Skip to main content

Full text of "Enzyme and metabolic inhibitors"

See other formats


^ 


Enzyme 
and  Metabolic 
Inhibitors 

Volume  II 

Malonate,  A  nalogs,  Dehydroacetate, 

Sulfhydryl  Reagents,  o-Iodosobenzoate,  Mercurials 


Volume  I         General  Principles  of  Enzyme  Inhibition 

Volume  III      lodoacetate 
Maleate 

A^-Ethylmaleimide 
Alloxan 
Ouinones 
Arsenicals 


Volume  IV 


in  preparation 


Uncouplers  of  Oxidative  Phosphorylation 

Dinitrophenol 

Arsenate 

Cyanide 

Carbon  Monoxide 

Azide 

Sulfide 

Antimycin 

Fluoroacetate 

Parapyruvate 

Urethanes 

Chelating  Agents 


Volume  V 


in  preparation 


Protein  Group  Reagents 

Heavy  Metals  (Copper,  Zinc,  Cadmium,  Silver,  etc.) 

Dyes 

Aldehydes 

Dimercaprol 

Antienzymes 

Phlorizin 

Selenite  and  Tellurite 

Naturally  Occurring  Inhibitors 

Cholinesterase  Inhibitors 

Monoamine  Oxidase  Inhibitors 

Drugs  as  Inhibitors 

Carbonic  Anhydrase  Inhibitors 

Borate 


Enzyme 
and  Metabolic 
Inhibitors 


Volume  II 

Malonate,  Analogs,  Dehydroacetate, 

Sulfhydryl  Reagents,  o-lodosobenzoate.  Mercurials 


J.  LEYDEN  WEBB 

School  of  Medicine 

University  of  Southern  California 

Los  Angeles,  California 


1966 


ACADEMIC  PRESS     New  York  and  London 


Copyright  ©  1966,  by  Academic  Press  Inc. 
all  rights  reserved. 

no  part  of  this  book  may  be  reproduced  in  any  form, 
by  photostat,  microfilm,  or  any  other  means,  without 
writt'^n  permission  from  the  publishers. 


ACADEMIC  PRESS  INC. 

Ill   Fifth  Avenue,  New  York,  .s'ew  York  10003 


United  Kingdom  Edition  published  by 
ACADEMIC  PRESS  INC.  (LONDON)  LTD. 
Berkeley  Square  House,  London  W.l 


Library  of  Congress  Catalog  Card  Number:  62-13126 


PRINTED   IN   THE   UNITED   STATES   OF  AMERICA 


This  volume  is  dedicated  with  sincere  gratitude  to  the  medical  librarians 
—  Vilma,  Ruth,  Lilian,  Clara,  Esther,  Michelle,  Rose,  Shari,  Nancy,  Nahida, 
and  others  —  who  have  not  only  helped  me  to  the  limit,  hut  have  made  each 
visit  to  the  library  a  pleasure  and  often  the  most  delightful  experience  of 
the  day. 


PREFACE 


For  those  rare  readers  who  may  feel  inclined  to  pursue  their  way  through 
Volumes  II  and  III  from  beginning  to  end,  I  have  tried  to  arrange  the 
chapters  and  sections  in  a  logical  and  interdependent  order.  Malonate  has 
been  approached  first  because  its  actions  so  well  illustrate  some  of  the 
general  principles  covered  in  Volume  I,  and,  indeed,  malonate  is  discussed 
in  greater  detail  than  any  other  inhibitor  in  order  to  suggest  how  one 
would  like  to  deal  with  all  inhibitors  if  one  had  either  the  time  or  space. 
Inasmuch  as  malonate  is  the  classic  substrate  analog,  the  next  chapter 
takes  up  various  types  of  analogs  and  here  we  are  able  to  obtain  some 
rough  idea  of  the  energies  involved  in  the  interactions  of  inhibitors  with 
enzyme  surfaces,  as  well  as  study  some  of  the  factors  which  determine 
specificity.  Some  readers  may  feel  that  too  much  attention  has  been  given 
to  these  analogs,  but  I  believe  they  represent  a  very  important  group  of 
inhibitors  and  illustrate  many  principles  —  competitive  behavior,  group 
specific  interactions,  protection  and  reversal,  and  even  nuituil  depk  Lion 
kinetics  since  some  analogs  are  extremely  potent  inhibitors  —  and,  in 
addition,  contribute  to  our  understanding  of  feedback  inhibition  and  me- 
tabolic regulation.  Most  of  the  remainder  of  the  volumes  is  devoted  to  sub- 
stances considered  to  react  with  SH  groups,  certainly  one  of  the  most 
commonly  used  and  important  classes  of  inhibitors,  about  which  it  is  sur- 
prisingly difficult  to  find  adequate  and  comprehensive  treatment.  Certain 
aspects  of  inhibition  have  been  treated  in  detail,  not  necessarily  because 
of  any  intrinsic  importance,  but  because  of  the  information  which  is 
provided  to  help  us  comprehend  the  general  phenomena  of  inhibition. 
There  are  many  ways  of  writing  about  inhibitors,  and  I  have  tried  to  alter 
the  approach  according  to  what  I  believe  to  be  the  most  interesting  aspects 
of  each  inhibitor.  These  aspects  may  not  happen  to  be  those  which  would 
have  been  chosen  by  the  reader,  but  it  is  impossible  to  cover  any  inhibitor 
completely  and  present  it  from  all  viewpoints.  On  the  other  hand,  there 
are  certain  sections  which  I  have  been  unable  to  make  very  interesting, 
to  organize  into  a  coherent  picture,  sometimes  because  the  data  are  insuf- 
ficient or  too  heterogeneous,  but  nevertheless  some  worthwhile  material 


vu 


Vlll  PREFACE 

can  often  be  included  in  these  areas.  One  finds  much  of  the  subject  to  be 
somewhat  disconnected  and  it  is  seldom  possible  to  present  an  orderly- 
clear  version  of  any  inhibition  because  of  the  gaps  in  our  knowledge,  but 
one  must  consider  that  these  isolated  strands  may  some  day  be  woven 
into  a  durable  fabric.  Each  chapter  has  in  general  been  organized  so  that 
the  treatment  proceeds  from  the  simplest  system  to  the  higher  levels  of 
organization,  since  this  generates  progressive  understanding  it  may  be 
hoped,  although  one  occasionally  wishes  that  the  effects  on  the  simpler 
systems  could  be  appreciated  against  a  background  of  the  actions  on  tissues 
and  animals.  Perhaps  to  some  extent  the  historical  introductions  at  the 
beginning  of  most  chapters  may  serve  as  provisional  backgrounds.  As  in 
many  fields  of  science  one  is  confronted  with  the  problem  of  vertical  or 
horizontal  presentations.  Efforts,  however  inadequate,  have  been  made  to 
correlate  the  results  at  the  different  levels,  and  it  is  hoped  that  unlike 
bacteria  under  certain  conditions  this  volume  does  not  too  much  exhibit 
the  phenomenon  of  accumulation  without  synthesis,  or  suffer  from  an  even 
worse  danger,  that  in  the  psychosynthesis  of  concepts  and  over-all  pictures 
some  abnormal  or  spurious  units  have  been  lethally  incorporated. 

This  periplus  of  the  field  of  enzyme  inhibition  presents  a  rather  large 
and  often  heterogeneous  group  of  information,  but  everything  has  been 
selected  for  some  reason;  the  reasons  may  be  debatable,  since  different 
readers  come  to  a  book  for  different  purposes,  but  occasionally  one  detects 
something  in  a  report,  perhaps  intuitively,  which  others  would  not,  and 
hence  includes  it  for  reasons  difficult  to  express.  The  half-life  for  the  general 
use  of  a  book  is,  indeed,  determined  in  part  by  the  ability  or  good  fortune 
of  the  author  to  select  that  which  will  have  the  most  value  or  pertinence 
in  the  pseudopodal  fronts  of  science.  One  has  no  time  for  justifications, 
since  some  decisions  have  to  be  made,  and  only  the  naive  think  they  can 
please  or  help  everyone,  but  there  is  perhaps  one  justification  I  feel  impelled 
to  make.  Certainly  there  will  be  those  who  ask  why  the  effects  of  an  inhib- 
itor on  the  blood  pressure  or  the  central  nervous  system  have  been  pre- 
sented when  there  is  little  or  no  obvious  correlation  with  any  metabolic 
inhibition,  or  why  I  have  made  up  tables  of  tolerated  or  lethal  doses,  and, 
in  general,  some  may  criticize  the  discussion  of  inhibitor  actions  which  are 
likely  to  be  unrelated  to  enzyme  inhibition,  or  at  least  for  which  there  is 
no  direct  evidence.  In  defense  of  this,  I  can  only  say  that  I  believe  we 
should  not  so  rigorously  categorize  the  actions  of  inhibitors.  The  refusal 
to  consider  the  nonmetabolic  actions  has  led  many  investigators  to  very 
biased  interpretations  of  their  data.  If  we  are  interested  in  the  mercurials, 
we  are,  I  assume,  interested  in  all  their  possible  actions,  whether  they  are 


PREFACE  IX 

based  on  metabolic  disturbances  or  not.  To  be  narrow  here  would  be  like 
discussing  only  the  beneficial  effects  of  drugs  and  omitting  the  toxic  actions. 
Of  course,  space  limitations  make  it  impossible  to  treat  all  these  actions 
equally,  and  I  have  tried  to  emphasize  those  actions  in  which  a  disturbance 
of  metabolism  is  the  most  likely  mechanism.  But  we  must  never  ignore  the 
possibilities  of  other  mechanisms  with  any  inhibitor,  particularly  those 
reacting  with  groups  on  proteins  and  other  cell  components.  The  mercurials 
offer  an  especially  clear  example  of  enzyme  inhibitors  producing  character- 
istic effects  on  many  tissues  (e.g.  kidney,  heart,  central  nervous  system, 
liver,  muscle,  etc.)  and  where  not  a  single  action  can  be  definitely  corre- 
lated with  a  mechanism  involving  enzyme  inhibition.  Nevertheless,  with 
further  improvements  in  techniques  and  more  knowledge,  it  is  quite  pos- 
sible that  in  the  future  at  least  some  of  these  actions  will  be  related  to 
effects  on  enzymes.  To  be  perfectly  honest,  at  the  present  time  we  cannot 
say  in  the  majority  of  cases  just  how  substances  called  enzyme  inhibitors 
act  to  produce  their  interesting  and  often  clinically  or  industrially  important 
effects  on  microorganisms  or  tissues,  and  it  is  necessary  to  realize  our  igno- 
rance so  that  progress  in  understanding  may  take  place.  Inhibitors  do 
produce  some  very  intriguing  effects  on  tissue  function  or  in  whole  animals, 
and  many  of  these  effects  are  unknown  to  those  who  look  upon  inhibitors 
merely  as  biochemical  tools;  so  by  reading  of  these  effects  some  may  be 
activated  to  study  the  ultimate  causes  in  greater  detail.  Incidentally, 
since  inhibitors  will  be  used  more  and  more  frequently  in  animals,  infor- 
mation on  dosage  ranges  to  produce  various  effects  may  serve  a  very  prac- 
tical purpose. 

I  would  like  to  express  my  gratitude  to  those  who  have  written  to  me 
saying  they  have  found  the  first  volume  of  interest  or  of  some  value  to 
them,  who  have  sent  me  unpublished  manuscripts  or  difficultly  obtainable 
material,  and  who  have  given  me  encouragement  during  those  periods 
when  I  sincerely  wished  I  were  in  a  monastery  in  Kyoto. 

J.  Leyden  Webb 
November,  1965 


CONTENTS 


Dedication v 

Preface vii 

Introduction xv 

Conventions xix 

Symbols xix 

CHAPTER  1 

Malonate  1 

Early  Historical  Development 1 

Chemical  Properties 3 

Inhibition  of  Succinate  Dehydrogenase 15 

Inhibition  of  Succinate  Oxidation  in  Cellular  Preparations        ....  50 

Inhibitions  of  Enzymes  Other  Than  Succinate  Dehydrogenase       ...  58 

Effects  of  Malonate  on  the  Operation  of  the  Tricarboxylic  Acid  Cycle      .  69 

Accumulation  of  Succinate  during  Malonate  Inhibition 90 

Accumulation  of  Cycle  Substrates  Other  than  Succinate 104 

Antagonism  of  Malonate  Inhibition  with  Fumarate 112 

Specificity  of  Malonate  Inhibition  in  the  Cycle 117 

Effects  of  Malonate  on  Oxidative  Phosphorylation 118 

Effects  of  Malonate  on  Glucose  Metabolism 122 

Effects  of  Malonate  on  Lipid  Metabolism 135 

Effects  of  Malonate  on  Amino  Acid  and  Protein  Metabolism     ....  151 

Effects  of  Malonate  on  Porphyrin  Synthesis 158 

Effects  of  Malonate  on  Miscellaneous  Metabolic  Pathways 163 

Effects  of  Malonate  on  the  Endogenous  Respiration 166 

Permeability  of  Cells  to  Malonate 186 

Growth,  Development,  and  Differentiation 192 

Cellular  and  Tissue  Function 202 

Effects  of  Malonate  in  the  Whole  Animal         217 

Effects  of  Malonate  on  Bacterial  Infections 221 

Metabolism  of  Malonate 224 

Inhibitors  Structurally  Related  to  Malonate 235 


XI 


xii  CONTENTS 

CHAPTER  2 

Analogs  of  Enzyme  Reaction  Components  245 

Terminology 246 

Possible  Sites  and  Mechanisms  of  Inhibition 246 

Kinetics  of  Analog  Inhibition 248 

Means  of  Expressing  Results 252 

Important  Types  of  Molecular  Alteration  Producing  Inhibiting  Analogs  .  255 

Development  of  the  Concept  of  Inhibition  by  Analogs 259 

Analog  Inhibition  of  Membrane  Transport 261 

Analogs  Which  Are  Isomers  of  Substrates 268 

Fumarase 274 

Inhibition  of  Xanthine  Oxidase  by  Purine  Analogs  and  Pteridines     .      .  279 

Choline  Oxidase 290 

Inhibition  of  Nitrogen  Fixation  by  Other  Gases         291 

Phenol  Oxidases 296 

Tyrosine  Metabolism 302 

Tryptophan  Metabolism 321 

Glutamate  Metabolism 327 

Arginase 335 

L-Amino  Acid  Oxidases 338 

D-Amino  Acid  Oxidase 340 

Analog  Inhibition  of  the  Metabolism  of  Various  Amino  Acids     ....  350 

Diamine  Oxidase  (Histaminase) 360 

Carboxypeptidase,  Aminopeptidases,  and  Dipeptidases 365 

Chymotrypsin  and  Other  Proteolytic  Enzymes 368 

Hexokinases 376 

Effects  of  2-Deoxy-D-Glucose  on  Carbohydrate  Metabolism        ....  386 

Effects  of  6-Deoxy-6-Fluoro-D-Glucose  on  Metabolism 403 

Various  Analog  Inhibitors  of  Carbohydrate  Metabolism 405 

Glycosidases 415 

Pyruvate  Metabolism         429 

Lactate  Metabolism 432 

Phosphatases 439 

Sulfatases 443 

Adenosinetriphosphatases  and  Transphosphorylases 444 

Hydroxysteroid  Dehydrogenases 447 

Nitrite  and  Sulfite  Metabolism 450 

Simple  Ion  Antagonisms 452 

Inhibition  by  Macroions 453 

Inhibitions  by  Nucleotides  and  Related  Substances 465 

Inhibitions  by  Coenzyme  Analogs 482 

Analogs  of  Nicotinamide  and  the  Pyridine  Nucleotides         484 

Analogs  of  Thiamine 514 

Analogs  of  Riboflavin  and  FAD 534 

Analogs  of  Pyridoxal 561 

Analogs  of  Pteroylglutamate  (Folate) 579 


CONTENTS  Xni 

Analogs  of  Other  Vitamins,  Coenzymes,  and  Their  Components     .      .      .  586 

Miscellaneous  Analog  Inhibitions 590 

CHAPTER  3 

Dehydroacetate  617 

Chemical  Properties 618 

Inhibition  of  Enzymes 620 

Effects  on  Respiration  and  Glycolysis 623 

Effects  on  Tissue  Functions 624 

Effects  on  the   Whole  Animal 627 

Distribution  and  Metabolism 629 

Antimicrobial    Activity 631 

CHAPTER  4 

Sulfhydryl  Reagents  635 

Role  of  SH  Groups  in  Metabolism  and  Function 636 

Chemical  Properties  of  SH  Groups         637 

Types  of  SH  Reaction  Important  in  Inhibition 642 

Factors  Determining  the  Reactivities  of  SH  Groups 643 

Interpretation  of  Inhibitions  by  SH  Reagents 647 

Protection  and  Inhibition  Reversal  by  Thiols 650 

General  Considerations  of  the  Uses  of  SH  Reagents 651 

CHAPTER  5 

Oxidants  655 

Disulfides 661 

Porphyrexide  and  Porphyrindin 664 

Ferricyanide 670 

Iodine 678 

Peroxides 690 

Tetrathionate 696 

CHAPTER  6 

o-lodosobenzoate  701 

Chemistry 701 

Reaction  with  Protein  SH   Groups 703 

Inhibition  of  Enzymes      . 704 

Inhibition  of  Metabolism 721 

Effects  on  Animal  Tissue  Functions 723 

Effects  in  Whole  Animals 724 

Effects  on  Sea  Urchin  Egg  Development 726 

Effects  on  Bacteria  and  Viruses 727 


XIV  CONTENTS 

CHAPTER  7 

Mercurials  729 

Chemical  Properties 730 

Reactions  with  Proteins 751 

Inhibition  of  Enzymes 768 

Electron  Transport  and  Oxidative  Phosphorylation 870 

Fermentation  and  Glycolysis 874 

Tricarboxylate  Cycle          877 

Respiration         879 

Various  Metabolic  Pathways 886 

The  Cell  Membrane  as  a  Site  for  Mercurial  Action 892 

Effects  on  Permeability  and  Active  Transport 907 

Effects  on  the  Kidney 917 

Effects  on  Tissue  Functions 937 

Effects  Observed  in  the  Whole  Animal 950 

Effects  on  Mitosis,  Growth,  and  Differentiation 963 

Effects  on  the  Growth  of  Microorganisms         970 

Development  of  Resistance  to  Mercurials 983 

References        987 

Author  Index    1071 

Subject  Index         1125 


INTRODUCTION 


Certain  principles  or  prejudices  in  the  approach  should  be  clearly  stated 
since  these  occasionally  heretical  opinions  have  been  the  basis  for  much  of 
the  organization  of  this  volume. 

What  is  hoped  to  be  pertinent  information  on  the  physical  and  chemical 
properties  of  the  inhibitors  has  been  given  in  the  belief  that  the  proper  use 
of  any  inhibitor  requires  as  much  knowledge  of  its  properties  as  possible. 
Such  data  are  often  difficult  to  find  and  I  am  afraid  that  much  important 
material  has  been  omitted. 

The  quantitative  formulation  of  inhibitions  has  been  stressed  because 
here,  as  in  every  science,  progress  often  depends  on  accurate  recording  and 
reporting  of  observations.  It  is,  for  example,  not  only  not  informative  but 
actually  misleading  to  state  that  aldehyde  oxidase  is  or  is  not  inhibited 
by  p-mercuribenzoate.  What  is  meant  by  "  inhibited  "  —  10%,  50%,  or 
100%?  What  is  the  concentration  of  the  inhibitor?  If  it  is  0.01  mM  it  may 
mean  something,  but  if  it  is  10  mM  probably  nothing.  What  is  the  source 
of  the  enzyme?  There  are  many  aldehyde  oxidases  and  they  differ  quite 
markedly  according  to  their  sources.  What  substrate  and  acceptor  were 
used?  Is  the  substrate  acetaldehyde,  glyceraldehyde,  formaldehyde,  reti- 
nene,  or  even  hypoxanthine,  and  are  electron  acceptor  dyes  used  or  the  Og 
uptake  measured?  These  and  other  factors  must  be  made  explicit.  Of 
course,  it  is  impossible  in  a  book  like  this  to  give  a  complete  picture  of 
each  inhibition  mentioned,  but  I  have  tried  to  state  the  source  of  the 
preparation,  the  substrate,  the  inhibitor  concentration,  and  the  per  cent 
inhibition  in  every  case,  as  well  as  the  pH  and  the  incubation  times  when 
necessary. 

One  can  understand  a  phenomenon  better  when  it  can  be  visualized  in 
some  manner  and  much  of  the  recent  work  on  the  inhibition  of  enzymes 
has  been  done  with  the  purpose  of  clarifying  the  topography  of  the  en- 
zyme surface  and  the  nature  of  the  interactions  occurring  there.  Thus  I 
have  tried  to  emphasize  the  interpretation  of  data  in  terms  of  an  accurate 
delineation  of  group  orientation  and  intermolecular  forces,  although  in  the 
present  state  of  our  knowledge  this  can  seldom  be  done  satisfactorily. 


XV 


XVI  INTRODUCTION 

Metabolism  within  cells  is  almost  always  a  matter  of  multienzyme  sys- 
tems and  so  the  effects  of  inhibitors  on  such  systems  have  been  discussed 
fully  wherever  possible,  although  this  is  even  more  difficult  to  describe 
quantitatively  than  the  behavior  of  single  enzymes. 

The  importance  of  the  specificity  of  inhibition  was  sufficiently  emphasiz- 
ed in  the  previous  volume  and  it  should  be  clear  that  this  is  a  critical  prob- 
lem which  has  been  neglected,  ignored,  or  abused  extensively.  It  is  not 
an  easy  matter  to  evaluate  the  specificity  of  an  inhibitor  under  various 
conditions,  particularly  when  the  necessary  data  are  lacking,  but  it  is 
hoped  that  at  least  a  provisional  picture  has  been  presented  in  some  in- 
stances. 

Certain  aspects  of  metabolism  (e.g.  glucose  utilization,  respiration,  pho- 
tosynthesis, protein  synthesis,  or  oxidative  phosphorylation)  and  cellular 
activity  (e.g.  active  transport,  membrane  potentials,  movement,  mitosis, 
or  proliferation)  are  obviously  of  general  significance,  and  the  effects  of 
inhibitors  on  these  have  been  emphasized.  This  is  not  to  say  that  other 
pathways  or  functions  are  unimportant,  and  indeed  where  necessary  they 
have  been  treated  as  adequately  as  possible,  but  one  cannot  discuss  all 
the  actions  of  each  inhibitor,  so  that  some  compromises  must  be  made. 

A  major  use  of  inhibitors  is  in  the  attempt  to  correlate  cellular  functions 
with  particular  enzymes  or  metabolic  pathways,  and  for  this  reason,  as 
well  as  the  fact  that  this  represents  one  of  the  most  fascinating  aspects  of 
inhibitor  study,  these  correlations  have  been  discussed  fully  if  the  infor- 
mation has  been  available,  and  the  effects  on  certain  organisms  or  processes 
have  often  been  given  in  the  hope  that  some  correlation  will  emerge  or 
further  work  will  be  stimulated.  It  is  believed  that  conceiving  inhibitor 
actions  in  terms  of  deviations  in  the  energy  flow  is  of  some  value  although 
an  accurate  formulation  of  this  must  await  the  development  of  a  new 
terminology. 

It  is  simpler  to  restrict  the  treatment  of  an  inhibitor's  action  to  a  par- 
ticular organism  or  tissue,  but  it  is  felt  that  a  great  deal  may  be  learned 
from  comparative  inhibitor  enzymology.  Therefore,  in  the  tables,  the  ef- 
fort has  been  made  to  present  the  results  from  as  many  sources  as  possible 
for  a  particular  enzyme  or  metabolic  pathway  since  by  doing  this  one  is 
better  able  to  see  the  great  extent  of  the  variability  in  responses;  only  a 
distorted  view  is  obtained  if  a  limited  range  of  action  is  considered. 

Paradoxical  actions  have  been  both  the  despair  and  delight  of  scientists 
in  many  fields,  and  it  is  recognized  that  some  of  our  finest  theories  have 
originated  in  the  observation  and  study  of  anomalies.  There  is  an  inherent 
desire  in  most  of  us  to  eliminate  anomalies  and  perhaps  devote  a  good  deal 


INTRODUCTION  XVU 

of  effort  to  this,  since  we  feel  that  an  anomaly  really  is  something  we  would 
expect  if  we  knew  the  system  or  mechanism  better,  or  as  Henry  Miller 
has  said  in  the  "  Tropic  of  Capricorn,"  "  confusion  is  a  word  we  have  in- 
vented for  an  order  which  is  not  understood."  I  have  thus  brought  up 
certain  so-called  anomalies,  not  only  for  their  interest  but  again  because 
they  often  stimulate  deeper  investigation,  although  at  present  they  may 
to  some  only  confuse  the  picture. 

Many  of  the  results  have  been  put  into  tabular  form,  first  because  this 
is  the  most  efficient  way  of  presenting  certain  types  of  data,  second  because 
such  simple  observations  are  often  the  sole  information  on  the  inhibitors 
provided  in  the  reports,  third  because  this  allows  a  more  convenient  com- 
parison of  results  (e.g.,  for  those  interested  in  possible  phylogenetic  rela- 
tionships, for  which  reason  the  source  organisms  have  usually  been  given 
in  the  classic  taxonomic  sequence,  or  for  studying  the  variability  in  re- 
sponses on  a  comparative  basis),  fourth  because  this  is  the  clearest  way 
to  provide  information  from  which  specificity  may  be  evaluated,  and  fifth 
because  these  tables  may  serve  as  reference  sources  for  those  interested  in 
the  actions  of  a  particular  inhibitor  on  a  certain  enzyme  or  organism. 
There  is  much  more  in  these  tables  than  anyone  can  assimilate  or  under- 
stand or  interpret  today,  but  it  is  these  data  which  could  possibly  con- 
tribute to  some  idea  or  concept  if  placed  against  the  proper  experience  or 
background.  Nothing  makes  some  data  look  more  miserable  or  incomplete 
than  putting  them  in  tables,  but  perhaps  this  is  an  asset,  since  it  shows 
what  is  missing,  what  should  have  been  done,  and  what  more  there  is  to 
do.  A  great  deal  of  information  could  not  be  included  in  the  tables,  for, 
although  some  of  them  look  formidably  long,  they  represent  only  a  frac- 
tion of  what  is  available  in  reports.  One  tries  to  include  only  that  which 
is  important,  but  the  definition  of  this  word  becomes  more  difficult  as  one 
applies  it.  There  are  so  many  very  specialized  and  unique  enzymes  being 
isolated  and  studied  these  days  that  it  becomes  more  of  a  problem  each 
year  to  determine  which  of  the  enzymes  are  generally  significant.  An  en- 
zyme which  at  first  sight  might  seem  esoteric,  if  for  no  other  reason  than 
its  gargantuan  name,  implying  a  specificity  of  catalysis  incommensurate 
with  anything  but  a  very  limited  role  in  metabolism,  may  well  be  of  great 
importance  in  a  particular  pathway,  a  pathway  perhaps  as  yet  undiscover- 
ed. Every  enzyme  is  of  some  importance  to  some  organism  or  tissue,  or  it 
would  not  be  there.  And  we  often  take  a  limited  viewpoint;  one  of  the 
numerous  enzymes  in  the  pathway  of  steroid  biosynthesis  is  recognized  as 
important  in  cholesterol  or  adrenal  corticoid  formation,  but  it  may  be  equal- 
ly important  to  some  microorganism  in  producing  steroids  which  function 


XVIU  INTRODUCTION 

in  their  membranes,  the  inhibition  of  the  formation  of  which  could  lead 
to  a  suppression  of  growth.  In  view  of  the  past  history  of  science,  anyone 
is  presumptuous  to  claim  they  can  distinguish  what  is  important  from  what 
is  not  —  we  have  to  do  this  much  of  the  time,  of  course,  but  we  should 
realize  we  are  presumptuous.  There  are  probably  some  errors  in  the  tables, 
since  it  is  often  difficult  to  determine  exactly  the  conditions  used;  one  is 
sometimes  referred  to  a  previous  report,  but  cannot  be  certain  that  all  the 
conditions  have  been  maintained  throughout  the  work.  One  must  often 
guess  a  parameter  from  other  work  the  investigators  have  done,  and  some- 
times calculate  results  from  heterogeneous  data.  There  has  been  a  good 
deal  of  calculation,  and  recalculation,  and  averaging,  and  I  take  full  re- 
sponsibility for  anything  right  or  wrong  I  may  have  done.  A  number  of 
curves  have  been  replotted  or  data  represented  in  a  way  that  differs  from 
that  of  the  original  investigator,  and  I  fully  realize  that  this  usuallj'  results 
in  nothing  but  animosity. 


CONVENTIONS 


The  naming  of  enzymes  is  not  an  easy  task.  On  the  one  hand,  there  are  the  more 
trivial  names  with  their  occasional  confusions  —  on  the  other,  there  are  the  official 
names  in  the  "  Report  of  the  Commission  on  Enzymes  "  (1961)  which  are  reasonably 
precise  but  often  unwieldy.  I  have  usually  chosen  the  former  because  I  feel  most 
readers  will  recognize  these  more  readily,  but  frequently  I  have  taken  an  interme- 
diate course  which  probably  will  not  please  anyone.  It  is  much  more  accurate  to 
write  NADH:menadione  oxidoreductase  than  to  use  the  designation  NADH  oxidase 
or  NADH  dehydrogenase,  since  the  former  name  indicates  the  substrate  and  acceptor 
used.  In  addition  it  is  cumbersome  to  use  D-xylulose-5-phosphate  D-glyceraldehyde- 
3-phosphate-lyase  (phosphate-acetylating)  instead  of  phosphoketolase,  yet  there  is  no 
doubt  that  this  longer  terra  accurately  describes  the  enzyme.  There  are  also  prefer- 
ences in  nomenclature,  for  various  reasons.  I  never  cared  much  for  the  term  invertase; 
I  prefer  to  call  it  [ii-fructofuranosidase,  although  it  is  clumsier,  but  not  as  much  so 
as  P-D-fructofuranoside  fructohydrolase.  In  other  instances  the  older  and  shorter 
names  are  more  pleasing  to  me  and  I  imagine  to  others.  I  have  tried  to  use  enzyme 
names  which,  at  least,  can  be  found  in  the  index  of  the  "  Report  of  the  Commission 
on  Enzymes,"  and  some  cross  referencing  of  names  has  been  included  in  the  index. 
There  are  certain  instances  of  inconsistency  which  I  do  not  particularly  regret. 

As  in  the  first  volume,  concentrations  have  been  given  as  millimolar  (m.M)  except 
when  designated  otherwise,  and  in  other  matters  the  conventions  given  there  have 
been  retained. 


SYMBOLS 


A 

absorbance 

DQ 

ADP 

adenosinediphosphate 

DQH, 

AMP 

adenosinemonophosphate 

Eo 

ATP 

adenosinetriphosphate 

9,10-AQ 

9,10-anthraquinone 

Eo' 

BAL 

dimercaprol 

ChE 

cholinesterase 

ED, 

CoA 

coenzyme  A 

6-DFG 

6-deoxy-6-fluoro-D-glucose 

EDTA 

2-DG 

2-deoxy-D-glucose 

EI 

DNA 

deoxyribonucleic  acid 

EM 

DNP 

2,4-dinitrophenol 

duroquinone 
durohydroquinone 
standard  oxidation-reduc- 
tion potential  (pH  =  0) 
oxidation-reduction  poten- 
tial at  specified  pH  (usually  7) 
effective  dose  or  concentra- 
tion for  X  per  cent 
ethylenediaminetetraacetate 
enzyme-inhibitor  complex 
Embden-Meyerhof  (path- 
way) 


XX 

SYMBOLS 

EP 

enzyme-product  complex 

NEM 

iV-ethylmaleimide 

Epi 

epinephrine 

1,2-NQ 

1 ,2-naphthoquinone 

ES 

enzyme-substrate  complex 

1,4-NQ 

1 ,4-naphthoquinone 

FAD 

flavin-adenine  dinucleotide 

pl 

-  log  (I) 

FDP 

fructose- 1 ,6-diphosphate 

p-MB 

p-mercuribenzoate  ion 

FMN 

flavin  mononucleotide 

p-MPS 

j9-mercuriphenylsulfonate 

GSH 

reduced  glutathione 

ion 

GSSG 

oxidized  glutathione 

P-Q 

2)-benzoquinone 

i 

fractional  inhibition 

P-QH, 

p-benzohydroquinone 

if 

final  fractional  inhibition 

(hydroquinone) 

lA 

iodoacetate 

pS 

-  log  (S) 

lAM 

iodoacetamide 

P-XQ 

p-xyloquinone 

IC 

intracutaneous 

9,10-PAQ 

9, 10-phenanthraquinone 

IM 

intramuscular 

3-PGDH 

3  -  phosphogly  ceraldehyde 

IMP 

inosinemonophosphate 

dehydrogenase 

IP 

intraperitoneal 

PM 

phenylmercuric  ion 

ISBZ 

o-iodosobenzoate 

Pyr 

pyruvate 

IV 

intravenous 

SC 

subcutaneous 

Ka 

ionization  constant 

SH 

sulfhydryl 

Ki 

inhibitor  constant 

S/M 

slice/medium  ratio 

Km 

Michaelis  constant 

s— s 

disulfide 

Ks 

substrate  constant 

TD 

tolerated  dose  or  concentra- 

LD. 

lethal  dose  or  concentration 

tion 

for  X  per  cent 

T/M 

tissue/medium  ratio 

MD 

menadione 

TQ 

toluquinone 

MHD2 

menadiol 

TQH3 

toluhydroquinone 

MLD 

minimal  lethal  dose 

V 

rate 

MM 

methylmercuric  ion 

Vm 

maximal  rate 

NAD 

nicotinamide-adenine 

P-XQ 

p-xyloquinone 

dinucleotide 

£ 

molar   extinction   coefficient 

NADP 

nicotinamide-adenine 

(p-AsO 

phenylarsenoxide 

dinucleotide  phosphate 

99-ASO2 

phenylarsonic  acid 

Enzyme 
and  Metabolic 
Inhibitors 

Volume  II 

Malonate,  Analogs,  Dehydroacetate, 

Suljhydryl  Reagents,  o-Iodosobenzoate,  Mercurials 


CHAPTER  1 

MALONATE 


Malonate  is  one  of  the  most  interesting,  specific,  useful,  and  well-known 
enzyme  inhibitors,  and  for  these  reasons  will  be  discussed  in  detail  in  order 
to  illustrate  some  of  the  general  principles  delineated  in  the  first  volume. 
It  will  be  valuable  perhaps  to  take  up  one  inhibitor  to  the  degree  necessary 
to  consider  many  of  the  various  problems  in  the  application  of  these  prin- 
ciples, and  to  examine  the  pitfalls  that  may  appear  even  in  the  use  of  an 
inhibitor  that  in  several  ways  approaches  what  one  would  want  ideally. 
Much  of  what  will  be  said  concerning  malonate  may  be  applied  to  the  other 
inhibitors.  This  discussion  will  emphasize  the  often  overlooked  fact  that 
the  effect  of  relatively  simple  inhibitors  in  cells  may  constitute  a  very 
complex  problem  and  that  their  use  in  elucidating  metabolic  relationships  in 
tissues  or  whole  organisms  should  not  be  undertaken  lightly.  It  is  a  simple 
matter  to  apply  an  inhibitor  such  as  malonate  but  it  is  often  very  difficult 
to  use  it  properly  and  to  interpret  the  results  accurately.  The  treatment 
of  malonate  will,  furthermore,  provide  a  foundation  for  the  more  general 
discussion  of  competitive  inhibitions  produced  by  analogs  in  the  following 
chapter. 

EARLY   HISTORICAL  DEVELOPMENT 

The  first  report  of  the  use  of  malonate  in  a  biological  system  was  made 
by  Heymans  (1889)  to  the  Physiological  Society  in  Berlin.  The  toxicity 
of  oxalate  to  animals  had  been  known  for  many  years  and  Heymans  be- 
lieved that  an  investigation  of  the  higher  homologs  of  the  dicarboxylate 
series  might  be  interesting.  Although  sodium  oxalate  was  quite  poisonous 
when  injected  into  the  frog  dorsal  lymph  sac,  the  sodium  salts  of  malonate, 
succinate,  and  glutarate  were  essentially  without  effect,  sodium  malonate 
being  nonlethal  at  a  dose  as  high  as  about  8  g/kg.  However,  Pohl  (1896), 
working  in  Prague,  found  that  the  urinary  excretion  of  oxalacetate  was 
increased  by  administering  malonate  to  dogs  and,  furthermore,  that  only 
a  small  portion  of  the  malonate  given  could  be  recovered  in  the  urine, 
indicating  that  the  dog  can  metabolize  malonate.  The  first  experiments 
showing  the  metabolic  inhibitory  action  of  malonate  were  done  by  Thun- 


2  1.    MALONATE 

berg  (1909)  in  Lund.  He  had  observed  the  inhibition  of  minced  frog  muscle 
respiration  by  oxalate  and  decided  to  study  the  higher  homologs.  Thus  the 
inhibitory  action  of  malonate  on  muscle  respiration  was  demonstrated, 
whereas  succinate  instead  stimulated  the  oxygen  uptake.  Apparently  this 
observation  went  unnoticed  and  the  inhibitory  activity  had  to  be  rediscov- 
ered later.  Rose  (1924)  at  Illinois  showed  that  malonate  exhibits  no  nephro- 
toxic action,  as  does  glutarate,  when  given  orally  to  rabbits,  although  he 
later  (Corley  and  Rose,  1926)  found  that  the  methyl  and  ethyl  derivatives 
depress  renal  function.  At  about  the  same  time,  Momose  (1925)  in  Japan, 
continuing  the  work  of  Pohl,  observed  that  malonate  when  perfused  through 
dog  liver,  gives  rise  to  acetoacetate,  acetone,  and  aldol.  He  postulated  that 
these  substances  arise  from  malonate  after  decarboxylation  to  acetate,  but 
it  is  more  likely  from  our  present  knowledge  that  malonate  gives  rise  to 
these  substances  by  a  disturbance  of  the  metabolism.  During  the  next 
few  years  evidence  was  accumulated  that  malonate  can  arise  from  normal 
tissue  metabolism  —  occurring  in  alfalfa  (Turner  and  Hartman,  1925) 
and  wheat  (Nelson  and  Hasselbring,  1931),  and  appearing  during  citrate 
fermentation  in  the  mold  Aspergillus  (Challenger  et  al.,  1927)  —  and  be 
metabolized  by  certain  microorganisms,  such  as  Escherichia  coli  (Grey, 
1924). 

Our  present  concepts  of  the  inhibitory  action  of  malonate,  however, 
arose  from  the  work  on  bacterial  dehydrogenations  by  Quastel  and  Whetham 
(1925)  at  Cambridge.  They  tested  the  abilities  of  various  dicarboxylic  acids 
to  reduce  methylene  blue  in  suspensions  of  E.  coli  and  found  that  only  suc- 
cinate is  active.  Malonate  inhibited  this  reduction  by  succinate.  As  stated 
in  their  own  words,  "Oxalic,  glutaric,  and  adipic  acids  (when  mixed  with 
succinic  acid)  do  not  retard  the  reduction  due  to  the  succinic  acid,  but 
malonic  acid  has  a  definite  retarding  effect.  It  is  difficult  to  explain  the 
anomalous  behaviour  of  malonic  acid,  but  there  is  no  doubt  as  to  the  reality 
of  the  effect."  They  found  the  methylene  blue  reduction  time  with  succinate 
bo  be  tripled  in  the  presence  of  77  mM  malonate.  Quastel  and  Wooldridge 
(1928)  extended  this  work  to  show  that  the  action  on  succinate  oxidation  is 
rather  specific  in  that  malonate  does  not  appreciably  inhibit  the  oxidation 
of  several  other  substrates  by  E.  coli.  But  in  addition  they  demonstrated 
that  increasing  succinate  concentrations  would  counteract  the  malonate 
inhibition,  leading  them  to  suggest  that  both  substances  are  adsorbed  to 
the  enzyme  reversibly,  probably  competing  for  the  same  active  site.  Final- 
ly, Quastel  and  Wheatley  (1931),  now  at  the  Cardiff  City  Mental  Hospital, 
reported  that  the  malonate  inhibition  of  succinate  oxidation  occurs  in 
many  bacteria,  and  in  mammalian  brain  and  muscle  as  well,  the  enzymes 
from  the  mammalian  tissue  being  even  more  sensitive. 

The  concept  of  the  competitive  inhibition  of  an  enzyme  by  a  substance 
structurally  related  to  the  normal  substrate  was  first  clearly  demonstrated 


CHEMICAL   PROPERTIES  6 

and  expressed  for  this  inhibition  of  succinate  oxidation  by  malonate  in  the 
work  of  Quastel  and  Wooldridge,  although  competition  specifically  was  not 
mentioned.  Cook  (1930),  also  at  Cambridge,  however,  stated  that  a  "com- 
petitive" mechanism  had  been  established,  presumably  referring  to  the 
work  of  Quastel  inasmuch  as  Cook  performed  no  experiments  indicating 
a  competitive  relationship.  The  competitive  nature  of  the  malonate  inhibi- 
tion has  been  substantiated  many  times  and  placed  on  a  quantitative  basis, 
so  that  malonate  has  come  to  be  recognized  as  the  classical  example  of 
inhibition  by  a  pvirely  competitive  mechanism.  The  development  and  ap- 
plications of  this  concept  will  be  discussed  in  more  detail  in  Chapter  2. 
For  20  years  malonate  was  the  only  available  specific  inhibitor  of  succinate 
dehydrogenase,  and  later  of  the  tricarboxylic  acid  cycle,  and  actually  played 
an  important  role  in  the  elucidation  of  the  cycle  sequence.  Other  cycle  inhib- 
itors have  been  described  recently,  but  no  other  inhibitors  of  the  succinate 
oxidation  step  as  specific  and  useful  as  malonate  have  been  found. 

CHEMICAL    PROPERTIES 

Malonic  acid  and  its  salts  when  obtained  commercially  are  often  not 
sufficiently  pure  for  accurate  work  and  it  has  been  the  practice  in  our  lab- 
oratory to  recrystallize  all  material.  Malonic  acid  may  be  recrystallized 
from  ethyl  acetate  and  benzene  (Adell,  1940),  or  ether  and  benzene  con- 
taining 5%  light  petroleum  (Vogel,  1929),  or  simply  from  a  hot  concentrated 
benzene  solution  by  cooling  to  50-10°.  The  sodium  and  potassium  salts  may 
be  dissolved  in  small  amounts  of  warm  water  and  precipitated  by  the  ad- 
dition of  ethanol,  as  is  commonly  done  with  other  dicarboxylate  salts 
(Potter  and  Schneider,  1942),  or,  with  somewhat  less  yield,  may  be  crystal- 
lized by  cooling  hot  concentrated  aqueous  solutions.  In  all  cases  we  have 
decolorized  with  activated  charcoal  in  the  solutions  before  recrystallization 
and  have  washed  the  products  with  ether  preparatory  to  drying.  It  should 
be  emphasized  that  the  choice  of  the  sodium  or  the  potassium  salt  will 
depend  on  whether  the  preparation  to  be  tested  is  cellular  or  subcellular. 

Stability 

Malonic  acid  and  its  salts  are  quite  stable  and  chemically  unreactive. 
Decarboxylation  to  acetate  proceeds  very  slowly  under  ordinary  conditions. 
Aqueous  solutions  of  sodium  malonate  heated  to  125°  for  48  hr  show  no 
perceptible  decomposition  (Fairclough,  1938),  and  the  half-life  of  sodium 
hydrogen  malonate  in  solutions  5-50  mM  is  at  80°  around  40  days,  cal- 
culated from  the  rate  constant  for  decarboxylation  (Hall,  1949).  The  free 
energy  change  for  the  reaction 

Malonate   ->  acetate  +  CO, 


4  1.    MALONATE 

is  approximately  —7  kcal/mole  but  the  activation  energy  is  27.9  kcal/mole 
(Gelles,  1956).  Thus  at  physiological  temperatures  one  may  consider  mal- 
onate  as  completely  stable.  Nevertheless,  there  are  enzyme  systems  which 
catalyze  the  decarboxylation  (see  page  227)  and  one  should  be  reasonably 
certain  in  the  use  of  malonate  that  it  is  stable  in  the  system  investigated, 
since  the  formation  of  acetate  might  well  confuse  the  results. 

Molecular  Structure 

In  crystals  of  malonic  acid  the  molecules  are  arranged  in  zigzag  chains 
with  the  carboxyl  groups  linked  through  two  hydrogen  bonds  (Goedkoop 
and  MacGillavry,  1957).  The  following  bond  parameters  were  observed 
(the  two  values  refer  to  the  two  carboxyl  groups,  since  the  molecule  in  the 
crystal  is  apparently  not  symmetrical):  C — C — C  angle  =  110^;  C — C 
distance  =  1.54,  1.52;  C— 0  distance  =  1.29,  1.31;  C=0  distance  =  1.24, 
1.22;  0— C— 0  angle  =  128^,  128°;  and  H-bond  distance  =  2.68-2.71. 
The  malonate  ion  in  solution  would  probably  deviate  somewhat  from  this 
configuration  but  not  a  great  deal  inasmuch  as  malonate  is  fixed  in  a  rather 
rigid  structure,  because  the  C — C — C  angle  is  determined  by  the  electronic 
tetrahedral  orbitals  and  can  be  distorted  only  with  difficulty.  The  carbox- 
ylate  groups  can  rotate  around  the  C — C  axis  but  they  presumably  ster- 
ically  interfere  with  each  other  when  both  lie  in  the  plane  of  the  molecule, 
since  the  centers  of  the  oxygen  atoms  would  be  2.2  A  apart  and  the  van 
der  Waals'  radius  of  the  oxygen  atom  is  1.4  A  (Goedkoop  and  MacGillavry, 
1957).  In  malonic  acid  crystals  one  carboxyl  seems  to  be  in  the  molecular 
plane  and  the  other  is  at  right  angles;  in  the  malonate  ion  it  may  well  be 
that  neither  is  in  the  C — C — C  plane.  However,  another  factor  must  be 
considered;  it  is  possible  that  in  solution  there  is  intramolecular  hydrogen 
bonding  (Gelles,  1956),  at  least  for  the  hydrogen  malonate  ion.  When  the 
carboxyl  group  ionizes,  the  equivalence  of  the  structures: 

R-cf  R_c--° 

allows  a  greater  resonance  than  in  the  unionized  state  (equivalent  to  an 
extra  8  kcal/mole  energy),  and  this  high  resonance  would  indicate  an  inter- 
mediate structure  in  which  the  center  of  negative  charge  lies  midway  be- 
tween the  two  oxygen  atoms.  Although  keto-enol  tautomerism  occurs 
(Hofling  et  al.,  1952)  in  the  esters  of  malonic  acid: 

O  OH 

-CH^-C-OEt    ^     -CH=C-OEt 

it  is  probably  not  significant  in  the  malonate  ion  because  it  would  reduce  the 
electronic  resonance. 


CHEMICAL   PROPERTIES  5 

The  distance  between  the  two  centers  of  negative  charge  in  malonate 
is  of  importance  in  the  binding  to  succinate  dehydrogenase.  Calculation  of 
this  distance  (using  the  following  values:  C — C — C  angle  =  111.7°;  0 — C — 0 
angle  =  125.8°;  C— C  distance  =  1.544,  and  C— 0  distance  =  1.273) 
leads  to  a  value  of  3.28  A.  Intercharge  distances  for  various  dicarboxylates 
and  other  compounds  known  to  inhibit  succinate  dehydrogenase  are  given 
in  Table  1-1,  and  these  values  will  be  of  interest  in  comparisons  of  inhibi- 
tory activity.  Dicarboxylic  acids  with  more  than  one  methylene  group  are 
flexible  and  the  intercharge  distances  may  vary  between  the  limits  of  great- 
est bending  and  extension  of  the  molecules.  As  the  chain  length  increases 
there  will  be  greater  tendency  for  the  intercharge  distance  in  the  ions  to 
be  less  than  that  of  the  maximal  extension,  since  the  electrostatic  repulsion 
will  decrease.  The  mean  statistical  intercharge  distance  in  the  succinate  ion 
will  probably  be  closer  to  4.75  A  than  the  contracted  distance  of  3.81  A. 
Indeed,  it  may  be  calculated  that  it  would  require  at  least  2.3  kcal/mole 
to  bring  succinate  from  the  extended  to  the  contracted  configuration, 
using  the  dielectric  constant  obtained  from  D  =  6d  —  7  (Eq.  1-6-72)*; 
since  the  dielectric  constant  is  probably  less  due  to  the  hydrocarbon  groups 
between  the  charges,  this  would  be  a  minimal  energy  value.  The  mean 
intercharge  distance  in  the  succinate  ion  may  be  estimated  as  not  less  than 
4.20  A  (Gane  and  Ingold,  1931;  Eyring,  1932;  Westheimer  and  Shookhoff, 
1939)  and  possibly  closer  to  4.75  A.  The  glutarate  intercharge  distance  is 
probably  around  5.2  A.  These  considerations  are  of  importance  in  comparing 
the  interactions  of  these  substances  with  the  active  center  of  succinate 
dehydrogenase.  It  must  be  remembered  that  the  bound  ions  are  undoubt- 
edly held  in  a  configuration  different  from  the  statistical  mean  in  solution. 
The  flexibility  of  the  higher  homologs  allows  them  to  adjust  to  a  specified 
configuration,  but  at  the  expense  of  the  energy  and  entropy  changes  neces- 
sary to  bring  them  from  their  free  configurations. 

Acidic  Ionization 

The  ionizations  of  malonic  and  succinic  acids  are  important  in  the  inter- 
actions with  succinate  dehydrogenase  and  with  regard  to  the  penetration  of 
these  substances  into  cells.  The  pKJs  for  the  simple  dicarboxylic  acids,  in- 
cluding various  derivatives  of  succinate  and  malonate,  and  related  succinic 
dehydrogenase  inhibitors,  are  given  in  Table  1-2.  The  dissociation  constants 
change  slightly  with  ionic  strength;  for  malonic  acid,  dpK^  Ids  =  —  0.32, 
and  dpK^  jds  =  —  0.98,  for  ionic  strengths   around  0.15,  where  s  is  the 

*  In  cross  references  of  this  type,  Eq.  1-6-72,  the  roman  number  indicates  the  volume 
of  this  treatise  in  which  the  equation,  table,  or  figure  (as  the  case  may  be)  may  be 
found,  the  first  arabic  number  indicates  the  chapter  number,  and  the  second  arabic 
jiumber  the  equation,  table,  or  figure  number. 


1.    MALONATE 


Table  1-1 
Intekchakge  Distances  in  Dianions  op  Interest  " 


Compound 


Oxalate 
Malonate 


Succinate 


Glutarate 


(c) 


Adipate 


Hydroxy  malonate 
(tartronate) 


Fumarate 

Maleate 

Acetylene-dicar  boxy  late 
o-Phthalate 

Isophthalate 
Terephthalate 


Structure 


'ooc-coo 

'ooc      coo" 

\   / 

CHj 
"ooc  -CHj 


"OOC 


CH,-  COO" 
COO" 


\  / 

CH2 CH2 


"OOC— CHj 


CH2  CH2 

COO" 


"OOC-CHj 

"ooc    coo" 

/       \ 

CH9      CHa 

^  / 

CHj 
"OOC-CHj 

CHj— CH. 

\ 


COO 


CHj-  COO" 


"OOC         COO 

\  / 

CH 

I 
OH 

"OOC  H 

\  / 

c=c 

/        \      - 

H  COO 

"OOC  COO" 

\       / 
c  =  c 

/       \ 

H  H 

"OOC-CEC-COO" 

COO" 


^^^^-^  COO- 


COO" 


Distance  (A) 


2.42 
3.28 

4.75 

3.01 

5.83 

4.84 

1.70 

6.87 
3.28 

4.87 

3.66 
5.16 
3.55 

5.84 
7.33 


COO 


CHEMICAL   PROPERTIES 


Table  1-1  (continued) 


Compound 


Structure 


Distance  (A) 


Cyclobutane-dicar  boxy  late 


Cyclopentane-dicar  boxy  late 


COO 


coo 

COO' 


Cyclohexane-dicarboxylate 


COO' 
COO' 


COO 


/3-Sulfopropionate         < 

(a) 
(b) 

OOC-CH, 
\ 
CHj-SO. 

"OOC                   SO,' 

CHj— CHj 

Methanedisulfonate 
(methionate) 

'O3S         SO3' 
CH, 

1,2-  Ethanedisulf  onate  < 

(a) 
(b) 

'OjS-CHj 

CHj— SO3' 

'0,8                    SO,' 
'  \               /      ' 
CHj—  CHj 

/::?X^COO" 

0  -Sulfobenzoate 

^SO,' 

Arsonoacetate 

'OOC         ASO3H' 
CH, 

^-Phosphonopropionate  < 


(a) 


(b) 


'OOC-CHj 


"OOC 


CH,— PO; 

po: 


\  / 

CH2       CH2 


3.39 


cis- 

3.52 

trans- 

4.91 

cis  - 

3.05 

trans  - 

4.51 

5.05 

3.13 

3.40 

5.38 

3.25 

3.72 

4.05 
5.09 
3.14 


The  distances  were  calculated  on  the  basis  of  bond  lengths  and  angles  given  In  Tables  1-6-12  and  1-6-13 
except  where  direct  measurements  were  available.  It  was  assumed  that  the  center  of  negative  charge  lies 
midway  between  the  resonating  oxygen  atoms.  For  o-phthalate  and  cyclohexane-dicarboxylate  a  3°  distortion 
of  the  bond  angles  due  to  electrostatic  repulsion  was  assumed.  For  cyclopentane-dicarboxylate  a  further 
5°  widening  was  estimated  from  the  bending  of  the  ring  angles.  The  value  for  cyclobutane-dicarboxylate 
is  only  approximate  and  is  based  on  a  7°  distortion  of  the  bond  angle  in  comparison  to  malonate.  It  should 
be  pointed  out  that  the  values  in  this  table  are  smaller  than  generally  used;  one  reason  is  that  Intercarboxy- 
late  distances  have  usually  been  based  on  the  distances  between  associating  or  dissociating  protons ,  ra- 
ther than  between  centers  of  negative  charge. 

ionic  strength.  Since  AH  for  the  dissociation  of  weak  acids  is  quite  small, 
the  constants  do  not  change  much  with  temperature;  d^K^JdT  =  0.0031, 
and  d])Kg  jdT  =  0.0038,  approximately.  Three  species  will  be  present  in 
any  solution  of  malonate  —  HOOC— CHg— COOH,  HOOC— CHg— C00-, 
and  -OOC — CHg — COO"  —  the  relative  concentrations  being  determined 


1.    MALONATE 

Table  1-2 
Ionization  Constants  of  Dicarboxylic  Acids  " 


Acid 


pKa, 


P^., 


Oxalic 

Malonic 

Succinic 

Glutaric 

Adipic 

Pimelic 

Methylmalonic 

Ethylmalonic 

Dimethylmalonic 

Diethylmalonic 

n.-Propylmalonic 

i«o-Propylmalonic 

Methylethylmalonic 

Di-w-propylraalonic 

Phenylmalonic 

Hydroxymalonic  (tartronic) 

Maleic 

Methylmaleic  (citraconic) 

Fumaric 

Methylfumaric  (mesaconic) 

Malic 

Tartaric 

2,2  '-Dimethylsuccinic 

2,2  '-Diethylsuccinic 

Tetramethylsuccinic 

Methylenesuccinic  (itaconic) 

Cyclopropane- 1 , 1  -dicarboxylic 
Cyclobutane- 1 , 1  -dicarboxylic 
Cyclopentane- 1 , 1  -dicarboxylic 
Cyclohexane- 1 , 1  -dicarboxylic 
<raw5-Cyclopropanedicarboxylic 
cis-Cyclopropanedicarboxylic 
frarw-Cyclopentanedicarboxylic 
<raws-Cyclohexanedicarboxylic 

Phthalic 
Isophthalic 


1.09 

3.79 

2.58 

5.17 

3.95 

5.16 

4.07 

4.93 

4.17 

4.95 

4.23 

4.98 

2.86 

5.24 

2.79 

5.34 

2.97 

5.59 

2.04 

6.90 

2.83 

5.38 

2.78 

5.46 

2.71 

6.07 

1.92 

7.13 

2.43 

4.68 

2.93 

— 

1.67 

5.75 

2.23 

5.89 

2.85 

4.00 

2.93 

4.82 

3.21 

4.62 

2.83 

3.88 

3.77 

5.82 

3.34 

6.22 

3.33 

6.90 

3.67 

5.19 

1.77 

7.37 

3.08 

5.82 

3.18 

6.02 

3.40 

6.05 

3.49 

4.75 

3.16 

6.09 

— 

4.22 

— 

4.45 

2.95 

5.23 

2.15 

4.49 

"  These  values  have  been  obtained  from  a  variety  of  sources  and  have  been  corrected 
to  a  temperature  of  37°  and  an  ionic  strength  of  0.15  so  as  to  be  applicable  to  physio- 
logical conditions.  These  corrections  were  obtained  from  studies  on  the  variations  of 
dicarboxylic  ionization  constants  with  temperature  and  ionic  strength  (e.g.  AdeU, 
1940).  They  are  not  absolutely  correct  but  probably  allow  a  closer  approximation 
than  the  values  in  the  literature,  which  are  usually  for  25°  and  extrapolated  to  zero 
ionic  strength. 


CHEMICAL   PROPERTIES 


by  the  pH  (see  Eqs.  1-14-6  to  1-14-8).  The  variations  of  these  species  with 
pH  are  shown  in  Fig.  1-1,  and  actual  concentrations  are  given  in  Table  1-3 
for  malonate  and  succinate  at  a  total  concentration  of  10  mM.  In  the 
usual  range  of  physiological  pH,  over  95%  of  these  acids  are  in  the  form 
of  the  completely  dissociated  doubly-charged  anion;  this  is  the  active  form 
for  the  inhibition  of  succinate  dehydrogenase.  However,  it  is  the  concen- 
trations of  the  other  forms  which  are  important  in  the  rates  and  degrees 
of  penetration  into  cells,  and  these  change  appreciably  with  pH  (e.  g., 
from  pH  7.4  to  6.8  there  is  a  4-fold  increase  in  the  singly  dissociated  form 
and  a  16-fold  increase  in  the  undissociated  form). 


"x 

/^ 

,cooh\ 

^COOH      \ 

^< 

_  cooA 

' COOH  \ 

/           COO' 

/         ^coo- 

;    ^ 

I 

'                        Malonate 

J 

\ 

V 

pH 

Fig.  1-1.  Concentrations  of  the  various  ionic  species  of  malonic 

acid  at  different  values  of  the  pH  expressed  as  per  cent  as  of 

the  total  concentration. 


Das  and  Ives  (1961)  on  the  basis  of  thermodynamic  evidence  suggested 
that  an  internal  symmetrical  hydrogen  bond  occurs  in  H-malonate~,  and 
that  this  would  affect  to  some  extent  the  piiC^  values  and  reduce  hydration. 
However,  Lloyd  and  Prince  (1961)  examined  the  infrared  spectra  of  malonic 
acid  and  its  ions  in  DgO,  compared  the  data  with  those  obtained  with 
fumaric  acid  (in  which  no  hydrogen  bonds  could  occur),  and  concluded  that 
if  hydrogen  bonding  exists  in  H-malonate~,  the  bond  is  very  weak  and  not 
symmetrical.  Eberson  and  Wadsd  (1963)  determined  the  ionization  enthal- 
pies in  water  and  ethanol,  and  also  concluded  that  hydrogen  bonding  is 
not  important  in  the  dicarboxylates  when  zlpiiC^  is  less  than  4.  It  would 
thus  appear  that  intramolecular  hydrogen  bonding  is  not  a  significant 
factor  in  stabilizing  the  H-malonate~  ion. 


10 


1.    MALONATE 


> 

ft 

E-t 


<  ^ 


d      d 


o      o 


-H  05  -< 


^       -^       oo 


ft 


CHEMICAL   PROPERTIES  11 

Metabolic  studies  of  the  substituted  malonates  and  malonic  esters  wiU  be 
taken  up  after  the  actions  of  malonate  have  been  discussed  (see  page  235). 
It  is  interesting  to  note  (Table  1-2)  that  the  pK^  's  of  the  substituted 
malonic  acids  are  generally  higher  than  for  malonic  acid  itself.  This  is 
mainly  the  result  of  the  reduction  of  the  dielectric  constant  of  the  region 
between  the  interacting  carboxyl  groups,  and  is  particularly  evident  for 
the  disubstituted  ethyl  and  w-propyl  derivatives.  This  increase  in  pK^ 
should  facilitate  penetration  of  these  compounds  into  cells;  in  addition  they 
are  more  lipid-soluble,  which  will  also  favor  penetration.  The  esters  are  not 
active  inhibitors  of  succinate  dehydrogenase,  at  least  by  the  same  mechanism 
as  malonate,  but  have  been  used  because  of  their  ability  to  enter  cells  and 
tissues  readily,  some  hydrolysis  to  active  malonate  within  the  cells  being 
assumed.  The  presence  of  two  keto  groups  on  either  side  of  the  methylene 
group  makes  this  latter  group  more  reactive  and,  indeed,  imparts  some 
acidic  character  to  it,  malonic  diethyl  ester  having  a  p^^  of  approximately 
5  X  10~^^  (Pearson  and  Mills,  1950).  The  rate  of  ionization  is,  however, 
quite  slow  (A;  =  1.8  X  10-^  min-^). 

Chelation  with  Metal  Cations 

Malonate  is  able  to  form  fairly  stable  complexes  with  various  cations 
normally  present  in  media  used  in  metabolic  studies.  The  importance  of 
this  in  malonate  inhibition  will  be  discussed  later  (see  page  66),  and  in  the 
present  section  we  shall  investigate  the  magnitudes  of  the  effects  expected. 
These  complexes  are  chelates  with  a  six-membered  ring  structure,  accounting 

o      or  o.   /O 

for  the  relatively  high  stability  compared  to  complexes  with  the  monocar- 
boxylates.  The  chelate  dissociation  constant  is  given  by: 

(M++)  (A=) 

K=- — -  (1-1) 

(MA) 

where  A=  represents  the  anion  of  any  dicarboxylic  acid.  The  values  for  the 
piii's  of  some  of  the  more  important  chelates  are  given  in  Table  1-4.  These 
constants  are  dependent  on  temperature  and  ionic  strength.  The  pK  for 
Mg-malonate  is  related  to  the  temperature  at  zero  ionic  strength  in  the 
following  way:  pZ  =  2.92  -  0.008  (35  -  «oC)  (Evans  and  Monk,  1952). 
Thus  the  pii's  at  37°  are  approximately  0.1  unit  higher  than  at  25°,  the 
temperature  at  which  the  constants  are  most  commonly  determined.  It 
was  calculated  from  the  data  on  the  complexes  of  malonate  with  Mg++, 
Ca++,  Ba++,  and  Zn++  that  the  pK  at  an  ionic  strength  of  0.15  is  about 


12 


1.    MALONATE 


«       >S 


O 


-M  — I  -- 


O       cc        -^ 
CI         -^         — 1 


(N  -H  rt 


-H  (M  P- 


c 

2 

c5 

o 

o 

+s 

P< 

cS 

3 

3 

T3 

^ 

02 

O 

<Jl 

P^ 


^  CS 


a 

c 

CO 

>> 

S 

^ 

ki 

■^ 

'-^S 

cS 

43 

fcT 

3 

O 

i 

^ 
V 

.S 

V 

s 

is 

E3 

3 

CO 

o 

O 

'53 

CD 

W 

^ 

o 

-1-i 

§ 

ft 

C 

CO 

M 

g 

S 

a 

c3 

•r" 

1 

O 

o 

CO 

a; 

O 
S 

o 

^ 

_C 

to 

o3 

3 

to 

o 

to 

to 

c 

c3 

C 

o 

_o 

'-P 

73 

'-3 

^ 

01 

+3 

eS 

4^ 

^ 

s 

S 
« 

OJ 

s 

V. 

o 

iH 

o 

e 
o 

o 

s 

o 

o 

o 

1 

a 

c 

ft 

>. 

_bi 

J2 

5b 

'S 

CO 

[o 

"c 

c 

eS 

'm 

o 

o 

C 
c3 

CO 

73 

CO 
1 

"s 

P 

<D 

^ 
S 

s 

4^ 

,::— ' 

2 

C 

t4H 

."'^ 

JS 

o 

o 

II 

-»^ 

o 

^ 

o 

+ 
+ 

§ 

3 

03 

to 
3 

aj 

■     ' 

O 

3 

le 

II 

^ 

^ 

3 

C 
> 

+3 
S 

a 

>. 

0; 

'So 

aj 

^ 

J2 

CO 

j3 

'T3 

o 

-t> 

« 

-ti 

_s 

Cm 

C 
O 

o 

CO 

> 

to 

c 

-C 

O 

S) 

_o 

03 

CO 

-C 

'-5 

»o 

-tJ 

S 

C 

CO 

C 
O 

d 
o 

a 

73 

a) 

C 

o 
o 

■t^ 

C 

(h 

C 

M 

a 

,o 

_o 

C 

c« 

"-3 

2 

m 

w 

2 

(0 

a> 

'o 

c3 

CO 

2 

'S 

J5 

a; 

73 

.2 

o 

> 

s 

j; 

'-3 

c 

-p 

cS 

eS 

bl) 

-i-i 

H 

C 

e 

TS 

?> 

13 

C 

3 

cS 

CO 

a* 

CHEMICAL    PKOPERTIES  13 

0.74  unit  lower  than  at  an  ionic  strength  of  zero.  The  free  energy  of  forma- 
tion of  Mg-malonate  is  —  3.90  kcal/mole,  while  .1//°  =  3.2  kcal/mole, 
and  J>S'o  =  23.9  cal/degree  (Evans  and  Monk,  1952;  Chaberek  and  Martell, 
1959,  p.  139).  The  entropy  term  is  large  and  probably  results  mainly  from 
displacement  of  water  from  the  charged  groups. 

The  importance  of  this  chelation  in  inhibition  work  lies  in  the  reductions 
in  the  concentrations  of  the  free  ions  it  may  bring  about,  both  the  metal 
ions  and  malonate.  The  decrease  in  metal  ion  concentration  can  easily  alter 
enzyme  activity  or  cellular  function  and  such  changes  are  apt  to  be  attrib- 
uted to  a  direct  action  of  malonate.  Conversely,  the  malonate  concentra- 
tion may  be  reduced  appreciably.  Examples  of  mutual  concentration  re- 
duction are  given  in  Table  1-5.  The  concentrations  of  Mg++  chosen  approxi- 
mate those  in  Tyrode  solution  (0.11  niM),  Krebs-Ringer  phosphate  medium 
(1.18  mM),  the  usual  media  for  mitochondria  (5  mM),  and  sea  water  (53.6 
TCiM),  while  the  concentrations  of  Ca"'""'"  correspond  to  Krebs-Ringer  phos- 
phate medium  (2.54  mM)  and  sea  water  (10.24  mM).  It  may  be  noted  that 
very  significant  reductions  in  Mg^^  and  Ca++  can  occur  with  concentrations 
of  malonate  commonly  used;  e.g.,  malonate  at  10  mM  will  reduce  the  Mg++ 
49%  and  the  Ca++  23%  in  Krebs-Ringer  medium,  and  higher  concentra- 
tions may  almost  deplete  these  ions  from  the  solution.  When  the  concen- 
trations of  these  cations  are  high,  as  in  sea  water,  the  effective  malonate 
concentration  may  be  reduced  markedly;  e.  g.  malonate  added  to  sea  water 
at  a  total  concentration  of  10  mM  will  result  in  a  1.5  mM  solution  of  free 
malonate  ion.  Such  phenomena  have  usually  been  ignored  or  forgotten 
despite  their  possibly  large  magnitudes.  One  way  of  determining  the  impor- 
tance of  cation  reduction  in  malonate  studies  is  to  calculate  the  reduction 
to  be  expected  in  the  medium  used  and  at  the  malonate  concentration,  and 
then  to  test  the  effects  of  lowering  the  cation  concentration  to  this  extent 
(Rice  and  Berman,  1961).  Other  metal  cations  may  be  reduced  to  a  greater 
extent  than  Mg++  and  Ca++.  Media  initially  1  mM  in  Co++,  Mn++,  or  Cu++ 
will  in  the  presence  of  5  mM  malonate  contain  these  ions  at  concentrations 
of  0.65  mM,  0.34  mM,  and  0.0052  mM,  respectively.  If  such  metal  ions  are 
normally  bound  to  enzymes,  the  degree  of  removal  from  the  enzyme  will, 
of  course,  depend  on  the  relative  affinities  of  the  metal  ion  for  the  enzyme 
and  malonate.  There  is  also  the  possibility  that  malonate  may  chelate  with 
metal  ions  combined  with  the  enzyme,  inactivating  them  for  their  catalytic 
role.  It  should  also  be  remembered  that  similar  phenomena  may  occur  with 
succinate,  and  the  inhibition  kinetics  may  be  distorted  when  malonate  is 
used  due  to  the  differential  reduction  in  the  concentrations  of  these  anions. 

Detection  and  Determination  of  Malonate 

Methods  have  been  developed  for  the  separation  and  identification  of 
organic  acids  from  animal  and  plant  tissues.  Earlier  determinations  involved 


14 


1.    MALONATE 


r^       q 


K 


,— V 

c 

lO 

o 

CO 

CO 

t^ 

CD 

00 

LCl 

(M 

c<l 

o 

+* 

o 

o 

r- 

00 

o: 

in 

LO 

+ 

O 

g 

d 

d 

CO 

t^ 

d 

C5 

Vm 

O 

o 

'^ 

lO 

o 

CO 

CO 

CO 

Oi 

o 

+ 

t- 

CD 

GO 

LO 

l-H 

(^ 

+ 

ai 

t- 

00 

OJ 

i-O 

^_ 

lO 

"-3 

O 

05 

05 

00 

r^ 

CD 

■* 

(M 

cS 

;-i 

-fj 

c 

0) 

o 

t^ 

C3 

g 

<M 

CD 

c- 

t- 

O 

o 

^ 

Is 

05 

t^ 

CD 

o 

CO 

LO 

CD 

o 

GO 

CD 

Tt< 

o 

Tt< 

O 

„_ 

g 

o 

d 

Tji 

d 

d 

GO 

oo 

o3 
g 

-* 

^ 

Tt< 

05 

'Si 

lO 

°3 

C<l 

+ 

CO 

CD 

r^ 

t^ 

i?q 

CI 

o 

o 

+ 

fO 

r-H 

o 

Tt< 

t^ 

00 

o 

cS 

lO 

Tt< 

(M 

05 

LO 

C5 

CD 

O 

c^ 

■M 

Csl 

-' 

-< 

d 

d 

II 

lO 

g 

c 

CD 

lO 

t- 

o 

'N 

_o 

CD 

t^ 

C5 

'M 

CD 

CD 

^ 

Is 

O 

d 

d 

00 

d 

00 

^ 
^ 

o 

CO 

(M 
t^ 

j;^ 

^^- 

-M 

»c 

_o 

o 

^_^ 

'-3 

IC 

+ 

+ 

'M 

CD 

00 

Ci 

(M 

CD 

CD 

h 

Ci 

GO 

00 

"* 

O 

(M 

"S 

^ 

Ol 

C5 

O' 

_h' 

tH 

CO 

t^ 

Tl* 

^ 

-<* 

'* 

CO 

'>^ 

o 

o 

"ip 
c 
o 

t^ 

t^ 

'^^ 

CO 

^ 

CC 

t^ 

•* 

GO 

Tt< 

00 

CD 

,o 

+' 

o 

CD 

CD 

t~ 

00 

00 

Tj< 

Qi) 

+ 

SB 

^ 

1 

d 

O 

CO 

c- 

d 

LO 

lO 

£ 

§ 

^ 

'^ 

-* 

c: 

CJ 

13 

O 

>o 

4^ 

t^ 

l^ 

^M 

CO 

CD 

t^ 

LO 

c2 

o 

+ 

CD 

t^ 

rti 

00 

CO 

t^ 

CD 

E 

Ci 

CD 

CD 

t— 

00 

00 

■* 

'-3 

(-1 

"* 

■^ 

CO 

•M 

-i 

d 

d 

-iJ 

!=l 

^„^ 

0) 

o 

C 

o 

II 

_l 

-* 

t- 

t^ 

C 

CI 

C^ 

O 

•* 

CD 

CS 

^ 

o 

C5 

CD 

in 

CO 

o 

o 

^ 

'^ 

d 

d 

^ 

Ci 

C5 

o 
•^ 

C5 
C5 

"^ 

_g 

S 

'Si 

^H 

+ 

,_, 

rH 

r- 

t^ 

CD 

LO 

o 

'S 

+ 

C5 

t- 

o 

CO 

CD 

C5 

o 

M 

o 

C2 

CD 

>o 

CO 

o 

S 

d 

d 

d 

d 

d 

d 

o 

■~" 

11 

t- 

c 

00 

o 

o 

■* 

C5 

t^ 

ri* 

CO 

<— < 

1— t 

^ 

o 

C5 

C5 

C5 

05 

C5 

CI 

o 

C' 

■* 

d 

d 

d 

d 

g 

•* 

c:> 

^_ 

_^ 

t^ 

d 

+ 

GO 

(^ 

O 

^ 

o 

LO 

t- 

+ 

OS 

o 

CD' 

Tf 

CO 

i-H 

o 

t* 
^ 

o 

o 

O 

o 

o 

o 

d 

d 

d 

d 

d 

d 

d 

Oi 

1  1  i 
?  ^  1 

o 

lO 

o 

o 

o 

o 

OJ 

lO 

o 

<:   s  ^ 

" 

o 


hI 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE  15 

oxidative  titrations  with  hot  acid  permanganate  or  eerie  sulfate  (Willard 
and  Young,  1930;  Christensen  and  Ross,  1941)  and  such  methods  have 
been  used  for  the  analysis  of  malonate  in  the  presence  of  proteins  (Ross  and 
Green,  1941).  Colorimetric  tests,  for  example  with  tetrahydroquinoline-iV^- 
propenal  to  form  blue- violet  compounds  (sensitive  to  0.01  mg  malonic  acid) 
(Dieterle  and  Wenzel,  1944),  have  been  used,  and  several  microcolorimetric 
and  spot  tests  are  available  (Feigl,  1960).  However,  the  most  valuable 
methods  are  chromatographic.  A  large  variety  of  solvents  and  spraying 
agents  have  been  utihzed,  and  different  techniques  have  been  applied: 
ion-exchange  resin  columns  and  partition  chromatography  (Stark  et  al.,  1951; 
Phares  et  al,  1952;  Owens  et  al,  1953;  Shkol'nik,  1954;  Reinbothe,  1957), 
strip  paper  chromatography,  both  one-  and  two-dimensional  (Buch  et  al., 
1952;  Cheftel  et  al,  1952;  Denison  and  Phares,  1952;  Duperon,  1956;  van 
Duuren,  1953;  Jermstad  and  Jensen,  1950;  Kalyankar  et  al,  1952;  Ladd 
and  Nossal,  1954;  Overell.  1952;  Smith  and  Spriestersbach,  1954),  and  cir- 
cular paper  chromatography  (Airan  and  Barnabas,  1953;  Airan  et  al,  1953; 
Barnabas,  1955).  The  use  of  chromatographic  methods  for  the  determina- 
tion of  succinate  and  malonate  in  animal  tissues  is  well  illustrated  in  the 
work  of  Busch  and  Potter  (1952  a,  b;  Busch  et  al.,  1952).  A  moderately  sen- 
sitive method  for  estimating  10-100  //g  of  malonate  by  applying  iodine 
vapor  to  paper  chromatograms  was  developed  by  Dittrich  (1963). 


INHIBITION  OF  SUCCINATE  DEHYDROGENASE 

The  oxidation  of  succinate  in  cells  is  catalyzed  by  a  mitochondrial 
multienzyme  system  which  is  structurally  organized  into  units  that  can  be 
isolated  as  a  particulate  suspension  capable  of  transferring  electrons  from 
succinate  to  oxygen:  this  complex  is  known  as  succinate  oxidase.  Evidence 
will  be  presented  that  the  inhibition  of  this  system  by  malonate  is  related 
to  the  binding  of  the  malonate  to  the  most  proximal  site  in  the  sequence, 
namely,  the  active  center  for  the  attachment  and  dehydrogenation  of  succi- 
nate: this  component  is  known  as  succinate  dehydrogenase.  It  is  impossible 
at  the  present  time  to  define  the  succinate  dehydrogenase  accurately,  be- 
cause it  is  assayed  with  various  electron-accepting  dyes  and  the  basic  mini- 
mal unit  has  not  been  completely  characterized.  The  inhibition  of  succinate 
dehydrogenase  by  malonate  has  generally  been  determined  with  various 
preparations  of  succinate  oxidase  and  not  with  the  isolated  dehydrogenase, 
and  thus  it  is  necessary  to  discuss  briefly  the  nature  of  the  entire  system. 

Properties  of  Succinate   Oxidase 

Knowledge  of  the  components  of  succinate  oxidase  and  the  pathways  of 
electron  flow  has  been  advanced  markedly  in  the  past  few  years  (Mahler, 


16  1.    MALONATE 

1956;  Singer,  et  al.,  1957;  Green  and  Crane,  1958;  Singer  and  Lara,  1957; 
Green  and  Flieseher  1960;  Green,  1960;  Redfearn,  1960).  The  particulate  elec- 
tron transport  particle  (ETP)  of  Green,  obtained  from  heart  mitochondria,  is 
a  succinate  oxidase  preparation  and  has  been  shown  to  contain  the  following 
components  (per  molecular  weight  of  about  5  x  10^):  flavin  dinucleotide  2, 
nonheme  iron  64,  heme  (equal  amounts  of  cytochrome  a,  cytochrome 
b,  and  cytochromes  c  +  Cj)  6,  copper  8,  ubiquinone  (coenzyme  Q)  10,  and 
lipid  34,5%.  In  addition,  there  are  the  several  proteins  with  which  these 
substances  are  bound.  Such  preparations  also  oxidize  NADH,  ascorbate, 
p-phenylenediamine,  and  hydroquinones,  these  substrates  supplying  elec- 
trons at  various  sites  in  the  electron  transport  chain.  The  use  of  various 
electron  donors  and  acceptors,  the  application  of  specific  inhibitors,  and  the 
fragmentation  of  the  succinate  oxidase  complex  have  led  to  several  postu- 
lates of  the  pathways  of  electron  flow.  One  difficulty  is  the  probable  dif- 
ference in  pathways  between  mitochondrial  phosphorylating  systems  and 
submitochondrial  nonphosphorylating  preparations.  A  second  difficulty  is 
the  possibility  of  alternate  pathways  of  electron  flow  rather  than  a  single  lin- 
ear sequence.  Scheme  1-2,  (Green,  1960;  Redfearn,  1960)  might  be  assumed 
provisionally: 

UQ  UQ 

cs       .  i'«/    ^  ^^         ^  ^^f,| 

Succinate  »-  ASF  ^  -• NADH 

iFe)   \^  /^    I     >\  ^   lFe\ 


(1-2) 


Fe  represents  nonheme  iron,  f,  is  the  flavoprotein  associated  with  succinate 
dehydrogenase,  f^  is  the  flavoprotein  associated  with  NADH  dehydrogenase, 
UQ  is  ubiquinone  (coenzyme  Q),  ASF  is  the  antimycin-sensitive  factor,  and 
the  other  symbols  indicate  the  usual  cytochromes.  It  is  possible  that  cyto- 
chrome b  and  ubiquinone  are  common  to  the  two  pathways  from  succinate 
and  NADH,  but  Green  believes  the  evidence  points  to  fusion  of  the  chains 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE  17 

only  at  the  antimycin-sensitive  factor.  It  is  also  possible  that  ubiquinone 
and  cytochrome  b  are  on  a  linear  pathway  rather  than  on  alternate  path- 
ways. In  nonphosphorylating  systems,  the  role  of  cytochrome  b  is  debat- 
able. The  sites  of  oxidation  of  p-phenylenediamine  and  ascorbate  are  distal 
to  ASF.  It  should  also  be  pointed  out  that  succinate  oxidase  from  other 
sources  may  have  different  ratios  of  components  and  somewhat  different 
pathways.  Thus  the  oxidation  of  succinate  and  NADH  by  the  electron 
transport  system  of  Azotobacter  is  not  sensitive  to  antimycin  (Bruemmer  et 
al.,  1957).  Two  recent  observations  have  extended  our  knowledge  of  suc- 
cinate oxidase  and  related  pathways.  Azzone  and  Ernster  (1961)  have 
demonstrated  an  ATP  requirement  for  mitochondrial  phosphorylating 
succinate  oxidation  and  propose  reactions  such  as  the  following: 

Succinate  +  A  -l  ATP    ->     fumarate  +  AH  ~  P  +  ADP 
AH  ~  P  +  B  +  ADP   ->     A  +  BH2  +  ATP 

Succinate  +  B    ->     BHj  +  fumarate 

A  is  possibly  a  flavoprotein,  a  quinone,  or  some  other  component  of  succinate 
dehydrogenase,  and  B  may  be  cytochrome  b.  In  nonphosphorylating  systems 
the  electrons  may  reach  B  more  directly.  An  alternate  pathway  for  AH  -^  P 
is  the  reduction  of  NAD: 

AH  ~  P  +  NAD    ->    A  +  NADH  +  P^ 

Succinate  has  been  shown  to  reduce  NAD  in  mitochondria  by  Chance  and 
Hollunger  (1961  a,  b,  c)  and  this  is  dependent  on  ATP.  Since  cndogen^asly 
formed  succinate  is  more  effective  than  added  succinate,  it  is  possible  that 
this  reaction  may  involve  succinyl-CoA.  The  reduction  of  NAD  passes 
through  antimycin-sensitive  and  Amytal-sensitive  links  and  seems  to  involve 
a  third  flavoprotein  component.  These  findings  are  not  only  important  with 
respect  to  the  behavior  of  succinate  oxidase,  but  also  have  possible  bearing 
on  the  responses  to  malonate  in  mitochondrial  and  cellular  systems. 

The  succinate  oxidase  particles  have  been  fragmented  in  various  ways  to 
yield  smaller  particles  or  soluble  preparations  with  different  compositions 
and  properties.  One  such  preparation  is  the  succinate  dehydrogenase  complex 
(SDC)  from  heart  mitochondria  or  ETP,  which  perhaps  represents  that  frac- 
tion of  the  complex  up  to  cytochrome  c,  since  it  oxidizes  both  succinate  and 
NADH  and  possesses  an  antimycin-sensitive  step.  Of  more  interest  for 
malonate  inhibition  are  the  several  forms  of  soluble  succinate  dehydrogenase 
that  have  been  prepared.  The  purest  contain  no  heme  or  lipid;  four  atoms 
of  tightly  bound  nonheme  iron  and  one  flavin  occur  in  a  molecule  (assuming 
a  molecular  weight  of  around  200,000).  The  flavin  is  apparently  not  ribofla- 
vin but  occurs  in  a  dinucleotide  form  covalently  attached  to  peptide  chains 
of  the  apoenzyme  (Kearney,  1960).  There  is  spectral  evidence  that  both 


18  1.    MALONATE 

flavin  and  nonheme  iron  are  reduced  by  succinate,  and  the  functional  role 
of  iron  in  the  catalysis  is  further  indicated  by  the  inhibition  produced  by 
complexing  the  iron  with  1,10-phenanthroline  or  /J^-globulin  (Singer  et  al., 
1957).  Electron  spin  resonance  studies  of  succinate  oxidase  have  demonstrat- 
ed signals  when  succinate  is  added;  these  free  radicals  seem  to  be  associat- 
ed with  the  dehydrogenase  and  possibly  reflect  changes  in  the  states  of 
iron  or  flavin  (Commoner  and  Hollocher,  1960;  Hollocher  and  Commoner, 
1960).  The  high  sensitivity  of  succinate  dehydrogenase  to  most  sulfhydryl 
reagents  indicates  the  presence  of  an  SH  group  at  or  near  the  succinate- 
binding  site.  One  may,  therefore,  characterize  succinate  dehydrogenase 
from  our  present  knowledge  as  containing  two  cationic  groups  for  the 
binding  of  succinate,  an  SH  group  nearby,  some  nonheme  iron,  and  a  unique 
flavin  dinucleotide  in  a  tight  peptide  complex. 

Relationship  of  Malonate  Inhibition  to  the  Electron  Acceptor  Used 

Measurement  of  malonate  inhibition  involves  either  the  oxygen  uptake  of 
the  complete  succinate  oxidase  system,  or  the  determination  spectroscopi- 
cally  of  the  reduction  of  one  of  the  normal  components  (such  as  cytochrome 
c),  or  the  reduction  of  some  artificial  electron-acceptor  dye.  The  complete 
system  can  reduce  a  variety  of  substances  in  the  presence  of  succinate,  and 
malonate  has  been  shown  to  inhibit  such  reductions  whatever  the  acceptor 
used:  methylene  blue  (Quastel  and  Whetham,  1925;  Hopkins  et  al.,  1938; 
Forssman,  1941;  Franke  1944  a;  Kaltenbach  and  Kalnitsky,  1951  a;  Wad- 
kins  and  Mills,  1955),  ferricyanide  (Stoppani,  1948;  Thorn,  1953),  tetrazolium 
dyes  (Barker,  1953;  Becker  and  Rauschke,  1951;  Zollner  and  Rothemund 
1954;  Waterhouse,  1955),  manganese  dioxide  (Hochster  and  Quastel,  1952), 
2,6-dichlorophenolindophenol  (Repaske,  1954;  Wadkins  and  Mills,  1955, 
Millerd,  1951),  janus  red  (Agosin  and  von  Brand,  1955),  brilliant  cresyl 
blue  (Agosin  and  von  Brand,  1955),  and  lY-methylphenazine  sulfate  (Singer 
et  al.,  1956  b).  Inhibition  of  cytochrome  c  reduction  has  also  been  observed 
Seaman,  1954).  There  is  thus  substantial  evidence  that  malonate  blocks 
electron  flow  very  early  in  the  sequence,  as  would  be  expected  if  it  prevents 
the  binding  of  succinate  to  the  dehydrogenase.  The  most  proximal  location 
of  the  site  of  malonate  inhibition  comes  from  the  work  of  Ziegler  (1961)  on 
electron  transport  particles  from  heart  mitochondria.  Some  of  the  nonheme 
iron  is  reduced  by  succinate  and  this  is  blocked  by  20  raM  malonate.  It 
may  also  be  noted  that  succinate  reduces  NAD  and  NADP  in  submito- 
chondrial  particles  from  heart  through  an  ATP-dependent  system  and  this 
is  readily  inhibited  by  malonate  (Snoswell,  1962;  Hommes,  1963;  Lee  et  al., 
1964),  which  possibly  indicates  that  succinate  dehydrogenase  is  involved. 
The  addition  of  malonate  to  succinate  dehydrogenase  brings  about  changes 
in  the  absorption  spectrum  in  the  flavin  region,  as  do  succinate,  fumarate, 
and  other  competitive  inhibitors  (Dervartanian  and  Veeger,  1962).  There  is 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE  19 

a  decrease  in  absorption  between  400  and  470  m//  and  an  increase  between 
480  and  540  m//,  with  a  maximum  at  510  m//  in  the  difference  spectrum. 
It  is  not  known  what  this  implies  relative  to  the  site  of  malonate  binding. 
The  dyes  may  accept  electrons  from  various  sites  in  the  succinate  oxi- 
dase system.  The  fragments  obtained  from  the  oxidase  particles  differ  in 
their  abilities  to  reduce  these  dyes,  and  as  one  approaches  the  purest  succi- 
nate dehydrogenase  the  number  of  possible  electron  acceptors  is  reduced. 
Indeed,  soluble  succinate  dehydrogenase  reduces  only  the  A^-alkylphenazines 
and  ferricyanide  at  appreciable  rates.  Some  dyes  accept  electrons  chiefly 
from  the  nonheme  iron,  some  accept  more  efficiently  from  the  flavin  com- 
ponent, and  some,  such  as  the  indophenol  dyes,  involve  ubiquinone.  It  may 
well  be  that  none  of  the  commonly  used  dyes  is  completely  selective.  Fer- 
ricyanide, for  example,  may  react  at  other  sites  down  the  chain  in  addition 
to  the  nonheme  iron,  since  its  reduction  has  a  partially  antimycin-sensitive 
component,  and  even  the  A^-alkylphenazines,  considered  to  be  the  most 
reliable  acceptors  at  the  dehydrogenase  level,  may  be  able  to  react  at  other 
sites. 

The  question  of  importance  with  respect  to  malonate  inhibition  is  whether 
all  of  the  methods  of  measuring  succinate  dehydrogenase  activity  are  equiva- 
lent for  the  purpose  of  obtaining  accurate  kinetic  results.  Unfortunately, 
there  have  been  very  few  reliable  investigations  wherein  different  methods 
or  acceptors  have  been  compared.  The  inhibitions  of  succinate  oxidase 
(manometric)  and  succinate  dehydrogenase  (dye  acceptors)  at  the  same 
concentrations  of  succinate  and  malonate  have  been  reported  sporadically. 
It  has  generally  been  found  that  the  oxidase  is  inhibited  somewhat  more 
strongly:  Arum  spadix  (Simon,  1957),  Limidus  gill  cartilage  (Person  and 
Fine,  1959),  and  potato  tubers  (Millerd,  1951).  But  in  the  enzymes  from 
Xanthomonas,  this  is  reversed  (Madsen,  1960).  The  results  of  Millerd  are 
particularly  difficult  to  understand,  inasmuch  as  she  found  a  40%  inhibi- 
tion of  the  oxidase  at  0.1  raM  malonate  whereas  the  dehydrogenase  (using 
2,6-dichlorophenolindophenol  as  acceptor)  was  not  inhibited  at  all.  The 
author  is  not  aware  of  any  quantitative  studies  comparing  malonate  inhi- 
bition with  different  dye  acceptors.  It  would  appear  that  most  workers 
have  assumed  there  would  be  no  difference. 

Actually,  from  kinetic  consideration,  it  is  not  at  all  necessary  that  the 
inhibition  produced  by  a  chosen  concentration  of  malonate  be  the  same 
when  different  acceptors  are  used.  In  fact,  the  inhibition  may  depend  on  the 
acceptor  concentration.  Thorn  (1953)  showed  that  the  inhibition  of  pig  heart 
succinate  dehj^drogenase  by  0.536  mM  malonate  increases  with  the  con- 
centration of  methylene  blue:  at  0.15  n\M  methylene  blue  the  inhibition 
is  15.9%  and  at  3  vaM  methylene  blue  it  is  28.8%.  Succinate  oxidase,  is 
inhibited  52.6%  at  the  same  malonate  concentration.  The  transfer  of  hydro- 
gen atoms  from  succinate  to  a  dye  acceptor  always  involves  the  interaction 


20  1.    MALONATE 

of  these  two  substances  with  the  enzyme  surface,  undoubtedly  at  different 
sites.  At  low  concentrations  of  the  acceptor,  or  with  weak  acceptors,  the 
over-all  rate  may  not  be  determined  by  the  rate  at  which  the  hydrogen  atoms 
are  removed  from  the  succinate,  but  may  depend  also  on  the  rate  of  transfer 
to  the  acceptor.  The  malonate  inhibition  will  thus  vary  with  the  degree  of 
saturation  of  the  enzyme  with  acceptor.  On  this  basis  it  would  seem  reason- 
able to  use  those  acceptors  which  are  the  most  active  and  react  with 
sites  closest  to  the  succinate  site.  A  further  consideration  is  the  inhibition 
produced  by  the  acceptors  themselves.  Both  ferricyanide  and  iV-methylphe- 
nazine  begin  to  inhibit  succinate  dehydrogenase  as  the  concentration  is 
raised  above  certain  levels.  Such  systems  would  then  constitute  examples 
of  multiple  inhibition  and  the  kinetics  of  the  inhibition  due  to  malonate 
alone  may  be  distorted.  The  choice  of  the  acceptor  and  its  concentration 
is  thus  of  some  significance. 

Site  of  Inhibition  by  Malonate  in  the  Succinate  Oxidase  Sequence 

The  results  discussed  in  the  preceding  section  point  clearly  to  the  site 
of  inhibition  as  succinate  dehydrogenase.  Indeed,  inhibition  of  soluble 
succinate  dehydrogenase  by  malonate  has  been  demonstrated.  The  competi- 
tive nature  of  the  inhibition,  to  be  treated  in  the  following  section,  indicates 
the  inhibition  to  be  at  the  active  site  at  which  succinate  is  bound.  There  is 
thus  no  question  but  that  the  major  site  of  inhibition  is  at  the  very  begin- 
ning of  the  electron  transport  sequence  in  succinate  oxidase.  The  question 
that  now  must  be  considered  is  whether  malonate  can  inhibit  at  any  other 
step  of  the  electron  transport  chain. 

There  are  two  obvious  ways  to  examine  this.  One  is  to  test  the  action 
of  malonate  on  the  succinate  oxidase  system,  using  substrates  that  donate 
electrons  at  more  distal  sites  than  succinate.  The  other  way  is  to  determine 
the  response  of  other  oxidases  that  utilize  most  of  the  electron  carriers 
in  the  succinate  oxidase.  Quastel  and  Wheatley  (1931)  observed  that  ma- 
lonate at  67  mM  does  not  inhibit  the  oxidation  of  p-phenylenediamine 
and  hence  concluded  that  the  cytochrome  region  of  the  sequence  is  immune 
in  their  preparations.  Actually,  many  oxidases,  comprising  varying  segments 
of  the  electron  transport  chain,  have  been  found  to  be  insensitive  to  malo- 
nate at  concentrations  from  25  mlf  to  50  mM.  All  this  evidence  points  to 
a  rather  specific  action  on  the  dehydrogenase.  However,  there  are  data  in 
the  literature  which  indicate  that,  at  least  in  some  species  and  at  high  enough 
malonate  concentrations,  inhibition  at  other  sites  may  occur.  Although  30 
mM  malonate  does  not  inhibit  NADH  oxidation  in  beet  mitochondria 
(Wiskich  et  al.,  1960),  some  inhibition  has  been  observed  in  mosquito  parti- 
cles (10-15%  at  1-10  mM)  (Gonda  et  al,  1957)  and  in  Tetrahymena  NADH 
oxidase  (35%  at  6.2  mM  and  70%  at  18  mM)  (Eichel,  1959).  Such  inhi- 
bition could,  of  course,  be  on  the  NADH  dehydrogenase  rather  than  on 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE 


21 


enzymes  common  to  the  succinate  pathway.  Malonate  does  not  usually 
interfere  with  ascorbate  oxidation,  but  in  the  silkworm  it  was  found  both 
spectroscopically  and  manometrically  that  malonate  inhibits  the  oxida- 
tions of  succinate  and  ascorbate  equally  (Sanborn  and  Williams,  1950). 
Observations  such  as  these,  coupled  with  those  showing  greater  inhibition 
of  succinate  oxidase  compared  to  the  dehydrogenase,  make  it  necessary 
to  exert  some  caution  in  assuming  a  completely  specific  action  in  all  cases. 

Competitive   Nature  of  the   Inhibition 

It  has  been  often  stated  that  competitive  inhibition  was  first  demonstrat- 
ed by  Quastel  and  Wooldridge  (1928)  for  the  inhibition  of  E.  coli  succinate 
dehydrogenase  by  malonate.  However,  inhibitions  were  not  calculated 
and  the  data  presented  do  not  lend  themselves  to  quantitative  interpreta- 
tion. Indeed,  when  their  data  are  plotted  on  a  l/w,  —  1/(S)  graph  (Fig.  1-2)  a 
straight  line  is  not  obtained.  Since  no  comparable  control  studies  were  done 
in  the  absence  of  malonate,  the  fact  alone  that  increasing  the  succinate  con- 
centration increases  the  rate  in  the  presence  of  malonate  does  not  prove 
competition.  These  points  are  brought  out  not  to  criticize  pioneering  work 
but  to  illustrate  that  conclusions  about  the  type  of  inhibition  cannot  be 
made  so  readily  as  many  imagine. 


Fig.  1-2.  A  double  reciprocal  plot  for  the 
inhibition  of  E.  coli  succinate  dehydrogenase 
by  malonate  at  1.43  mM.  Succinate  concen- 
trations are  in  vaM.  (Data  from  Quastel  and 
Wooldridge,  1928). 


22 


1.    MALONATE 


Competition  between  succinate  and  malonate  has  been  claimed  to  occur  in 
the  following:  rat  liver  homogenates  (Potter  and  DuBois,  1943),  oyster 
muscle  homogenates  (Humphrey,  1947),  oyster  egg  homogenates  (Cleland, 
1949),  carrot  root  (Hanly  et  al,  1952),  yeast  (Krebs  et  al,  1952),  pig  heart 
particulates  (Thorn,  1953),  cockroach  muscle  homogenates  (Harvey  and 
Beck,  1953),  the  trypanosome  Crithidia  (Hunter,  1960),  and  the  soluble 
succinate  dehydrogenase  from  heart  (Keilin  and  King,  1960).  In  all  cases, 
the  inhibition  produced  by  a  certain  malonate  concentration  is  reduced  by 
increasing  the  succinate  concentration,  but  a  quantitative  analysis  of  the 
data  has  been  seldom  carried  out.  Most  of  these  studies  were  made  with 
the  usual  assay  methods,  but  competition  has  recently  been  shown  by  meas- 
uring the  reductions  in  the  electron  spin  resonance  signals  produced  by 
malonate  at  different  succinate  concentrations  (Commoner  and  Hollocher, 
1960). 


Fig.   1-3.  A  single-curve  plot  for  the  inhibition  of 

cockroach  muscle  succinate  oxidase  by  malonate  at 

0.33  mM.  Ki  =  0.105  mM  and  KJK^  =  275.  (Data 

from  Harvey  and  Beck,  1953). 


Single-curve  plots  (type  F,  see  Chapter  1-5)  were  made  for  two  of  the 
inhibitions  mentioned  above  (Figs.  1-3  and  1-4).  In  the  case  of  Crithidia, 
Hunter  showed  competition  with  a  1/'U-1/(S)  plot  and  the  single-curve  plot 
confirms  this,  K^  having  the  value  of  0.22  mM  and  K^^JK^  calculated  from 
the  slope  a  value  of  53.  The  results  with  cockroach  muscle  likewise  fall 
roughly  on  a  straight  line,  giving  K,  as  0.11  mM  and  KJK^  as  275,  values 
differing  somewhat  from  those  calculated  by  Harvey  and  Beck  using  a 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE 


23 


different  plotting  procedure  {K^  =  0.13  mM,  and  KJK^  =  200).  In  most 
studies  insufficient  data  are  available  for  plotting.  It  must  be  emphasized 
that  a  change  in  inhibition  observed  at  two  or  three  succinate  concentra- 
tions is  not  adequate  to  prove  a  purely  competitive  inhibition.  It  seems 
to  be  rarely  considered  that  an  inhibition  may  not  be  either  completely 
competitive  or  completely  noncompetitive.  The  conditions  for  partially 
competitive  inhibition  were  given  (Eqs.  1-3-14  and  1-3-15)  and  the  types 
of  plot  to  be  expected  discussed  (Chapter  1-5).  An  interesting  example 
of  this  is  provided  by  the  work  of  Honda  and  Muenster  (1961)  on  the  inhi- 
bition of  succinate  oxidation  in  lupine  mitochondria.  Here  the  osmolarity 
of  the  preparation  and  assay  media  was  varied  with  sucrose,  and  it  was 
found  that  the  interaction  constant,  a,  defined  in  Eqs.  1-3-5  and  1-3-6, 


Fig.   1-4.  A  single-curve  plot  for  the  inhibition   of 

Crithidia   succinate   dehydrogenase   by  malonate  at 

1  mM.  Ki  =  0.22  mM,  and  KJE^  =  .53.  (Data  from 

Hunter,  1960). 


varies  quite  markedly  from  values  indicating  nearly  completely  competitive 
inhibition  to  those  showing  noncompetitive  inhibition.  This  work  will  be 
discussed  in  greater  detail  in  a  later  section  (see  page  46),  but  it  suffices 
to  show  that  partial  competitive  inhibition  by  malonate  is  possible  and  that 
the  type  of  inhibition  may  vary  with  the  experimental  conditions. 

The  most  elegant  treatment  of  malonate  inhibition  is  by  Thorn  (1953) 
at  the  St.  Thomas's  Hospital  Medical  School  in  London,  using  succinate 
oxidase  preparations  from  pig  heart  muscle.  The  activity  was  measured  by 
the  reduction  of  ferricyanide  in  the  presence  of  cyanide  to  block  the  cyto- 


24 


1.    MALONATE 


chrome  pathway.  The  usual  Ijv  -  1/(S)  plots  (Fig.  1-5)  show  apparently 
completely  competitive  inhibition,  and  values  of  the  substrate  and  inhibi- 
tor constants,  to  be  discussed  in  the  next  section,  were  calculated.  Using 
average  values  of  K„^  and  K^  obtained  by  Thorn,  the  inhibition  curves  in 


(Malonate)  =  0  0089mM 


No  Malonate 


Fig.  1-5.  Double  reciprocal  plots  for  the  inhibition 
of  pig  heart  succinate  dehydrogenase  by  malonate, 
showing  pure  competitive  inhibition.  Succinate  con- 
centrations are  in  mM  and  v  in  spectrophotometric 
units.  (Data  from  Thorn,  1953). 


Fig.  1-6  were  plotted.  These  curves  show  the  expected  reduction  in  inhi- 
bition as  the  succinate  concentration  is  increased  at  constant  values  of 
malonate  concentration.  It  may  be  noted  that  overcoming  the  inhibition 
is  much  more  difficult  at  high  inhibitor  concentrations  and  this  must  be 
taken  into  account  in  experiments  designed  to  show  a  competitive  type  of 
action.  K„i  was  used  rather  than  K,.  because  under  the  usual  experimental 
conditions  it  is  this  constant  that  determines  the  behavior. 

A  word  must  now  be  said  about  malonate  reversibility.  One  of  the  cri- 
teria of  competitive  inhibition  is  that  the  inhibitor  should  leave  the  active 
site  readily  when  its  concentration  is  reduced,  or  that  it  should  be  displaced 
rapidly  when  more  substrate  is  added.  These  points  have  seldom  been 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE 


25 


tested  with  malonate.  However,  it  was  demonstrated  quite  early  that  the 
inhibition  of  rabbit  muscle  succinate  dehydrogenase  is  reversible  by  washing; 
the  preparation  was  incubated  for  30  min.  with  100  milf  malonate  anaerobi- 
cally  and  then  washed  3  times  on  a  filter  —  the  reduction  times  were  15 
min  for  the  control,  61  min  with  the  malonate,  and  17  min  for  the  washed 
preparation  (Hopkins  et  al.,  1938).  A  trypanosomal  succinate  dehydroge- 
nase, however,  showed  no  reversal  of  the  inhibition,  even  when  succinate 
was  added  to  50  times  the  malonate  concentration,  which  is  particularly 
surprising  since  in  the  living  cells  a  good  reversal  was  observed  (Agosin  and 
von  Brand,  1955).  The  concentrations  of  malonate  used  here  were  not 
unduly  high  and  so  these  data  are  unexplainable. 


Fig.  1-6.  Curves  showing  the  calculated  reductions 
of  the  malonate  inhibition  of  succinate  dehydrogen- 
ase by  various  concentrations  of  succinate,  using 
K„  =  0.366  mM,  and  K^  =  0.0076  mM,  as  given 
by  Thorn  (1953). 


Constants  of  the  Inhibition 

It  is  not  surprising  that  the  values  of  the  inhibitor  constant,  K^,  for 
malonate  inhibition  are  quite  variable  in  the  literature,  because  not  only 
do  the  experimental  conditions  affect  this  constant  markedly  but  the  suc- 
cinate dehydrogenase  varies  in  its  properties  from  species  to  species.  From 
the  values  in  the  accompanying  tabulation  and  from  reasonable  estimates 
based  on  assumed  or  determined  substrate  constants  for  the  data  in  Table 
1-6,  it  is  seen  that  the  K^'s  for  most  preparations  vary  between  0.005  mM 


26 


1.    MALONATE 


< 

1? 

8  s 

g 

s  "~" 

5 

C/3 

o 

u 

M 

C 

[i- 

_o 

O 

■-S 

Z 

(1 

o 

c3 

& 

? 

0; 

S 

li 

5 

;z 

P5 


s  ^ 


2 
'o 

3 

05 

:3;. 

-2 

c 

c5 

o 

S 

13 

:^ 

"e 

2 

c3 

-^ 

(» 

C 

C5 

"^ 

(h 

J3 
o 

c 

cS 

03 

2 

3 

c 

N 

'^ 

S 

7^ 

_« 

03 

0; 

(-1 

W 

O" 

<l 

o 

<; 

Qi 

m 

m 

O 

o 

(M    LO 

lO 

t- 

QO 

lO 

o 

t~ 

o 

(N    O    <N 

c5 

IC 

t>    Ol 

A 

05 

GO 

<r> 

o 

c- 

o 

^    C^J     t> 

13 

^ 

C 

e 

•*-* 

^ 

<u 

e 

73 
C 

'«3 

>. 

s 

o 
c 

3 

73 

o 

s 

O    O    O    -H    o    -^ 


o  o  o 

Tt<     (M     Tt< 


a>  73 


w  TO  ^  W  M 


c?    s;   .ii 


H 


H  c^ 


O     c3 
C»    Ph 


w 


w 


S 


fej 


^    W 


d  d 


O-^t^       coo       'MCC'M       CO      —    O 


s     s 


^  ^ 


-S 

ei 

"3 

TO 

f, 

_o 

o 

cS 

Ch 

'-^ 

b4 

Ch 

"c 

-»^ 

X 

c3 

o 

X 

W 

|1h 

M 

W 

g 

•ra 

s 

s 

o 

^«e 

S) 

a 

?■ 

INHIBITION    OF    SUCCINATE    DEHYDROGENASE 


27 


^    pq 


>> 

J 


Tf     C5     O     C-1 

o   ^  00  05 


§ 

o 

■*  r-  ci  o 
Tt<    o    u-j    t-    ^ 

o 

o 

o  o 

o 

o  o  -<  o  o 

o 

•— I      .^    c^    »o 


LO    L-^    CO    CO    CC    M 
C-l     (M     CO    CO    CO    CO 


-M    (N    Tff    00    O 


P^ 


a        3 


-^      ''I 


Ti 

o 

C 

o 

CS 

^ 

0) 

cS 

-ti> 

T-! 

c 

C 

p 

O 

W 

a 

00 

lO 

CO 

00 

o 
d    -H 


P     '-3 
5     Is 


05 


a 


s     ts 


PM 


« 

Ol 

•^ 

« 

S 

^ 

s 

>, 

■e 

a; 

s 

s 

s  a 


or,    e 
s   s 


^  -^    a,    ^ 


fin     P4 


M    2 


^    g 


05       O    C5    CO       Oi    O 
00  C<|    t^      CO    o 


-H      O    O    — I      o    ^ 


,S  ^         (U 


s  (^ 


« 


*    a 


H     ^ 


M 


28 


1.    MALONATE 


05 


1    S 


PL, 


OO    00    00 

M    00    CC 


a 

ft 


,0     .2 

c     o 


00 

^ 

05 

i-H 

'•-*' 

>> 

, , 

V 

CO 

^4 

lO 

^ 

Oi 

ft 

s 

■-^ 

s 

2 

W 

o 
o 

t^ 

T3 

s 

2 

C 
c3 

OO 

t3 

^^ 

>. 

05 

e3 

IN 

O 

>. 

<D 

05 

p 

(h 

c 

tx 

^ 

b 

o 

^_^ 

j3 

ft 

"^ 

ft 

-*^ 

T3 

C3 

c 

^ 

rt 

S 

^ 

^ 

1-5 

0) 

3 

« 

w 

05 

w 

i-O-^C-       M       00       f-      OOOCO       OOC<I 
MI>05      Oi      C5       >Cl       t-t>00      CDOO 


-^      O      OO-H      -HC^i      o    — 


m    o  -^  ic    s*5 


lO       W       lO 


«      2 


cS 

cS 

C 

C 

0) 

a> 

bC 

M 

O 

O 

a 

S 

o 

o 

w 

X 

■i 

•S 

Si 

>> 

O 

=0 

1 

1 

Si 

s 

5 

g 

e 

o 

=0 

>, 

s 

OQ 

+i 

g 

"S) 

S 

CI 

§ 

SQ 

^ 

^"^ 

a: 

O 

w 

INHIBITION    OF    SUCCINATE    DEHYDROGENASE  29 


m 


s 

i>         •-  .s; 

»c         s  ^ 

c^        ^  i: 


O    W    CO      -H 


o 

eS 

s 

« 

rj 

a 

cS 

o 

Ut 

o 

s 

o 

O 

cc 

p^ 

QO     CO 

Oi   <u   a 

fO    X 

CO 

lO    00 

CO    73 


00 

CO 

IM 

o 

CO    CO 

t^ 

CO 

ut 

»o 

9 

O    CO 

t^ 

o 

(M 

■<*  o 

s 

cS 

o3 

eS 

eS 

s 

^ 

c 

a 

C 

C 

C 

cS 

cS 

v 

o 

0) 

a> 

a> 

b£ 

M 

bD 

bc 

bc 

3 

"3 

O 

O 

O 

o 

O 

« 

o 

s 

S 

S 

S 

S 

'•S 

'•S 

o 

o 

o 

o 

o 

c3 

ei 

W 

W 

be 

W 

a 

K 

PLi 

Ph 

^ 

'-3 

j2 

CO 

3 

00 

3 

O 

o 

^ 

S 

s 

^ 

§ 

OOOCOO— 'O'CO— <o  O  OOO  OO-^iOO  ^ 


2  §    £    g     > 


^  -^     -8 


s 

s 

■4) 

1) 

e 

-iS 

s 

^ 

"a 

QB 

1 

bc 

3 

Co 

-2. 

J3 

•£ 
£ 

^^ 

3 

i 

t 

o 

1 
2 
o 

to 

-2 

a; 

o 

o 
O 

o 

o 

8 

i;5 

o 

o 

g 

3 

cr 

03 

o 

1 

a 

S-i 

O 

a 

c 
IS 

c 
o 

b£ 

W 

1 

s 

cc 

« 

PM 

§ 

30 


1.    MALONATE 


tf 


fe 


o 

M 

03 


OS     ,-. 

Oi    lO    ^^ 

^^     05     CO 


-ii 


^    -^ 

i;        O        O 

<  ffi  § 


Ah 


;?? 


I 


»CC0(M05O'^00'^20O 


Tti>p-Hco-*t-?cOQOirjTt*i>os 


o  o  o  o 


o  o  o  o  — < 


oooooooo-^ooo 
o^ooooooooooo 


OOOO-HOOOO^ 


^   ■^   lO         CO         CO   CO         CO 

-^OOO'-HO^C0O'-<C0 


^ 


G     Si     S 
O     («1     o 

ffi  H  W 


c 
? 
bo 
o 

s 

o 

w 


w 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE 


31 


C5    fi^  2 


^  ^  ^ 

CO  CO  o 

IC  »0  lO 

05  0  05 


^  ^  -Q 

^  X!  ^ 

(B  <U  4) 

T3  T3  TS 

C  C  C 


b    TJ      t^.     t^.     >•, 


<U      ^     C.     G     S     Oi     <b     a 

S    Si  -t^    o        £    S    S 

O-SfQMCoOO 
MO!.       O     *     MMM 

GOSOoOOO 


<N    (M 

«c  »o 

05     05 

i-H      t-H 

^_^ 

^.„^ 

^ —     ^ — 

uo 

>-o 

1— 1     f— 

lO 

o 

<0      <D 

05 

05 

-^     -^ 

03       tC 

c5      cS 

3     S 

o 

o 

GfC? 

m 

<0 

CO 

eS     c3 

73 

-o 

;h     Ld 

C 

c 

o    o 


COOO»CCO(MfOC<lfOM-^Ot'^        00050COIOO 
00O51O00O5GOO500  -hC000O5CO        -HF^cOlOOt^M 


>>    >> 


j3 

-S3 

^^ 

^ 

T3 

73 

00 

o 

C 

C 

05 

eS 

c3 

^ 

^^ 

^^ 

^~^ 

<U 

OJ 

;h 

^ 

-IJ 

IB 

in 

^ 

P 

p 

cS 

O 

a  o-  PM 

Ph 

t-      -H 

o 

»o 

o 

o 

Tf      'O 

t- 

C5 

C5 

o 

o 

^H  O  O-HfCCOO-^fOO 

CO  O  0-H(N  lOcOCOOOOOOOOO 


lO    -^    CO    lO    o 


o 
in  lo  lo  o 


o-Hioo>o      -H-Moo-^eoo 


CO.  CO  CO 

O    ^    CO 


^iW       1.,/       -V       ^ig       ^g       I.U 


T3 

C     C 

rt  -3  -r!  -c 
c    a    c    c 

O 

+2 

o    o 

S<.   o    o    o 

j: 

o 

js  j= 

o 

rt 

o     o 

2     o     o     o 

o 

o    o 

s    o    o    o 

■i^    -t^ 

—     -u     HJ     -u 

S 

w 

s  s 

W  §  S  S 

g  =« 

eS 

r; 

bC    P 

3 

1 

O     o 

■-3 

><     O     cS 

w  W  pm 


^1 


3  §    w 


■ti     '-H      r-'     -*^ 


•3     "    .3 


w 


Oh  S  pq  W 


t-  CI  .3 

b.  1^  3 

.£  fc;  (B 

i-:i  M  cc 


2    2 


-5       rH     -tJ 


pq  M 


(30 

■ft 

eS 
(D 

C 

'3 


32 


1.    MALONATE 


tf 


c8 
C 

^ 

s 

_o 

•r^ 

"3 

s 

o 

§ 

^ 

CO 

+3 

J^ 

eS 

Q 

,  J5 

"^i 

u 

o 

c 

tS) 

o 

s 

n 

3 

"" — 

< 

W 

H 

Oh 


fq 


Pi  ;^ 


aj    3 


-C      ^  k-H 


CO  05 


«    o 


CO 

__^ 

S 

05 

;o 

tH 

»o 

m 

■^-^ 

05 

C 
O 

>> 
-2 

^ 

o 

a;) 

pq 

_N 

rt 

« 

"3 

CO 
CO 

s 

52    !- 

c 

73 

c 

o 

3 

ci 

^ 

c 

03 

^ 

oj     3 

c3 

cS 

b 

d    ^ 

Lh 

o 

3 

03 

Q  H 

fe 

W 

o*  m 

O    CO    CO 

o  o 

!M 

t- 

-H      -H      CO 

lO     05     05 

05    oo 

00 

00 

O    00    00 

rt     O  -H  O 


rt  (M  -H 


■* 

(M 

»0    >^    ■*    'M 

o 

o  Tf<  c^  CO 

o 

O    <M    ^    O    '^1    O         -H 


CO    -^    -^    "^  CO 

CO   t^   t^   t-~   o    -^ 


-i3     a> 

2   ^ 

3 

c    ce 

55    '^ 

eS 

O   -3 

0)     — . 
M     3 

-2  -3 

o 

+3 

3  t 

O    .O 
3    -k^ 

11 

1 

V3      cS 

O      c3 

3 

O     ci 

"S 

>1 

S   Ah 

W    P^ 

;^ 

M   Ah 

w 

w 

>. 

3 

a)    *3 

p 

3    >- 

■C    S5 

K»> 

.«     o 

M  W 

H 

M 


Ul 

<*- 

o 

"eS 

X 

Q 

o 

o 

lO    -H    (N 

-H    CO    CO 


-^      -H      ^  CO 


3     -►^ 


"3  s 


13 

o 

_3 

3 

'o 

^3 

;h 

"i 

'& 

PQ 

W 

INHIBITION    OF    SUCCINATE    DEHYDROGENASE 


33 


Preparation 


Ki  {mM) 


Reference 


Pig  heart 

0.0076 

Rat  heart 

0.01 

Yeast 

0.0105 

Ehizobium  japonicum 

0.017 

Claviceps  purpurea 

0.03 

Corynebacterium  diphtheriae 

0.037 

Beef  heart 

0.041 

Beef  heart 

0.045 

Mytilus  edulis 

0.06 

Beef  heart 

0.13 

Micrococcus  lactilyticus 

0.23 

Phaseolus  vulgaris 

0.24 

Lupinus  albus 

0.91 

Thorn  (1953) 

ITyhn  and  Matsumoto   (1964) 

Ryan  and  King  (1962a) 

Cheniae  and  Evans  (1959) 

McDonald  et  al.  (1963) 

Strauss  and  Jann  (1956) 

Kearney  (1957) 

Keihn  and  King  (1960) 

Ryan  and  King  (1962  a) 

Lee  et  al.  (1964) 

Warringa  and  Giuditta  (1958) 

Hiatt  (1961) 

Honda  and  Muenster  (1961) 


and  0.05  vaM,  and  mammalian  tissues  generally  yield  quite  low  constants. 
The  interest  in  the  K,  lies  in  its  relationship  to  the  binding  energy  of  mal- 
onate  to  the  active  center,  so  that  variations  in  the  K^,  unless  due  to  exper- 
imental conditions,  may  be  attributed  to  differences  in  the  topography 
of  the  enzyme  surface.  The  values  of  £",„  for  different  preparations  are  not 
particularly  significant,  except  for  characterizing  a  certain  preparation 
under  specified  conditions,  because  .K",,,  does  not  usually  equal  K^..  Inasmuch 
as  the  ratio  of  K^  to  K j  is  of  some  significance  in  interpreting  interactions 
with  the  active  center,  it  will  be  worthwhile  to  discuss  the  only  instance  in 
which  this  has  been  determined. 

The  ratio,  K„,!K„  has  been  given  by  various  investigators  as  ranging 
from  10  to  60.  Thorn  (1953)  pointed  out  that  K,„  for  succinate  dehydroge- 
nase is  often  quite  different  than  K,  (Slater  and  Bonner,  1952).  K,„  thus 
equals  (A-j  +  A'2)/A:j,  and  A-j  is  dependent  on  the  experimental  conditions. 
Thorn  obtained  values  of  KJK^  from  3  to  60  by  varying  the  electron- 
acceptor  dye  and  the  reaction  rate.  The  faster  the  rate,  the  greater  the  de- 
viation of  ^,„  from  K^.  Thus  extrapolation  to  zero  rate  from  a  series  of 
experiments  at  different  methylene  blue  concentrations  enabled  Thorn  to 
determine  the  true  ratio  of  succinate  and  malonate  affinities;  KJK,  turned 
out  to  be  3.  Since  K,  is  approximately  0.0076  milf .  K,  =  0.023  mM.  a  value 
which  checks  well  with  directly  determined  values  of  the  rate  constants 
(A'i  =  3.35x  10^,  and  A-_i  =  0.99).  Similar  low  ratios  would  probably  be  found 
in  most  succinate  dehydrogenase  preparations.  The  difference  in  binding 
between  succinate  and  malonate  is  thus  not  so  great  as  previously  believed. 

It  is  of  great  practical  importance  to  realize  that  the  degree  of  inhi- 
bition of  succinate  oxidase  by  malonate  varies  with  the  rate  of  succinate 
oxidation  and  the  electron  acceptors  present.  This  may  explain  some  of  the 


34  1.    MALONATE 

differences  in  Table  1-6.  If  one  is  to  compare  the  inhibitory  potencies  of 
malonate  on  succinate  oxidases  from  different  tissues  or  species,  equivalent 
rates  of  succinate  oxidation  should  be  used  and  similar  methods  of  deter- 
mining the  rate  should  be  employed. 

Inhibition   of  Succinate   Dehydrogenase   by   Other   Dicarboxylates 

Before  discussing  the  more  intimate  nature  of  the  inhibition  and  the 
possible  ways  by  which  malonate  is  bound  to  the  active  center,  it  will  be 
useful  to  consider  the  inhibitory  potencies  of  other  dicarboxylate  ions. 
The  configuration  of  an  active  center  may  often  be  approached  by  comparing 
the  relative  affinities  of  analogous  compounds  for  the  enzyme.  At  the  end  of 
this  chapter  a  more  complete  discussion  of  inhibitors  related  to  malonate 
will  be  given;  for  the  present  we  shall  be  interested  only  in  the  inhibition 
of  succinate  dehydrogenase.  Inhibitions  and  inhibitor  constants  are  summa- 
rized in  Tables  1-7  and  1-8. 

(a)  U nsubstituted  dicarboxylate  ions.  It  was  stated  by  Quastel  and  Whet- 
ham  (1925)  that  oxalate,  glutarate,  and  adipate  do  not  interfere  with 
succinate  oxidation,  in  contrast  to  malonate,  and,  although  in  later  work 
they  found  some  inhibition,  this  has  generally  been  confirmed.  Accurate 
comparisons  must  be  made  using  inhibitor  constants,  and  these  are  not 
available,  but  there  is  no  doubt  that  in  the  series  ~00C — (CHa)^ — C00~ 
the  inhibition  reaches  a  sharp  maximum  at  w  =  1.  One  must  assume  that 
malonate  best  fits  the  intercationic  distance  on  the  active  center  of  suc- 
cinate dehydrogenase,  and  this  includes  succinate  itself  since  K^  is  generally 
larger  than  K,  for  malonate. 

(6)  Unsaturated  dicarboxylate  ions.  It  is  interesting  to  compare  the  two 
isomers,  fumarate  and  maleate,  since  they  differ  in  the  intercarboxylate 
distance  (Table  1-1).  For  mammalian  succinate  dehydrogenase  it  would 
appear  that  fumarate  is  bound  much  less  tightly  than  either  succinate  or 
malonate  (Table  1-8),  but  this  does  not  necessarily  hold  for  the  bacterial 
enzymes,  providing  evidence  that  the  active  center  configurations  may  be 
quite  different  in  different  dehydrogenases.  One  might  expect  a  rather  low 
affinity  for  fumarate  because  of  the  fairly  long  intercarboxylate  distance, 
but  it  is  surprising  that  maleate  is  such  a  poor  inhibitor  inasmuch  as  its  inter- 
carboxylate distance  in  only  0.38  A  greater  than  in  malonate.  A  complicating 
factor  is  the  ability  of  maleate  to  react  with  SH  groups,  and  actually  the 
inhibitions  observed  by  Hopkins  et  al.  (1938)  and  Morgan  and  Friedmann 
(1938  b)  were  obtained  only  after  prolonged  incubation.  The  possible 
significance  of  these  observations  will  be  discussed  in  the  next  section. 
On  the  other  hand,  acetylene-dicarboxylate,  with  a  distance  of  5.16  A 
between  carboxylate  groups  and  a  restriction  to  linearity,  does  inhibit  and 
this  is  completely  competitive  (Thomson,  1959). 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE 


35 


P5 


IS  ^ 


«>  lii 


S^      2 


P    T3 


s 

15 

5 

T3 

T) 

S 

3 

C 

cS 

c3 

t4 

^2 

CO 

S 

.^ 

(S 

OI 

o 

^ 

O 

(1^ 

c? 

tf 

w 


Q  Ec5 


<M    2    r-  05 

Oi  _  eo        -^ 


-  QU  -— - 

S  "«  - 

S  'E  S 

<u  "o  X 

^  ^  ^ 

^  ^  H 

XS  ^5  _ 

eS  cS  C 

—  —  ** 

11^ 

5  3  O 


t>      o 


■<*    00   o    o    o 
ec  <M  ^  lo  o 


Eq 


«     ^      +3 

C     o     t- 

W  §  K 


■"     O     o3 
O    ffi    Ph 


e<5    <3^  M 


.1 

s 

s 

cT 

c 
o 
o 
pq 

c 
So    i 

CO      S 

'a 

> 

5s 

> 

2 

o 

2  1 

T3 

"^ 

Tj 

"e 

T3 

<4>      r< 

C 

s 

fi 

<u 

cS 

cS 

d 

cS 

CO 

'3 

3 

o 

(1h 

2. 

o    o 

t^ 


fcii  ffi  ffi      S  K 


-M 


o  &3       o  P5  Ph       Ph  Pm 


o 


a 

'^ 

3 

<3 

fM 

36 


1.    MALONATE 


^ 


CD 

CO 

lO 

U5 

__^ 

05 

OS 

<s 

^^ 

1— 1 

■* 

■ — ■ 

' — ' 

05 

Xi 

Xi 

^ 

X5 

-S 

a 

-^   aj 

_c 

^ 

fli 

CJ 

J« 

73 

-H    73 

^ 

C 

^   c 

^ 

o 

ll 

9^ 

M 

(h 

-ki 

t-i    -p 

0^ 

o 

.£    o 

fc 

§ 

Q  § 

5  2      - 


fcH      2 


W 


-0 

c 

73 

C 

-c 

03 

a; 

01 

« 

-k^i 

TS 

CO 

;h 

c3 

-t-l 

cS 

S 

o 

fin 

o- 

fin 

.S  cc 


i>aooc<iooiMOiio 


:5  ^ 


CO    M    IC    O 
rt    CO  (M 


T»<C^)        -hM:0000<NO        '-h 


TO    r 


-*  o 

CO 

CO    o 

o 

o 

CO    K> 

CO 

-H      lO 

CO 

IC 

PQ 


g  a 


.2       « 


m 


t^ 


05 


Ph    Oh 


O   Oh 


a,  0:5 


&q 


3  -^ 


o3 
C 

o  ^^ 

S  ce 

><  S 

2  -^ 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE  37 


^  lO 


CC    ^         CO 

-  -     -*       -  -       CO       CO        ^"^        ''^        ''^        ^^        '•^        "-^ 


O  OOOOOOOOOO  O  "^^iSkS 


-x      ^       c 


-u        .-. 


c5 

o 

(S 

C 
O 

e3 

c 
o 

eS 

c 

O 

cS 

5= 

o 

o3 

a 

s 

42 

M 

< 

M 

W 

-T  «      S   «      c   ;>> 


5-3 


o  tc  CO 


-^        ^  o 


"'2       -       -_,       ._.       <~- 

_  ..^_  „  -  «  C  O  >)'rt  '^T-l  B-^  J3 


(D 

CS 

C 

CC 

& 

>. 

OJ 

>> 

o 

CS 

y, 

P- 

y. 

o 
Id 

o 

XI 
o 

O 

t3 

-0 

1 

38 


1.    MALONATE 


Table  1-8 
Inhibitor  Constants  for  Dicarboxylate  Ions  on  Succinate  Dehydrogenase 


Source  of  enzyme 

Inliibitor 

Ki  (mM) 

Reference 

Corynebacterium 

Malonate 

0.037 

Strauss  and  Jann 

diphtheriae 

Fumarate 

1.8 

(1956) 

Micrococcus  lactilyticus 

Malonate 

0.23 

Warringa  and  Giix- 

Fumarate 

0.22 

ditta  (1958) 

Claviceps  purpurea 

Malonate 

0.03 

McDonald  et  al. 

(ergot) 

Fumarate 

0.93 

(1963) 

Yeast 

Malonate 

0.0105 

Ryan  and  King 

Fumarate 

1.03 

(1962  a) 

Mytilus  edulis  (mussel) 

Malonate 

0.06 

Ryan  and  King 

Fumarate 

0.15 

(1962  a) 

Rat  kidney 

Oxalacetate 

0.0015 

Pardee  and  Potter 

(1948) 

Acetylene-dicar- 

0.81 

Thomson  (1959) 

boxylate 

Pig  heart 

Malonate 

0.05 

Hellerman  et  al. 

Fumarate 

3.5 

(1960) 

Oxalacetate 

0.0016 

Ethyloxalacetate 

0.04 

Diethyloxalacetate 

No  inh. 

Acetylene-dicar- 

1.4 

boxylate 

ci5-Cyclohexane- 

132 

1,2-dicarboxylate 

Beef  heart 

Malonate 

0.041 

Kearney  (1957) 

Fumarate 

1.9 

Itaconate 

1.8 

Dervartanian  and 

Veeger  (1962) 

(c)  Oxalacetate.  The  marked  inhibition  exerted  by  this  substance  was  first 
noted  by  Das  (1937  b)  and  this  has  been  confirmed  on  the  succinate  dehydro- 
genases from  a  variety  of  organisms.  It  is  at  present  recognized  as  the  most 
potent  succinate  dehydrogenase  inhibitor  among  the  dicarboxylate  ions. 
The  binding  of  the  oxalacetate  to  the  enzyme  would  appear  to  be  at  the 
active  center  because  the  inhibition  is  competitive  with  succinate  (Pardee 
and  Potter,  1948;  Kearney  and  Singer,  1956;  Hellerman  et  al.,  1960),  and 
the  presence  of  oxalacetate  on  the  enzyme  protects  the  active  center  from 
various  sulfhydryl  reagents  (Stoppani  and  Brignone,  1957).  The  reason  for 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE  39 

this  strong  inhibition  is  still  not  clear  but  Hellerman  and  his  group  have 
advanced  some  interesting  speculations.  Oxalacetate  in  aqueous  solution 
may  exist  around  neutrality  in  three  forms  in  equilibrium.  It  was  suggested 
that  the  enolate  form  may  be  the  potent  inhibitor,  especially  the  trans- 

'CX)C  coo"  ^  "OOC^  H 

^c=c'^      -* »^     'ooc-c-CH,— coo"     -* ^       ,c=c^ 

-Q^        ^n  'o^        ^coo" 

cJs  -Enolate  Keto  tautomer  trans  -Enolate 

tautomer  tautomer 

enolate  tautomer  where  the  enolate  group  and  the  carboxylate  group  are 
on  the  same  side.  Evidence  for  this  was  provided  by  showing  that  the 
oxalacetate  monoethyl  ester  is  inhibitory  (about  as  potent  as  malonate) 

EtOCO^  H 

^C  =  C^ 

~o^        ^coo' 

^raws- Enolate 
/  tautomer  of 

monoethyl  ester 

(Table  1-8).  The  diethyl  ester  in  inactive.  Actually  there  is  a  difference  in 
binding  energy  of  about  2.1  kcal/mole  between  oxalacetate  and  either 
malonate  or  the  monoethyl  ester.  This  extra  energy  might  be  due  to  the 
doubly  negative  charge  on  one  end  of  the  molecule.  It  could  well  be  that 
both  the  cis-  and  ^ra ws-enolate  tautomers  are  bound,  the  charge  distri- 
butions being  almost  equivalent  if  one  takes  the  center  of  negative  charge 
as  lying  between  the  carboxylate  and  enolate  groups.  It  is  unfortunate  that 
the  per  cent  of  the  oxalacetate  in  the  enolate  forms  at  pH  7.6  and  30°  in 
aqueous  solution  is  not  known.  Another  possibility  is  that  the  keto  tautomer 
combines  with  the  enzyme,  the  extra  binding  energy  arising  from  an  inter- 
action between  the  keto  group  and  the  enzyme;  a  hydrogen  bond  could  ac- 
count for  the  2.8  kcal/mole  difference  between  the  binding  energies  of  ox- 
alacetate and  succinate.  One  difficulty  in  assuming  the  enolate  form  as  the 
active  inhibitor  is  the  fact  that  maleate  inhibits  very  poorly  and  fumarate 
not  a  great  deal  better.  The  difference  in  binding  energy  between  oxalace- 
tate and  fumarate  for  the  pig  heart  succinate  dehydrogenase  is  close  to 
4.8  kcal/mole,  and  it  would  be  difficult  to  account  for  this  large  difference 
simply  on  the  basis  of  additional  ionic  interactions.  A  final  possibility  must 
be  entertained,  although  no  evidence  for  it  exists,  namely,  the  complexing 
of  the  enolate  group  with  one  of  the  nonheme  iron  atoms  near  the  cationic 
groups  of  the  enzyme  surface,  since  this  could  provide  the  extra  energy 
for  the  binding  of  oxalacetate.  It  should  be  noted  that  although  oxalacetate 
strongly  inhibits  succinate  oxidation  in  rapidly  respiring  liver  mitochondria, 
both  coupled  and  uncoupled,  it  stimulates  when  the  mitochondria  are  in  a 


40  1.    MALONATE 

state  of  respiratory  control  (Kunz,  1963).  Malonate  does  not  stimulate 
under  these  conditions  and  thus  the  effects  of  oxalacetate  on  intact  mito- 
chondrial succinate  oxidation  differ  in  some  manner  from  those  of  malonate. 
This,  however,  is  probably  explained  by  other  reactions  of  oxalacetate,  e.g., 
the  oxidation  of  NADH  to  NAD,  which  would  release  the  mitochondria 
from  respiratory  control. 

{d)  Substituted  malonates  and  succinates.  Franke  (1944  a)  found  that  alkyl- 
malonates  are  inactive  on  heart  succinate  dehydrogenase,  confirming  the 
weak  inhibitions  reported  by  Thunberg  (1933).  Alkylsuccinates  likewise  are 
inactive  until  the  length  of  the  alkyl  chain  is  greater  than  eight  carbon 
atoms.  These  higher  alkylsuccinates  may  exhibit  some  degree  of  competitive 
inhibition  but  there  is  also  inhibition  of  the  oxidation  of  p-phenylenediamine 
and,  hence,  of  the  cytochrome  system.  Therefore,  it  is  clear  that  additions 
of  even  small  alkyl  groups  to  malonate  and  succinate  depress  or  abolish 
the  normal  inhibitory  activity.  Hydroxymalonate  (tartronate)  also  is  es- 
sentially inactive.  These  results  indicate  the  importance  of  that  part  of  the 
molecule  between  the  two  carboxylate  groups  and  would  seem  to  argue 
against  a  simple  interaction  of  the  molecules  with  cationic  groups  on  a 
flat  surface,  the  substituted  groups  protruding  outward.  Of  course,  it  is 
necessary  that  the  — CHgCHo —  grouping  of  succinate  interact  with  the 
enzyme  in  order  for  dehydrogenation  to  take  place.  The  failure  of  short- 
chain  alkylmalonates  to  inhibit  appreciably  must  be  attributed  to  some 
manner  of  steric  interference  by  the  alkyl  groups. 

(e)  Cyclic  dicarhoxylate  ions.  Cyclobutane  and  cyclopentane  dicarboxylates 
are  weak  inhibitors  of  succinate  dehydrogenase.  Since  the  intercarboxylate 
distance  in  cyclobutane- 1,1-dicarboxy late  and  CT's-cyclopentane-l,2-dicar- 
boxylate  are  not  too  far  from  that  in  malonate  (Table  1-1),  the  poor  inhibi- 
tion may  be  due  to  the  bulkiness  of  the  rings  interfering  sterically,  as  do  the 
alkyl  groiips  discussed  in  the  previous  section.  It  is  strange,  however,  that 
there  is  no  inhibitory  difference  between  the  cis  and  trans  isomers  of  cyclo- 
pentane-1,2-dicarboxylate,  since  the  intercarboxylate  distances  differ  by 
1.39  A.  Thus  the  inhibition  may  not  be  by  the  same  mechanism  as  for  mal- 
onate; indeed,  cw-cyclohexane-l,2-dicarboxylate  inhibits  succinate  dehy- 
drogenase noncompetitively  (Hellerman  et  al.,  1960). 

Nature  of  the  Active  Center  and   the   Binding  of  Malonate 

The  evidence  indicates  the  presence  at  the  active  center  of  two  cationic 
groups  and  a  nearby  SH  group.  The  cationic  groups,  perhaps  3-4  A  apart, 
are  suggested  by  the  very  weak  inhibitions  exerted  by  monocarboxylates 
(Quastel  and  Wooldridge,  1928;  Dietrich  et  al.,  1952)  and  the  complete 
lack  of  a  competitive  inhibition  by  compounds  in  which  the  negative 
charges  on  the  carboxylate  groups  are  eliminated.  Malondialdehyde  (Holt- 


INHIBITION    OF    SUCCINATE    DEHYDKOGENASE  41 

kamp  and  Hill,  1951),  malondiamide  (Fawaz  and  Fawaz,  1954),  and  var- 
ious derivatives  of  succinate,  in  which  one  or  both  of  the  carboxylate 
groups  have  been  replaced  with  nonionic  groups  (Dietrich  et  al.,  1952), 


/COO 

/OH 

^CHO 

CONH, 

H,C 

H-C 

H^C^ 

H,C^ 

1 

1 

^CHO 

CONHj 

"^^-Br 

"^*^^CN 

Malondialdehyde  Malondiamide  /3-Bromopropionate     /3- Hydroxy  propiononitrile 

are  all  lacking  in  inhibitory  activity.  An  SH  group  close  to  the  cationic 
attachment  points  is  proved  by  the  high  sensitivity  of  the  dehydrogenase  to 
substances  reacting  with  SH  groups  and  the  protection  that  malonate  af- 
fords against  such  substances.  The  latter  is  actually  better  evidence  for  the 
proximity  of  the  SH  group  because  sulfhydryl  reagents  can  alter  protein 
structure  and  exert  effects  for  some  distance  over  the  enzyme,  whereas  the 
blockade  of  these  substances  by  a  small  molecule  such  as  malonate  would 
be  almost  certain  proof.  Hopkins  et  al.  (1938)  showed  that  malonate,  can 
protect  succinate  dehydrogenase  from  oxidized  glutathione  (GSSG),  which 
is  a  potent  inhibitor.  Incubation  of  the  enzyme  with  GSSG  increased  the 
methylene  blue  reduction  time  from  10  min  to  3  hr;  malonate  at  concentra- 
tions from  0.2  to  100  mM  gave  almost  complete  protection.  That  the 
SH  groups  are  protected  by  malonate  was  also  shown  by  titration  of  these 
groups  with  iodine.  Potter  and  DuBois  (1943)  reported  protection  against 
quinone  inhibition  and  Barron  and  Singer  (1945)  against  arsenicals.  It  is 
interesting  that  malonate  also  protects  against  oxygen  poisoning  of  suc- 
cinate oxidase,  which  was  interpreted  to  mean  that  the  SH  groups  of  the 
enzyme  are  involved  in  this  inactivation  (Dickens,  1946  b).  Oxalacetate 
has  also  been  shown  to  be  protective  against  mercurials  and  arsenicals 
(Stoppani  and  Brignone,  1957),  providing  evidence  that  it  binds  to  the 
same  site  as  succinate  and  malonate.  The  degree  of  protection  depends  on 
several  factors,  including  the  concentrations  of  the  inhibitors  and  the  rates 
at  which  they  react  with  the  enzyme;  the  less  protection  seen  against  the 
mercurials  is  attributed  to  their  comparatively  rapid  action  whereas  the 


Inhibitor 

Oxalacetate 

Inhibitor 

concentration 
{mM) 

concentration 
(mM) 

0/ 

/o 

Protection 

jj-MB 

0.5 

1.3 

21.8 

p-Chloromercuriphenol 

0.76 

3.4 

24.0 

HgCl, 

0.38 

3.4 

53.8 

Methylarsenoxide 

2.0 

3.4 

67.2 

Oxophenarsine 

0.36 

1.3 

81.2 

42  1.    MALONATE 

arsenicals  require  at  least  30  min  to  reach  their  maximal  inhibition.  Such 
protection  on  a  competitive  basis  was  called  interference  inhibition  by 
Ackermann  and  Potter  (1949).  Once  the  sulfhydryl  reagents  have  reacted 
with  the  enzyme,  malonate  will  not  reverse  the  inhibition,  but  only  slows 
down  the  rate  at  which  the  substance  acts  on  the  enzyme.  (I  find  it  difficult 
to  understand  how,  in  some  cases,  such  low  concentrations  of  malonate 
afford  protection  against  irreversible  inhibitors.  For  example.  Potter  and 
DuBois  reported  that  0.33  mM  malonate  protects  quite  well  against 
p-quinone,  and  yet  this  concentration  of  malonate  inhibits  only  around 
20%,  showing  most  of  the  enzyme  uncombined  with  malonate). 

Succinate  dehydrogenase  also  contains  nonheme  iron  and  flavin  dinucleo- 
tide  but  the  locations  of  these  components  relative  to  the  succinate-binding 
site  are  not  known.  Since  the  iron  and  the  flavin  both  participate  in  the 
electron  transfer,  it  is  reasonable  to  assume  that  at  least  one  of  them  is 
close,  or  even  part  of  the  active  center.  Most  formulations  have  pictured 
the  initial  step  as  a  transfer  of  hydrogen  atoms  from  succinate  to  the  fla- 
vin; if  this  is  so,  the  topography  of  the  active  center  must  be  rather 
complex. 

We  shall  now  turn  to  the  energetics  of  the  binding  of  malonate  in  order 
to  determine  if  the  ionic  interactions  generally  assumed  are  reasonable. 
An  immediate  difficulty  is  the  variability  in  the  values  of  K^  reported,  even 
for  the  same  tissue;  for  example,  0.0076  mM  (Thorn,  1953)  and  0.05  milf 
(Hellerman  et  al.,  1960)  for  the  enzyme  from  pig  heart,  and  0.041  mM  for 
beef  heart  (Kearney,  1957).  A  K^  of  0.0076  mM  would  indicate  an  over  all 
binding  energy  of  7.24  kcal/mole,  or  3.62  kcal/mole  for  each  carboxylate 
interaction  assuming  only  ion-ion  contribution.  This  is  a  reasonably  high 
value  for  the  interaction  of  C00~  and  NH3+  groups  and  corresponds  to 
an  intercharge  separation  of  around  4.30  A  (Fig.  1-6-16),  which  is  not  far 
from  contact  of  the  groups.  It  is  unlikely  that  other  types  of  interaction 
are  important  for  malonate.  The  corresponding  K,  for  succinate  is  0.0028 
mM,  giving  an  interaction  energy  of  3.28  kcal/mole  per  carboxylate  group 
and  a  separation  of  4.45  A.  A  difference  of  fit  of  0.15  A  would  thus  account 
for  the  relative  bindings  of  malonate  and  succinate.  However,  in  the  case 
of  succinate  it  is  more  likely  that  other  energy  terms  are  involved.  There 
is  undoubtedly  some  interaction  between  the  — CH2CH2 —  region  and  the 
enzyme,  and  it  is  probable  that  distortion  of  the  succinate  molecule  occurs 
upon  binding.  In  any  event,  these  rough  estimates  point  to  a  fairly  close 
fit  of  malonate  to  the  enzyme  cationic  groups  for  the  pig  heart  enzyme. 
The  ability  of  malonate  to  bind  to  the  active  center  of  succinate  dehydro- 
genases from  bacteria  is  apparently  much  less  (Table  1-8). 

The  next  question  is:  what  is  the  most  probable  distance  between  the 
two  enzyme  cationic  binding  sites?  Information  on  this  must  be  obtained 
from  the  relative  bindings  of  substances  having  negatively  charged  groups 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE  43 

different  distances  apart.  Since  malonate  is  usually  bound  more  tightly 
than  succinate,  and  much  more  than  oxalate,  it  is  reasonable  to  assume  an 
intercationic  distance  approximating  the  intercarboxylate  distance  in  mal- 
onate. It  is  by  no  means  necessary  that  the  substrate  in  its  free  configu- 
ration exactly  fits  the  enzyme  site.  Pauling  (1946,  1948)  has  suggested, 
"  an  active  region  of  the  surface  of  the  enzyme...  is  closely  complementary 
in  structure  not  to  the  substrate  molecule  itself,  in  its  normal  configuration, 
but  rather  to  the  substrate  molecule  in  a  strained  configuration,  correspond- 
ing to  the  activated  complex  for  the  reaction  catalyzed  by  the  enzyme." 
Now,  the  fact  that  malonate  fits  the  active  site  well  does  not  mean  that  the 
enzyme  cationic  groups  are  the  same  distance  apart  as  the  carboxylate 
groups  (3.28  A).  The  calculations  above  indicate  a  distance  of  4.3  A 
between  carboxylate  and  cationic  groups  and  thus,  depending  on  the 
geometry  of  the  binding,  the  cationic  groups  could  be  much  farther  apart 
than  3.28  A.  Extreme  situations  are  shown  in  Fig.  1-7,  where  the  intercat- 


FiG.  1-7.  Representations  for  the  ex- 
treme situa/tions  in  the  interaction  of 
malonate  with  the  two  cationic  groups 
on  the  surface  of  succinate  dehydro- 
genase. In  both  cases  (A  and  B)  the 
interaction  distances  and  the  energies 
between  the  ionic  groups  are  the  same. 

ionic  distance  may  vary  from  3.28  to  13.1  A,  approximately  the  same  energy 
of  binding  being  expected  in  either  case.  Situation  A  is  not  very  likely 
because  it  is  improbable  that  protein  cationic  groups  would  occur  so  close, 
and,  furthermore,  in  this  case  oxalate  might  be  expected  to  bind  quite  well. 
Also,  situation  B  would  provide  more  opportunity  for  succinate  to  be  dehy- 
drogenated  at  the  enzyme  surface.  Of  course,  the  enzyme  surface  at  the 


44  1.    MALONATE 

active  center  may  not  be  smoothly  curved  as  shown,  and  we  shall  soon 
examine  evidence  that  it  is  not. 

The  problem  of  the  interactions  of  di-ionic  substances  with  receptor 
groups  has  been  recently  treated  by  Schueler  (1960,  p.  448)  and  on  the  basis 
of  statistical  calculations  he  has  concluded,  "The  most  dramatic  alteration 
in  activity  should  occur  upon  approaching  that  agent  in  the  series  which 
possesses  a  length  distribution  just  capable  of  overlapping  the  negative- 
charge  spacing  in  the  receptor,  and  this  should  be  followed  by  a  relatively 
slow  rate  of  loss  in  activity  with  respect  to  increasing  length"  (he  is  assum- 
ing a  positively  charged  drug).  In  view  of  what  has  just  been  said  above, 
there  is  some  doubt  if  predictions  like  this  can  be  made  with  confidence. 
The  large  distances  between  the  interacting  charged  groups  make  it  very 
difficult  to  assign  receptor  configuration  and  much  will  depend  on  the  over 
all  configuration  of  the  protein  surface.  Other  factors,  such  as  distortion  of 
long-chain  molecules,  interactions  of  regions  between  the  end  charged 
groups  the  surface,  and  possible  steric  repulsions,  must  be  considered. 

Inasmuch  as  little  accurate  information  on  the  intercationic  distance  can 
be  obtained  from  Kj  and  K^.  values  alone,  let  us  now  turn  to  more  profitable 
considerations  of  the  topography  of  the  active  center.  There  is  evidence 
from  several  lines  that  the  active  center  is  not  a  flat  or  slightly  convex 
surface.  In  the  first  place,  alkylsuccinates  are  bound  to  the  enzyme  very 
poorly;  methylsuccinate  is  oxidized  at  23%  the  rate  for  succinate,  and 
ethylsuccinate  at  18%  the  rate  for  succinate,  while  higher  members  are 
neither  oxidized  nor  are  they  inhibitory  (Franke,  1944  a).  In  the  second 
place,  as  we  have  already  seen,  alkylmalonates  are  very  poor  inhibitors. 
Indeed,  even  the  introduction  of  a  hydroxyl  group  (tartronate)  reduces  the 
inhibition  markedly.  These  observations  indicate  a  rather  close  fit  for 
malonate  and  succinate  at  the  active  center,  additional  groups  giving  rise 
to  steric  repulsion,  as  frequently  reported  for  antigen-antibody  reactions. 
In  the  third  place,  fumarate  is  bound  fairly  well  while  maleate  is  not,  indi- 
cating again  some  steric  repulsion  since  the  intercarboxylate  distances 
alone  would  certainly  allow  predictions  that  maleate  would  be  bound  more 
tightly.  All  molecules  that  bind  appreciably  to  succinate  dehydrogenase 
seem  to  be  simple  linear  substances,  or  substances  capable  of  assuming  a 
linear  configuration.  All  of  this  evidence  points  to  a  slit  or  tubular  structure 
for  the  active  center,  such  that  compounds  with  added  groups  or  rigid  non- 
linear molecules  cannot  enter.  Such  a  situation  is  pictured  in  Fig.  1-8  in 
two  dimensions.  This  is  not  to  be  construed  as  an  attempt  to  represent 
the  actual  configuration  but  merely  to  show  the  steric  barriers  impeding 
attachment  of  larger  or  nonlinear  molecules.  Glutarate  and  adipate  could 
not  fit  well,  not  because  of  unsatisfactory  intercarboxylate  distances,  but 
because  of  the  bulkiness  of  the  longer  hydrocarbon  chains.  Fumarate  is 
able  to  bind  because  its  configuration  is  much  like  that  of  succinate  in  the 


INHIBITION    OF    SUCCINATE    DEHYDROGENASE 


45 


extended  form  shown  in  the  figure,  whereas  maleate  might  not  fit  because 
of  its  nonlinear  structure.  Acetylene-dicarboxylate  would  be  expected  to  in- 
hibit to  some  extent  because  of  its  linearity.  Such  a  model  would  also  ex- 
plain why  small  alkyl  groups  added  to  succinate  do  not  completely  abolish 
the  binding.  It  may  also  be  mentioned  that  this  type  of  configuration  would 
allow  the  flavin  and  iron  components  of  the  dehydrogenase  to  be  in  posi- 
tions close  to  the  — CH2CH2 —  group  and  thus  able  to  participate  in  the 
removal  of  the  hydrogen  atoms. 


Succinate 


Fig.  1-8.  Representations  of  the  binding 
of  malonate  and  succinate  at  the  active 
site  of  succinate  dehydrogenase,  indicat- 
ing the  steric  barriers  possibly  surround- 
ing the  region  of  the  two  cationic  sites. 
The  actual  situation  must  be  visualized 
in  three  dimensions. 


Activation   of  Succinate   Dehydrogenase   by   Malonate 

Preparations  of  beef  heart  succinate  oxidase  obtained  using  borate  buf- 
fer are  not  fully  active  but  may  be  activated  by  the  addition  of  phosphate. 
This  interesting  discovery  by  Kearney  (1957,  1958)  may  have  important 
bearings  on  the  understanding  of  the  active  center  of  this  enzyme,  especially 
as  she  later  found  that  succinate,  fumarate,  and  malonate  also  activate, 
and  indeed  are  much  more  potent  than  phosphate  (see  accompanying 
tabulation).  Once  the  enzyme  has  been  activated,  the  activator  can  be 
removed  without  loss  of  the  activity;  in  fact,  malonate  must  be  dialyzed 
away  if  the  full  activity  of  the  enzyme  is  to  be  measured.  The  activation 
constants  are  quite  different  from  the  Michaelis  or  inhibitor  constants  for 
these  substances.  It  would  appear  that  malonate  binds  more  tightly  to  the 
less  active  form  of  the  enzyme.  Kearney  favors  the  view  that  these  activa- 


46  1.    MALONATE 

tors  convert  a  less  active  form  of  the  enzyme  to  a  more  active  form,  and 
that  this  transformation  may  involve  a  localized  change  in  the  protein 


Activator 


(mJf)  (mM) 


Malonate 

0.0072 

0.025 

Succinate 

0.12 

0.52 

Fumarate 

5.6 

0.80 

Phosphate 

100 

— 

structure  because  of  the  high  energy  of  activation.  Whatever  the  explanation 
it  is  important  to  remember  that  malonate  can  exert  an  activating  effect, 
as  weU  as  inhibiting,  in  media  low  in  phosphate. 

Effects  of  Various  Factors  on  the   Inhibition   of  Succinate   Dehydrogenase 

Very  few  illuminating  studies  on  the  modification  of  malonate  inhibition 
are  available.  The  effects  of  temperature  were  mentioned  in  Volume  I,  where 
the  following  thermodynamic  parameters  for  the  inhibition  of  beef  heart 
succinate  dehydrogenase  at  38°  were  calculated:  AF  =  —  6.26  kcal/mole, 
AH  =  —  5.48  kcal/mole,  and  AS  =  2.6  cal/mole/degree.  The  K^^  for  malonate 
was  found  to  be  0.025  mM  at  20o-23o  and  0.041  mM  at  38o  (Kearney,  1957). 
The  effects  of  osmolarity  on  the  inhibition  are  surprisingly  large  (Honda  and 
Muenster,  1961).  Lupine  mitochondria  were  prepared  in  media  of  different 
sucrose  concentrations  and  assayed  in  media  of  two  osmolarities  (Table 
1-9).  These  results  were  obtamed  on  mitochondria  and  it  is  possible  that 
the  effects  are  not  directly  on  succinate  oxidase  but  on  the  permeability 
or  structural  properties  of  the  mitochondria.  In  this  connection  it  has  been 
pointed  out  by  Singer  and  Lusty  (1960)  that  iV-methylphenazine  measures 
the  full  activity  of  succinate  dehydrogenase  in  mitochondrial  fragments, 
but  in  intact  mitochondria  it  measures  only  a  fraction  of  the  activity. 
Various  ways  of  damaging  the  mitochondria  lead  to  increased  succinate 
dehydrogenase  activity  (such  as  increase  in  Ca++  concentration).  This  was 
interpreted  in  terms  of  permeability  barriers  to  A^-methylphenazine  (and 
also  FMNHj),  limiting  the  rate  in  intact  mitochondria.  It  could  also  be 
explained  on  the  basis  of  structural  changes  in  the  enzyme  complexes.  In 
any  case,  these  observations  demonstrate  the  importance  of  the  mitochon- 
drial state  in  the  functioning  of  succinate  oxidase,  and  it  would  not  be 
surprising  if  malonate  inhibition  were  similarly  sensitive.  The  inhibition 
of  succinate  oxidation  by  malonate  in  rat  heart  mitochondria  in  KCl  me- 
dium is  not  altered  by  either  halving  or  doubling  the  KCl  concentration  in 
the  assay  medium  (Montgomery  and  Webb,  1956  b).  However,  the  effects 


INHIBITION    or    SUCCINATE    DEHYDROGENASE 


47 


Table  1-9 

Effect  of  Osmolarity  in   the  Properties  of  Succinate   Oxidase  in  Lupine 

Mitochondria  " 


Preparation 

Assay 

K 

K- 

osmolarity 

osmolarity 
(M) 

(mM) 

(mM) 

KJKi 

a 

0.15 

0.22 

5.10 

0.91 

5.6 

2 

0.60 

12.34 

0.64 

19.3 

1 

0.40 

0.22 

2.87 

0.19 

15.1 

8 

0.60 

5.47 

0.16 

34.2 

191 

0.60 

0.22 

1.20 

0.05 

24.4 

36 

0.60 

5.47 

0.11 

49.7 

268 

"  Kj„  is  the  Michaelis  constant  for  succinate,  iT,  the  inhibitor  constant  for  malonate, 
and  a  is  the  interaction  constant  defined  in  Eqs.  1-3-5  and  1-3-6,  indicating  the  type 
of  inhibition.  The  osmolarity  is  given  in  terms  of  sucrose  concentration.  (From  Honda 
and  Muenster,  1961.) 

of  Ca'^^  concentration  on  both  succinate  oxidation  and  malonate  inhibition 
are  complex  (Fig.  1-9)  and  difficult  to  explain.  The  decrease  in  the  rate 
of  oxidation  beyond  Ca+"^  concentrations  around  5  m.M  might  be  attributed 
to  a  complexing  of  the  succinate,  but  for  the  same  reason,  namely,  the 
complexing  of  malonate,  which  has  a  higher  affinity  for  Ca+''"  than  does 
succinate,   the  inhibition   would  be  expected  to  decrease.  Whether  the 


0.001  0  01 


Fig.  1-9.  Effects  of  Ca++  on  the  rate  of  succinate  oxidation  and 
the  inhibition  by  malonate  in  rat  heart  mitochondria.  Succinate  is 
5  mM  and  malonate  is  1  mM.  (From  Montgomery  and  Webb,  1956  b). 


48  1.    MALONATE 

modest  stimulation  of  succinate  oxidation  by  low  concentrations  of  Ca^^ 
is  a  permeability  or  structure-opening  effect  on  the  mitochondria,  or  due 
to  a  more  direct  effect  on  the  enzyme,  is  not  known;  certainly  the  oxidations 
of  other  cycle  substrates  and  pyruvate  are  strongly  depressed  by  Ca++. 
The  effects  of  Mg++  concentratioti  from  6.2  to  12.4  mM  on  trypanosomal 
succinate  dehydrogenase  inhibition  by  malonate  have  been  reported  as 
negligible  (Agosin  and  von  Brand,  1955).  In  connection  with  the  relation- 
ship between  mitochondrial  structure  and  malonate  inhibition,  it  is  interest- 
ing to  examine  the  effect  of  ATP,  inasmuch  as  ATP  protects  or  stabilizes 
mitochondria  after  isolation  from  cells.  ATP  at  1  mM  has  very  little  effect 
on  the  rate  of  succinate  oxidation  in  rat  heart  mitochondria  (in  five  exper- 
iments a  mean  depression  of  3%),  but  in  the  presence  of  succinate  (5 
mM)  and  malonate  (5  mM)  it  stimulated  the  rate  67%,  thus  antagonizing 
the  inhibition  by  malonate  (Montgomery  and  Webb,  1956  b).  Similar  re- 
sults were  seen  with  acetylene-dicarboxylate  inhibition.  If  the  effect  of 
ATP  is  to  reduce  the  permeability  to  these  inhibitors,  it  is  surprising  that 
interference  with  succinate  penetration  does  not  also  occur.  Thus  some 
other  explanation  may  have  to  be  sought. 

Inhibition   of  Fumarate   Reduction 

If  succinate,  fumarate,  and  malonate  bind  at  the  same  site  on  succinate 
dehydrogenase,  the  reverse  reaction  —  the  hydrogenation  of  fumarate  to 
succinate  —  should  be  inhibited  by  malonate;  that  is,  malonate  should 
compete  with  fumarate  as  well  as  with  succinate.  The  relative  potencies 
of  the  inhibitions  on  the  two  reactions  would  depend  on  the  Michaelis 
constants  for  succinate  and  fumarate,  so  that  the  inhibitions  would  not 
necessarily  be  identical.  Malonate  was,  indeed,  found  to  inhibit  the  reduction 
of  fumarate  by  horse  muscle  succinate  dehydrogenase,  but  less  potently 
than  the  oxidation  of  succinate  (18  mM  malonate  required  to  inhibit  50% 
in  the  former  case  and  3.6  mM  in  the  latter)  (Das,  1937  b).  A  more  detailed 
analysis  was  made  by  Forssman  (1941)  in  Lund,  using  pig  heart  succinate 
dehydrogenase  and  leucomethylene  blue  as  a  hydrogen  donor,  in  this  case 
the  inhibition  by  malonate  being  essentially  equivalent  for  both  forward 
and  backward  reactions.  More  recent  studies  of  beef  heart  succinate  dehy- 
drogenase, however,  have  given  different  values  for  K,:  0.025  mM  for  suc- 
cinate oxidation  and  0.12  mM  for  fumarate  reduction  (Singer  et  al.,  1956  a). 
This  difference  is  unexpected  and  the  suggestion  was  made  that  the  binding 
of  malonate  may  be  effected  by  the  state  of  oxidation  of  the  electron  trans- 
port components  adjacent  to  the  binding  site.  The  affinities  of  fumarate 
for  the  enzyme  were  also  found  to  be  different  for  each  reaction.  It  may  also 
be  of  significance  that  iV-methylphenazine  was  the  acceptor  in  the  oxida- 
tion of  succinate,  and  FMNH,  or  leucodiethylsafranin  the  donor  in  the  re- 
duction of  fumarate. 


INHIBITION    OF    SUCCINATE    DEHYDKOGENASE  49 

The  succinate  dehydrogenase  of  Micrococcus  lactilyticus  behaves  quite 
differently  and  is  poorly  inhibited  by  malonate,  a  ratio  of  (malonate)/(fu- 
marate)  =  20  being  required  for  29%  inhibition  (Peck  et  al,  1957).  It  might 
be  thought  that  this  enzyme  is  not  succinate  dehydrogenase,  but  another 
enzyme  that  could  be  called  "  fumarate  reductase,"  especially  as  fumarate 
is  reduced  at  a  faster  rate  than  succinate  is  oxidized,  in  contrast  to  the 
mammalian  enzymes.  However,  it  has  been  conclusively  demonstrated 
it  is  not  a  separate  enzyme  and  that  the  failure  of  malonate  to  inhibit 
is  to  be  attributed  to  a  very  high  affinity  for  fumarate  coupled  with  a  rela- 
tively low  affinity  for  malonate  (Warringa  et  al.,  1958).  The  configuration 
of  the  active  center  of  the  bacterial  enzyme  must  differ  from  that  of  the 
mammalian  enzymes.  This  may  also  explain  the  "fumarate  reductases" 
obtained  from  yeast  (Fischer  and  Eysenbach,  1937;  Kovac,  1960)  which  are 
rather  insensitive  to  malonate. 

Variations   of  Malonate   Inhibition   of  Succinate   Dehydrogenases 
from  Different  Tissues  and  Species 

The  comparative  biochemistry  of  enzyme  inhibition  is  in  its  infancy  and 
accurate  comparison  of  results  is  usually  impossible  due  to  the  different 
conditions  under  w^hich  the  inhibitions  w^ere  studied.  Examination  of  Table 
1-6  with  a  view  to  establishing  phylogenetic  relationships  is  made  difficult 
by  the  different  types  of  preparation  and  assay  procedure  used.  A  correla- 
tion graph,  made  by  plotting  inhibitions  against  (I)/(S)  ratios,  shows  a  very 
marked  scatter  of  the  points.  For  example,  at  an  (I)/(S)  ratio  of  0.1.  the 
inhibitions  range  from  10%  to  100%.  This  variation  cannot  all  be  due  to 
the  differences  in  technique.  All  of  those  cases  in  which  the  malonate  inhibi- 
tion is  significantly  below  the  mean  turn  out  to  be  in  the  bacteria,  inverte- 
brates, or  plants.  However,  this  is  not  a  strict  correlation  because  some  of  the 
potent  inhibitions  have  been  found  in  such  organisms  {e.g.,  Azotobacter,  E. 
colt,  and  Trypanosoma).  The  possibility  of  a  relationship  between  succinate 
dehydrogenase  type  in  the  bacteria  and  the  oxygen  requirements  for  growth 
has  been  proposed.  The  enzyme  from  the  obligate  anaerobe  Micrococcus 
lactilyticus  has  low  affinities  for  succinate  and  malonate,  as  discussed  in  the 
previous  section,  whereas  the  enzyme  from  the  facultative  anaerobe, 
Propionibacterium  pentosaceum  is  intermediate  in  properties  between  Mi- 
crococcus and  the  aerobic  mammalian  tissues  (Singer  and  Lara,  1958).  Cer- 
tainly many  invertebrate  and  plant  tissues  can  withstand  anaerobiosis 
better  than  mammalian  tissues,  but  at  the  present  state  of  our  knowledge 
such  a  correlation  is  dangerous  to  make. 

Comparisons  of  the  malonate  inhibitions  of  succinate  dehydrogenases 
from  different  tissues  have  been  reported  in  a  few  cases.  The  epithelium 
and  muscle  of  guinea  pig  seminal  vesicle  were  separated  and  the  inhibitions 
by  malonate  at  four  different  concentrations  were  similar  (Levey  and  Szego, 


50  1.    MALONATE 

1955).  However,  the  inhibitions  of  epithelial  dehydrogenase  appear  to  be 
about  5%  higher  than  for  the  muscle  enzyme,  although  this  may  not  be 
statistically  significant.  Malonate  was  found  to  inhibit  Hepatoma  134 
tumor  succinate  dehydrogenase  more  than  the  enzyme  from  normal  mouse 
liver  (Fishgold,  1957)  over  a  range  of  five  malonate  concentrations;  for 
example,  0.21  mM  malonate  inhibited  the  liver  enzyme  42%  and  the  he- 
patoma enzyme  76%  at  a  succinate  concentration  of  1  vaM.  Killer  and 
sensitive  stocks  of  paramecia  may  have  succinate  dehydrogenases  with 
different  sensitivities  to  malonate,  but  it  is  difficult  to  draw  conclusions 
from  the  data  published  (Simonsen  and  van  Wagtendonk,  1956).  The  Og 
uptake  of  homogenates  of  the  two  strains  was  increased  to  different  de- 
grees by  50  mM  succinate  and  malonate  inhibited  both  quite  well  (see 
accompanying  tabulation).  The  authors  concluded  that  the  enzyme  from 
the  killer  strain,  is  inhibited  more,  based  on  absolute  reduction,  but  actually 

Increase  in  Oj  uptake  from  succinate 


Sensitive  strain  Killer  strain 


Control  1.4  14.0 

With  malonate  80  mM  0.2  3.8 

%  Inhibition  86  73 

the  enzyme  from  the  sensitive  strain  seems  to  be  inhibited  as  w^ell.  Our 
work  with  succinate  dehydrogenase  from  various  rat  tissues  (Table  1-6) 
indicates  no  significant  difference  in  susceptibility  to  malonate. 

INHIBITION    OF  SUCCINATE   OXIDATION 
IN   CELLULAR  PREPARATIONS 

Attention  will  now  be  turned  to  the  inhibition  of  the  succinate  oxidase 
system  when  it  is  located  in  the  normal  cellular  structure  and  succinate  is 
added  exogenously  to  the  preparations.  When  succinate  is  added  to  most 
cell  suspensions,  minces,  or  slices,  there  is  an  increase  in  the  O2  uptake, 
and  this  response  is  inhibited  to  varying  degrees  by  malonate  (Table  1-10). 
It  is  particularly  important  in  cellular  preparations  to  take  account  of  the 
endogenous  respiration  and  the  effect  of  malonate  on  it  (see  Chapter  1-9). 
In  many  studies  this  has  not  been  done  and  this  is  one  factor  that  makes 
it  difficult  to  compare  accurately  the  malonate  inhibitions  m  vitro  and  in 
vivo.  Since  the  endogenous  respiration  is  generally  inhibited  less  than 
succinate  oxidation  by  malonate,  failure  to  correct  for  endogenous  respi- 
ration usually  leads  to  low  values  for  the  inhibition.  This  is  illustrated  in 


INHIBITION  OF  SUCCINATE   OXIDATION  51 

the  four  examples  given  in  Table  1-11.  The  importance  of  the  endogenous 
correction  is  seen  to  vary  with  the  effect  of  malonate  on  the  endogenous 
respiration.  When  malonate  inhibits  the  endogenous  O2  uptake  poorly, 
as  in  rat  liver,  or  actually  stimulates  the  O2  uptake  (due  to  its  metabolism), 
as  in  Euglena,  the  true  inhibition  of  succinate  oxidation  is  much  higher  than 
would  be  calculated  simply  from  the  data  on  succinate  and  succinate  + 
malonate. 

Comparison  of  the  inhibitions  in  Tables  1-6  and  1-10  for  the  same  species 
or  tissues,  although  this  is  qualitative  only,  shows  that  in  several  instances 
the  inhibitory  potency  of  malonate  seems  to  be  significantly  less  in  cellular 
preparations.  This  is  true  for  E.  coli,  Rhodospirilliim,  Crithidia,  Zygor- 
rhynchus,  pigeon  muscle,  and  rat  liver.  Moses  (1955)  points  out  that  in 
Zygorrhynchus  the  inhibition  of  succinate  oxidation  in  cell  suspensions  is 
very  weak,  even  at  the  low  pH  of  3.4,  but  when  the  cells  are  treated  with 
liquid  nitrogen  to  destroy  their  structure,  malonate  inhibits  normally. 
The  oxidation  of  succinate  by  cell  susj^ensions  of  Bacterium  succinicum  is 
not  inhibited  at  all  by  5  mM  malonate  whereas  such  oxidation  in  cell-free 
extracts  is  inhibited  completely  (Takahashi  and  Nomura,  1952).  There  are 
also  several  reports  in  which  malonate  was  found  to  be  ineffective  but  the 
extracted  succinate  oxidase  system  was  not  directly  tested;  however,  in 
these  cases  one  would  certainly  expect  the  enzyme  to  be  sensitive  to  mal- 
onate. For  example,  malonate  (2  mM)  does  not  inhibit  the  oxidation  of 
succinate  by  barley  roots  (Honda,  1957),  or  at  5-20  mM  in  dried  cells  of 
CJdorella  (Millbank,  1957),  while  in  beech  roots  malonate  (28.6  mM)  actually 
stimulates  the  rate  of  succinate  oxidation  (Harley  and  Ap  Rees,  1959). 
However,  in  some  cases  comparable  inhibitions  have  been  observed  in  vitro 
and  in  vivo.  Danforth  (1953)  showed  in  Euglena  that  malonate  is  quite  ef- 
fective in  intact  cells  if  the  pH  is  low  enough  (around  4.5)  and,  although  it 
is  difficult  to  compare  the  results  with  those  obtained  from  homogenates 
because  of  different  concentrations  of  succinate  and  malonate,  it  would 
appear  that  malonate  is  equally  inhibitory  in  the  two  preparations.  Similar 
effects  of  malonate  were  also  observed  in  our  work  (Montgomery  and  Webb, 
1956  b)  on  rat  heart  slices  and  mitochondrial  suspensions. 

The  failure  of  malonate  to  inhibit  the  oxidation  of  added  succinate  in 
cellular  preparations  well,  or  at  all,  has  usually  been  attributed  to  permea- 
bility factors.  However,  it  is  difficult  to  understand  how  permeability  could 
explain  these  results,  inasmuch  as  the  penetration  of  both  succinate  and  mal- 
onate would  be  controlled  by  the  same  factors,  presumably.  That  is,  if  there 
is  some  barrier  to  malonate  reaching  the  succinate  dehydrogenase,  how  can 
succinate  pass  this  barrier?  It  is  true  that  the  "pK,,  for  succinate  is  higher 
than  for  malonate  but  almost  identical  values  of  ^K,,  have  been  reported; 
thus  the  distribution  of  the  ionic  forms  around  neutrality  would  be  approxi- 
mately the  same  (Table  1-3).  If  only  the  uncharged  forms  of  these  acids  — 


52 


1.    MALONATE 


« 


<i) 


13    — 


m 


3 


3 


X 

-C 

^ 

^ 

T3 

T3 

,^ 

C 

c 

CD 

cS 

cS 

05 

S 

"a; 

^-' 

5S 

tc 

T3 

e8 

eS 

S 

S 

^ 

o- 

a 

t-i 

.S  '^ 


O    —<    -*    CD 
CO    »c    t^ 


i? 


r^  o        --<  o  ic  o 
CD  '—I  -H  irj   ic 


—        H 


o 

O 

_o 

_o 

M 

CO 

'S 

'« 

c 

C2 

C 

« 

£ 

5 

OJ 

ft 

ft 

ft 

ft 

ai 

t» 

m 

CO 

M   W         MM 


S     S 


5    e 


B3 


q; 


CQ 


INHIBITION  OF  SUCCINATE  OXIDATION 


53 


o  o  o 
'^  2  S 

I— I    CO 


o  o     o 


P5 


- — • 

:3 

^"^ 

ft 

o 

'e 

^ 

"e 

-< 

« 

'«3 

T3 

c 

>« 

c 
o 

eS 

02 

c3 

cS 

to 

_c 

>■. 

cS 

tn 

cS 

p 

eS 

cS 

'3 

O 

m 

W 

ffi 

§ 

m  oi 

o 

o 

GO 

O 

oo 

o   >o   IC 

iM    CO 

00 

IM 

t> 

o 

M 

Tt<    O    00 

a^ 


eo  CD  o 


CQ 


2Q 


I— I        '^ 


c3 


0)     o 


0)   .^ 


4)      <» 


ft     ft 
O      O 

'o     _ 


I 


ft 
ft 


ft 
02 


s 

s 
■a 

1 

1 

6q 

Q. 

54 


1.    MALONATE 


tf 


CM 


P^      P5 


W 


Oi    CO    -H    Oi    o 

t^  i>  !>  CO  ;d 


-o 

73 

^ 

C5 

73 

^ 

c 

C 

^ 

^_^ 

C 

<u 

cS 

c3 

Q) 

c3 

s 

s 

o 

'? 

4^ 
0^ 

CO 

1 

to 

2 

O 
1-5 

IS 

-* 

(M 

CO 

t^ 

00 

t^ 

t- 

00 

(N 

CO 

CO 

IC 

o 


73 
C 
cS 

73 

O 


3 
(1h 


K     g 


^ 


^- 


O 


1^ 

o 

eS 

J3 

■iJ 

-u 

t^ 

a 

ce 

P^ 

O 

S 

?5, 

."^ 

g 

3- 

'« 

£ 

o 

« 

tH 

S 

^ 

•§ 

!-) 

§ 

ci 

-§ 

+i 

^ 

a) 

CO 

o 

C/2 

to 
3 
o 
O 

C2s 
C3) 

-is! 
O 

INHIBITION  OF  SUCCINATE   OXIDATION 


55 


pq 


lO    CO    00    IC 
CO    O    ^    "O    l> 

O^HfOCSClCDOOO 

(M   00   -^   -<   — I 


-H      Tj<     O 


Ph 


o  ^ 


o 


cS 

^-v 

f^ 

o 

fC 

T3 

Ci 

S 

^ 

eS 

SI 

T3 

c« 

W 

fe 

■75     'IS     C5 


0)      « 


(Xi   a^   ^    lo  'Tit 
C5    :o    t-^    uo  C5 


CO      tJ^     -Tt^      -^ 

c^    c^    c^    t^ 


t-    c-    t> 


g 

t^ 

4J 

-p 

c 

O 

'3 

<0 

> 

c3 

1 

CO 

m 

i^ 

W 

K 

M 

1 

^       3 


O^    2 


^    H 


^     T3 


J     1 


_   -t^       — , 


O"  O"  Ph         O"  i:    W 


o 

o 

o 

i-O   o 

t^ 

t- 

>o 

t- 

o 

o 

(M 

(N 

<N 

-O 

o 

'M 

cr; 

o 

^H 

^    S    CC  S    M    S 


<»      c? 


W    Ph    Ph 


•-     .SP 


-J^ 

>. 

r^ 

,0 

o 

_^ 

>w' 

cS 

S 

<t)      OS 


56  1.    MALONATE 


Table  Ml 


Effects  of  Endogenous  Respikation  on  Calculations  of  Malonate  Inhibition 
OF  Succinate  Oxidation  " 


Pigeon 
muscle 

Rat  kidney 

Rat  liver 

Euglena 

O2  Uptake 

Endogenous 

11.8 

14.3 

14.7 

11.4 

Malonate 

3.0 

4.5 

12.4 

13.9 

Succinate 

19.0 

58.5 

20.2 

28.0 

Succinate  +  malonate 

6.0 

15.8 

14.1 

19.7 

%  Inhibition 

Endogenous  respiration 

74.6 

68.6 

15.6 

Stim21.9 

Total  respiration  in  the 

presence  of  succinate 

68.4 

73.0 

30.2 

29.6 

Succinate  oxidation 

58.3 

74.4 

69.1 

65.1 

°  The  true  inhibition  of  succinate  oxidation  is  given  in  the  bottom  row,  and  is  to 
be  compared  to  the  figures  in  the  row  immediately  above  where  the  correction  for  the 
endogenous  ejRFect  has  not  been  made.  The  concentrations  were:  pigeon  muscle  —  suc- 
cinate 1  vaM  and  malonate  10  iwM  (Stare  and  Baumann,  1939);  rat  kidney  —  suc- 
cinate 20  n\M  and  malonate  20  mM  (Fawaz  and  Fawaz,  1954);  rat  liver  —  succinate 
10  n\M  and  malonate  20  mM  (Edson  1936);  Euglena  —  succinate  10  mM  and  malonate 
10  ml/  (Danforth,  1953). 

HOOC — R — COOH  —  penetrated,  succinate  would  enter  cells  somewhat 
better  than  malonate,  but,  at  least  at  pH's  above  7,  it  seems  unlikely 
that  this  is  the  situation.  Also,  the  entrance  of  enough  of  the  undissociated 
caid  to  be  effective  would  presumably  decrease  the  intracellular  pH  signi- 
ficantly (see  Chaper  1-14).  Besides,  some  of  these  experiments  in  which 
malonate  was  inactive  were  done  at  low  pH's  (3.4  to  5.5).  A  question  that 
must  be  considered  is  whether  succinate  oxidation  is  always  entirely  intra- 
cellular. It  is  possible  that  succinate  oxidase  occurs  both  in  the  mitochondria 
and  in  the  plasma  membrane.  In  many  cases,  malonate  has  little  or  no  effect 
on  tissue  metabolism  or  function  at  concentrations  capable  of  inhibiting 
the  oxidation  of  added  succinate  completely.  This  is  well  seen  in  rat  ven- 
tricle slices  (Webb  et  al,  1949)  where  the  succinate  may  be  oxidized  at 
the  cell  surface,  and  it  is  interesting  that  in  this  tissue  the  potency  of  mal- 
onate in  vivo  and  in  vitro  is  the  same.  However,  in  most  instances,  the 
succinate  oxidase  seems  to  be  protected  in  some  manner  in  intact  cells. 
There  are  several  possible  factors  that  could  modify  the  malonate  inhi- 


INHIBITION  OF  SUCCINATE  OXIDATION  57 

bition  of  succinate  dehydrogenase  when  the  enzyme  is  isolated  from  the 
cells.  It  will  be  well  to  mention  some  of  these  in  order  to  emphasize  that 
there  are  usually  many  ways,  other  than  by  permeability,  by  which  unex- 
pected phenomena  may  be  explained. 

(a)  The  enzyme  environment  within  the  cell  is  different  from  the  artificial 
media  used  with  isolated  enzymes  (see  Chapter  1-9).  Many  substances  may 
be  able  to  alter  malonate  inhibition;  we  have  noted  the  effects  of  Ca+"'" 
and  ATP.  The  concentrations  of  such  substances  may  be  different  in  cell 
and  medium.  The  intracellular  pH  is  also  not  that  of  most  media  used  in 
enzyme  study  and  may  be  easily  changed  by  the  addition  of  external 
substrate  or  inhibitor. 

(b)  The  addition  of  succinate  to  cells  may  influence  the  endogenous  respiration; 
that  is,  the  change  in  Og  uptake  upon  adding  succinate  may  not  represent 
accurately  the  rate  of  succinate  oxidation.  In  other  words,  the  correction 
for  endogenous  respiration  may  be  in  error.  Also,  the  addition  of  succinate 
or  malonate  can  secondarily  alter  the  complex  balance  of  the  metabolic 
systems.  For  example,  oxalacetate  is  a  very  potent  inhibitor  of  succinate 
dehydrogenase  and  its  concentration  in  the  cell  may  be  a  controlling  factor 
in  the  operation  of  the  cycle.  Since  oxalacetate  can  be  formed  from  succin- 
ate, and  its  formation  inhibited  by  malonate,  secondary  changes  in  succinate 
dehydrogenase  activity  may  occur  that  are  easily  interpreted  as  due  to 
the  direct  effects  of  malonate. 

(c)  The  concentration  of  succinate  in  the  cell  may  already  be  appreciable 
and  the  addition  of  more  may  reduce  the  inhibition  by  malonate  because 
of  the  competitive  nature  of  the  inhibition. 

(d)  The  rate  of  succinate  oxidation  may  be  limited  by  the  rate  at  which  it 
can  enter  into  the  cells  or  tissus;  that  is,  the  succinate  oxidase  is  so  active 
that  the  succinate  is  oxidized  as  rapidly  as  it  enters.  In  such  a  case,  mal- 
onate would  be  quite  ineffective  until  the  enzyme  is  inhibited  sufficiently 
to  make  it  limiting. 

(e)  The  ratio  (malonate)  I  (succinate)  may  be  different  in  the  cell  than  in  the 
medium.  This  could  be  due  to  differences  in  the  permeabilities  of  the  cell 
membrane  towards  these  substances  for,  despite  the  fact  that  succinate  and 
malonate  have  similar  properties,  many  cases  of  differential  permeabilities 
to  more  closely  related  ions  are  known.  It  could  also  be  due  to  the  somewhat 
different  p/C^  values,  since  the  internal  concentrations  of  the  active  ions  are 
determined  by  the  ionization  constants  (see  Eq.  1-14-146  for  buffered 
cells). 

Whether  these  factors,  or  others,  are  responsible  for  the  anomalous  results 
mentioned  above  is  not  known.  It  would  be  very  useful  to  have  data  on  the 


58  1.    MALONATE 

rates  of  penetration  of  succinate  and  malonate  into  cells,  obtained  prefer- 
ably with  radioactive  material  and  under  the  same  conditions.  Quantitative 
studies  on  the  malonate  inhibition  of  intracellular  dye  reduction  resulting 
from  succinate  oxidation  might  also  be  informative  in  certain  instances. 
It  has  been  shown  that  malonate  inhibits  the  succinate-induced  reduction  of 
neotetrazolium  in  adipose  tissue  cells  (Fried  and  Antopol,  1957),  but  noth- 
ing otherwise  is  known  about  the  succinate  dehydrogenase  from  this  tissue. 

Competitive  Nature  of  the  Inhibition   in  Cellular  Systems 

It  has  been  shown  in  several  tissues  that  the  addition  of  succinate  will 
reverse  the  inhibition  of  respiration  produced  by  malonate.  Thus  in  Avena 
coleoptile  the  inhibition  by  50  mM  malonate  is  reduced  from  57.4%  to 
25.8%  upon  adding  succinate  (Bonner,  1948),  and  in  spinach  leaves  from 
75.4%  to  20.5%  (Bonner  and  Wildman,  1946).  In  chick  embryonic  carti- 
lage, the  depression  of  respiration  by  10  mM  malonate  is  reversed  by  100 
mM  succinate  but  not  by  10  mM  (Boyd  and  Neuman,  1954).  Such  results 
have  occasionally  been  stated  to  prove  the  competitive  nature  of  the  inhibi- 
tion but  this  reasoning  is  not  completely  valid.  The  mere  increase  in  O2  up- 
take seen  on  addition  of  succinate  to  malonate-inhibited  tissues  is  alone 
not  evidence  for  competition.  The  effects  of  succinate  on  uninhibited  tissue 
must  also  be  tested  and  it  must  be  shown  that  the  actual  inhibition  is  de- 
creased. A  decrease  in  the  inhibition  brought  about  by  increasing  succinate 
concentrations  has  indeed  been  reported  in  two  tissues,  pigeon  breast  muscle 
(Krebs  and  Johnson,  1948)  and  the  trypanosome  Crithidia  (Hunter,  1960) 
and  in  the  latter  a  true  competitive  inhibition  was  demonstrated  by  1/v — 1  /(S) 
plots.  The  data  are  given  in  Table  1-10.  It  is  probable  that  the  inhibition 
of  succinate  oxidation  in  cellular  systems  by  malonate  would  frequently 
not  obey  strictly  competitive  kinetics,  due  to  the  various  complexities  that 
arise,  as  discussed  in  the  previous  section,  even  though  the  primary  inhi- 
bition on  the  succinate  dehydrogenase  were  competitive.  Some  of  the  prob- 
lems involved  in  the  determination  of  the  type  of  inhibition  in  cells  have 
been  discussed  in  Chapter  1-9. 


INHIBITIONS    OF    ENZYMES 
OTHER  THAN  SUCCINATE  DEHYDROGENASE 

It  is  very  important  to  establish  the  degree  of  specificity  that  may  be 
achieved  in  the  use  of  malonate  under  various  conditions.  To  this  end 
we  shall  first  discuss  the  direct  evidence  for  the  inhibition  of  enzymes 
other  than  succinate  dehydrogenase,  and  then  proceed  to  the  effects  on  the 
operation  of  the  tricarboxylic  acid  cycle,  the  accumulation  of  succinate  and 
other  intermediates,  and  finally  the  antagonism  of  the  malonate  inhibition 


ENZYMES    OTHER   THAN   SUCCINATE    DEHYDROGENASE  59 

by  fiimarate.  We  shall  then  be  in  a  position  to  evaluate  the  specificity  of 
malonate.  There  are  other  reasons,  of  course,  for  taking  up  these  subjects; 
for  example,  malonate  is  frequently  used  to  block  the  cycle  in  living  tissue 
and  in  this  connection  it  is  essential  to  understand  how  malonate  can  alter 
cycle  activity  under  various  conditions. 

Some  effects  of  malonate  on  miscellaneous  enzymes  are  presented  in 
Table  1-12.  There  are  three  major  difficulties  in  the  establishment  of  the 
over  all  spectrum  of  action  of  malonate.  First,  there  are  many  quite  impor- 
tant enzymes  whose  response  to  malonate  has  never  been  investigated 
directly.  Second,  inspection  of  the  table  will  show  that  in  few  cases  has 
more  than  one  concentration  been  used,  and  often  the  single  concentration 
reported  is  either  too  high  or  too  low  to  be  of  much  value.  Third,  the  same 
enzyme  from  different  species  often  shows  widely  varying  susceptibility 
to  malonate  (e.g.  NADH  oxidase,  /5-glucuronidase,  lactate  dehydrogenase, 
oxalacetate  decarboxylase,  and  others  in  the  table),  making  it  clear  that 
there  are  many  different  spectra  of  malonate  inhibition  and  that  statements 
on  specificity  must  be  qualifield  by  naming  the  source  of  the  enzymes  in 
question. 

Ideally,  the  results  on  the  inhibition  of  an  enzyme  by  malonate  should 
be  given  for  several  concentrations,  preferably  convering  the  range  from  that 
concentrations  just  sufficient  to  produce  some  inhibitions,  through  that 
causing  approximately  50%  inhibition  to  higher  inhibitions  (unless  these 
latter  concentrations  are  unreasonably  high).  Another  way  of  looking  at  the 
problem  is  to  consider  that  range  of  malonate  concentrations  most  likely 
to  give  useful  information  when  tested  on  enzymes.  This  will  depend 
on  the  source  of  the  enzymes.  For  example,  in  mammalian  tissues  it  usu- 
ally requires  malonate  concentrations  between  2  and  5  mM  to  inhibit  suc- 
cinate dehydrogenase  around  90%  in  the  presence  of  5-10  mM  succinate. 
In  the  interests  of  establishing  the  degree  of  specificity,  it  would  thus  be 
most  important  to  test  the  effects  of  malonate  at  concentrations  around 
5  mM  on  enzymes  from  such  sources.  When  the  organism  studied  possesses  a 
succinate  dehydrogenase  less  sensitive  to  malonate,  correspondingly  higher 
concentrations  must  be  applied  to  the  other  enzymes. 

Instances  of  Competitive   Inhibition 

Until  it  is  time  to  discuss  the  matter  of  specificity,  there  is  little  to  say 
about  these  inhibitions  since  the  results  in  the  table  speak  for  themselves. 
It  is  evident  that  several  enzymes  other  than  succinate  dehydrogenase 
are  readily  inhibited.  One  would  not  be  surprised  if  enzymes  attacking  the 
dicarboxylate  anions,  where  the  carboxylate  groups  are  separated  by  two 
carbon  atoms,  were  inhibitable  by  malonate  to  some  extent,  since  it  is 
likely  that  these  enzymes  also  possess  cationic  groups  appropriately  spaced. 
Actually,  fumarase,  malate  dehydrogenase,  the  malic  enzyme,  oxalacetate 


60 


1.    MALONATE 


P5 


W 


^H  ^ 


^     .5     C 


m 


o 


^  iln        c3 


^       lO       ^ 


^^="•13 


C  ^  > 


c 

_ 

03 

S 

S 

O! 

S 

2 

>. 

"o 

§ 

O 

4j  :_, 


S     <D      5     ■:S     ^ 


~     .22 


fe     5     CO 


3       o 


-5      * 

t/3    W      N 


o 
o 

3 

g 

•S 

>> 

C 

15 

2 

> 

^ 
"o 

2 

-li 

-§ 

O 

k< 

^ 

IS 

Ld 
Q 

^ 

'JS 

so 

(0 
> 

3 

^ 
^ 

P5 

P^ 

05 

P5 

Q 


y     *- 


G       t«     .5 


2    'O 


'x 

05 

o 

cS 

<D 

T3 

r2 

"S 

o 

ft 

cS 

0 

'x 

ft 

o 

-.SO) 


OJ  03  03  OJ  C  r- 


X      S      G      c  ^ 


<j  <ii 


!-  O  uj 

O     O     Q 


o 


o  o     o   o     o 


G 


^ 

(B 

a 

03 

03 

cS 

O 

3 

0^ 

-0 

'x 

03 

ft 

12 

o 

^ 

» 

'x 

<© 

P       fM 


^     1 


fe5     ^   CM 


ENZYMES    OTHER   THAN   SUCCINATE  DEHYDROGENASE  61 


OS 

2 

1—4 
1— I 

„_^ 

>> 

^— ' 

00 

> 

b 

Tt^ 

> 

« 

c: 

o 

J3 

i-H 

J 

^.^ 

O 

'^-^ 

c^ 

^— V 

:S 

2 

05 

s 

0 

o 

T3 

s 

1     - 

is 

3 

a 

<H-I 

ft 

3 

C 
M 

T! 
>> 
jS 


5^  S 
CO  >o 

OS    OS 


2    X 


:r;     c    o 


■* 

lO 

OS 

^-v 

r- 

t^ 

tJ< 

fl 

iC 

OS 

o 

-»^ 

"-' 

— ^ 

e 

fe 

t3 

«^ 

"S 

s:; 

ilJ 

U-i 

c3 

cS 

& 

^ 

TS 

r^ 

V 

4^ 

O 

5 

^ 

o 

M 

S 

-•3       2-  ;3n-S  4)  =;o'- 


ic  ^  ^^  ec 

X    X 


~     ^  s    cs  .5 


cS    a>      o    «      o      3     7-^^     O  o     >      =1  .r?  .Sf     =c    c3  a>  ^2         v-  ^ 


® 

c3 

c 

cj 

« 

« 

"i 

00 

^ 

C 

T3 

ci 

>. 

o       ^      o    ^»  o  o 


3 

c3 

05 

>> 

c3 

p 

3 

_c3 

o 

5 

O 
>> 

X 

2 

CD 

-a 
>, 

X 

p 
>■, 

"S 

-^3 

cS 

u 
■*^ 

O 

CO 

1— 1 

<^        ^ 


g  +1       g      Ph        o 

3  3        rt        O        cS 

O  3         -li  tn        -2 

3  cS        3       «U        O 

-3  +j— .CJO  ^'Q.,^,_^_ocj  .-J 


62 


1.    MALONATE 


tf 


-^  -^    o 


J2      W) 
O    Z 


■r! 

-Q 

>jj 

C3 

t^ 

C5 

CO 

3 

,__ 

C5 

0^ 

2 

CO 



cS 

O 

03 

s 

3. 

^ 

C^ 

(S 

'S" 

a 

Q 

>^ 

o 

^         So 
O        cS      iS 


M  Q   C 


A    A 


o  o  o  o 

-M    Tt    00    Oi 


CD     O    O    CO     O     O     — H 
rt     -H     C<l    «N     CO     lO     ^ 


1^ 


K 


O        K-l 


i^    ^    S 


^ 


*     -,    - 
CO      S     i: 

?5     ^    e 


•»^  O        V. 


^   ^ 


• ^  "  (H 


c 

C 

« 

o 

o 

-=5 

(» 

Oj 

M 

tuD 

M 

£ 

e^ 

S 

S       cS 


ft 


ENZYMES    OTHER   THAN   SUCCINATE    DEHYDROGENASE 


63 


o    2  2 


73 


S  2   2  ^ 

-H    ^  "^   Ph 


ffi  -       -7- 


m 


P-        ^ 


m 


c  ^     5 


-a    rt 


^      ^   t; 


«:  ^  o 


^ 


?^         r-         r-         rH         >  > 


Si      o 


o 


2 

73 

1 

1^ 
>. 

4> 

CO 

a 

>-. 

">. 

:2 

ffl 

o 

X 

o 

o 

_cS 

13 

^ 

£ 

Q 

Is 

-c 

c« 

<i 

X 

§ 

^ 

§ 

"^ 

o 

Ph  CM  eu  2^     si 


C5        |l4 


i-H  03 


■^      .2 


«    c      §      2 


o  2 


9      c-     c 


;j    S  O 


lO       O       -H  ^H      ,—(      ^H 


p.    &    S     C     tj) 


-     -«^^a4^ 


S      «  —    =s 


!;^    p^ 


X     -S     -ii 


o    c3    3    3  CZ2  c:c  CO 

s  Pi  w  a  ^  ^  ^ 


«      3 


3         5 


O     fc 


-d 

>> 

^ 

0) 

S 

'S 

93 

S 

S 

OQ 

_>;. 

e3 

S 

■fci 

« 

o 

M 

O 

03 

73 

O 

>. 

-C 

-C 

Ph 

^ 

64 


1.    MALONATE 


cj     \2 


t— I         «  HH 


tf 


fe 


E^ 


5"    S 


S    P5 


O    PQ 


e  «5 


-^    .S 


(1-1     j!^        '-1 


=«    .2 


11^   eiH 


-c    ^       ju 


m 


^►, 

;j 

<D 

£2 

^ 

4^ 

73 

c3 

0 

IB 

)3 

J3 

1 

,4^ 

-ii! 

s 

0 

p       M) 


H 


-C3 


y,      g 


cs       rt 


o    o 


22    C     5   ^ 


-^ 

a 

a 

oi 

_c 

aj 

-t^ 

'c 

<33 

9J 

-u 

£ 

ft 

< 

M 

;^  p 


ENZYMES    OTHER   THAN   SUCCINATE    DEHYDROGENASE 


65 


decarboxylase,  and  the  condensing  enzyme  are  inhibited  by  malonate,  al- 
though usually  not  as  potently  as  is  succinate  dehydrogenase.  Enzymes 
catalyzing  reactions  of  the  dicarboxylates  in  which  the  charges  are  farther 
apart  (e.g.  a-ketoglutarate)  would  be  expected  to  be  less  susceptible.  It  is 
likely  that  these  inhibitions  are  mosth^  competitive  but  sufficient  data  to 
establish  this  are  generally  lacking. 

The  inhibition  oifumarase  by  malonate  has  been  shown  to  be  competitive 
in  the  thorough  study  of  Massey  (1953  b)  but  the  affinity  of  the  enzyme  for 
malonate  is  not  very  high  {K^  =  40  n\M).  This  implies  either  a  very  dif- 
ferent intercationic  distance  in  this  enzyme  from  that  in  succinate  dehydro- 
genase or  a  different  configuration  of  the  enzyme  surface  surrounding  these 
cationic  groups,  probably  the  latter.  Both  directions  of  the  reaction  cata- 
lyzed by  the  malic  enzyme  are  inhibited  by  malonate,  which  competes 
with  either  malate  (Stickland,  1959  b)  or  p\Tuvate  (Stickland,  1959  a)  (see 
accompanying  tabulation).  The  inhibition  of  this  enzyme  might  well  alter 


Malate 

%  Inhibition  by  malonate  at: 

Pyruvate 

%  Inhibition  by  malonate  at: 

(mif) 

2  mM 

5  mM 

20  mM 

{mM) 

1  mM 

10  mil/ 

0.1 

71 

88 

100 

1 

43 

62 

0.3 

48 

74 

91 

10 

37 

70 

1 

19 

48 

84 

50 

10 

46 

the  operation  of  the  tricarboxylic  acid  cycle  under  certain  conditions.  The 
situation  is  different  for  lactate  dehydrogenase,  since  malonate  is  competitive 
with  respect  to  lactate  but  noncompetitive  with  respect  to  pyruvate  (Ot- 
tolenghi  and  Denstedt,  1958),  leading  to  the  suggestion  that  these  two  sub- 
strates react  with  different  sites  on  the  enzyme.  The  X/s  are  6.4  m.M  for 
the  oxidation  of  lactate  and  27  mM  for  the  reduction  of  pyruvate.  Oxalate 
and  tartronate  are  much  better  inhibitors  of  this  enzyme.  The  D-a-hydroxy 
acid  dehydrogenase  of  yeast,  which  oxidizes  lactate,  malate,  a-hydroxybuty- 
rate,  and  glycerate,  is  competitively  inhibited  by  malonate  with  a  K^  of 
0.9  mil/  (Boeri  et  al.,  1960).  In  this  case,  oxalate  is  a  very  potent  inhibitor 
{Ki  =  0.0025  mM)  while  tartronate  is  of  similar  potency  to  malonate. 
Finally,  in  reducing  systems  in  which  succinate  can  serve  as  an  electron 
donor,  malonate  maj^  inhibit  competitively.  This  is  the  case  for  the  particu- 
late nitrate  reductase  of  soybean  root  nodules,  the  K^  for  malonate  being 
0.017  milf  (Cheniae  and  Evans,  1959).  The  nitrate  reductase  is  not  inhibited 
by  malonate  directly  but  the  results  on  the  over  aU  system  might  make  it 
appear  to  be  the  case.  In  all  cases  where  malonate  inhibits  competitively, 
the  susceptibility  of  the  reaction  in  complex  systems  will  depend  on  the 
concentration  of  the  substrate,  and  thus  may  be  quite  high  in  living  systems 
where  the  concentrations  of  intermediates  are  frequently  low. 


66  1.    MALONATE 

Inhibition  Due  to  Chelation  with  Metal  Ion  Cofactors 

Many  enzymes  are  dependent  on  dissociable  metal  ions  for  their  activity 
and  the  operation  of  most  of  the  important  metabolic  systems  thus  requires 
the  presence  of  these  cofactors.  The  list  of  enzymes  requiring  Mg++  is  a 
long  one  and  includes  the  oxidases  and  decarboxylases  for  the  keto  acids, 
most  of  the  enzymes  involved  in  phosphate  metabolism  (e.g.,  the  kinases, 
the  transphosphorylases,  the  phosphatases,  and  the  acyl — CoA  synthetases), 
some  dehydrogenases  e.g.,  phosphoglucose  dehydrogenase,  phosphogluco- 
nate  dehydrogenase,  and  isocitrate  dehydrogenase),  some  peptidases, 
phosphoglucomutase,  and  enolase.  Several  enzymes  require  Zn++,  such  as 
lactate  dehydrogenase,  glutamate  dehydrogenase,  alcohol  dehydrogenases, 
carboxy peptidase,  and  carbonic  anhydrase.  Since  malonate  is  able  to  chelate 
effectively  with  these  metal  ions,  inhibition  may  result  from  the  reduction 
of  metal  ion  concentration  in  the  medium  or  the  removal  of  the  metal 
ions  from  the  enzyme.  The  ability  of  malonate  to  inhibit  by  this  mechanism 
will  depend  on  the  affinity  of  the  enzyme  for  the  metal  ion.  The  binding 
of  Zn++  to  enzymes  is  usually  rather  strong  and  it  is  difficult  for  malonate 
to  deplete  the  enzyme  of  this  metal,  but  Mg++  is  more  loosely  bound  in  most 
cases  and  the  activity  of  enzymes  dependent  on  it  is  generally  related  to 
the  Mg+'*"  concentration  in  the  medium.  The  effect  of  malonate  on  Mg++- 
dependent  enzymes  will  thus  depend  on  the  concentrations  of  Mg++  and 
malonate,  and  on  the  relationship  between  enzyme  activity  and  Mg+*^ 
concentration.  In  the  use  of  malonate,  especially  at  higher  concentrations, 
it  is  imperative  to  consider  the  possibility  of  such  effects.  It  is  likely  that 
some  of  the  inhibitions  in  Table  1-12  are  due  to  metal  ion  depletion. 

Inhibition  by  reaction  with  an  activator  was  discussed  briefly  in  Chapter 
1-3  and  it  was  seen  that  in  the  general  case  no  simple  expression  for  the 
inhibition  is  possible.  Nevertheless,  it  should  be  reasonably  easy  to  deter- 
mine if  the  inhibition  is  purely  the  result  of  activator  depletion,  since  the 
concentration  of  free  activator  can  be  calculated  from  the  dissociation 
constant  of  the  activator-inhibitor  complex  by  an  equation  similar  to 
Eq.  1-3-72: 

(Mg)  =  ^  V  [(I,)  -  (Mg,)  +  K-\-  +  4(Mg)if  -  -^  [(I,)  -  (Mg,)  +  K]     (1-3) 

If  the  effect  on  the  enzyme  is  only  to  reduce  the  Mg++  concentration,  the 
addition  of  malonate  should  bring  the  activity  to  that  value  corresponding 
to  the  reduced  free  Mg++.  Another  possibility  is  that  the  Mg-malonate 
complex  is  the  active  inhibitor,  in  which  case  the  kinetics  should  be  investi- 
gated with  the  calculated  concentrations  of  this  complex  at  different  con- 
centrations of  Mg++  and  malonate. 

The  most  thorougly  studied  instance  of  the  possible  relationship  of  Mg++ 
to  inhibition  by  malonate  is  the  work  on  the  utilization  of  oxalacetate 


ENZYMES    OTHER   THAN  SUCCINATE   DEHYDROGENASE 


67 


in  rat  tissue  homogenates  by  Pardee  and  Potter  (1949).  The  formation 
of  citrate  here  probably  involves  decarboxylation  of  some  of  the  oxal- 
acetate  to  pyruvate,  with  subsequent  condensation  to  enter  the  cycle.  Al- 
though a  single  enzyme  was  not  studied,  it  is  possible  that  the  utilization 
of  oxalacetate  was  limited  by  the  decarboxylase.  This  reaction  is  activated 
by  low  concentrations  of  Mg++  and  inhibited  by  higher  concentrations 
(upper  curve  in  Fig.  1-10).  It  is  clear  that  the  inhibition  by  malonate  de- 


(    Mg*  •   ). 


30 
mM 


Fig.  1-10.  Utilization  of  oxalacetate   by  a  rat 

kidney  homogenate  at  diflferent   concentrations 

of   Mg++,   with   and   without   malonate.    (From 

Pardee  and  Potter,  1949). 


creases  as  the  Mg"*"*"  concentration  is  raised  (lower  curve).  This  was  interpret- 
ed to  mean  that  the  inhibition  is  mainly  due  to  depletion  of  Mg++.  Some 
arguments  may  be  brought  against  this  interpretation.  Table  1-13  gives 
my  calculations  of  the  concentrations  of  free  Mg^^,  free  malonate,  and  the 
complex  at  the  different  levels  of  total  Mg++  used  in  the  experiment  shown 
in  Fig.  1-10.  It  is  seen  that  the  malonate  at  10  mM  does  indeed  reduce  the 
free  Mg++,  but  in  the  higher  range  of  Mg++  concentrations  this  should  in- 
crease the  activity  rather  than  decrease  it,  because  in  this  range  Mg"'"+  is 
somewhat  inhibitory.  Pardee  and  Potter  did  not  consider  the  reduction  in 
free  malonate  concentration,  which  is  very  marked  as  shown  in  Table  1-13. 
This  could  well  acount  for  the  decrease  in  the  inhibition  from  47%  at  zero 
Mg++  concentration  to  12%  at  30  roM  Mg++.  One  might  speculate  that  the 
reduction  in  rate  at  high  Mg++  concentrations  could  be  due  to  the  complexing 
of  oxalacetate  so  that  it  is  unable  to  react  with  the  enzyme.  The  question 
of  the  mechanism  of  the  malonate  inhibition  is  thus  not  settled.  It  was 
claimed  that  the  inhibition  is  not  typically  competitive  because  increase 
in  the  oxalacetate  concentration  actually  increases  the  inhibition  somewhat, 


68  1.    MALONATE 


Table  1-13 


Effects  of  Total  Mg++  Concentration  on  the  Concentrations  of  Free  Mg++, 
Free  Malonate,  and  Ms-Malonate  Complex  " 


(Mg++,) 

(Mg++) 

(Mg 

-malonate) 

(Malonate) 

(milf) 

(mif) 

(mif) 

(mM) 

0 

0 

0 

10 

3 

1.59 

1.41 

8.59 

10 

6.13 

3.87 

6.13 

20 

14.09 

5.91 

4.09 

30 

22.98 

7.02 

2.98 

"  Conditions  as  in  Fig.  1-10  from  the  work  of  Pardee  and  Potter  (1949).  The  total 
malonate  concentration  is  10  mM  in  all  cases.  The  values  were  calculated  from  Eq. 
1-3  using  9.77  X  10^^  M  for  the  dissociation  constant  of  the  Mg-malonate  complex. 

although  no  figures  were  given  so  the  magnitude  of  the  effect  is  unknown. 
An  increase  of  oxalacetate  might  reduce  the  amount  of  Mg++  bound  to 
malonate  and  thereby  increase  the  free  malonate  concentration.  It  would 
appear  unlikely  that  the  Mg-malonate  complex  is  inhibitory  since  its  con- 
centration increases  with  Mg++  concentration  (Table  1-13),  whereas  the 
inhibition  decreases.  However,  it  is  evident  that  malonate  inhibits  this 
enzyme  system  in  some  manner  directly,  not  only  from  the  above  consid- 
erations but  also  because  of  the  marked  inhibition  observed  in  the  absence 
of  Mg++. 

Dialkylfluorophosphatase  is  activated  by  Mn++  and  its  inhibition  by 
malonate  was  attributed  by  Mounter  and  Chanutin  (1953)  to  the  chelation 
of  the  activator.  Since  no  concentrations  of  either  Mn++  or  malonate  were 
given,  it  is  impossible  to  evaluate  the  results.  However,  their  data  show 
that  malonate  inhibits  just  as  well,  if  not  better,  in  the  absence  of  Mn++ 
(21%  with  Mn++  added  and  28%  without  Mn++  at  15  min).  These  results 
might  better  be  interpreted  as  due  to  removal  of  free  malonate  by  the 
Mn++.  If  Mn-malonate  is  incubated  with  the  Mn-free  enzyme,  activity 
slowly  appears,  indicating  that  the  Mn++  is  transferred  from  the  malonate 
to  the  enzyme,  which  is  not  surprising  since  the  affinity  of  the  enzyme  for 
the  Mn++  is  much  greater,  as  shown  by  the  dissociation  constants  {pK  for 
enzyme-Mn++  complex  is  7.7).  It  would  be  very  surprising  under  these 
circumstances  if  malonate  were  able  to  reduce  the  Mn++  sufficiently  to 
inhibit  the  enzyme. 

There  are  several  other  claims  for  this  mechanism  of  malonate  inhibi- 
tion but  in  all  cases  there  is  either  inadequate  evidence  or  no  evidence  at 


EFFECTS  ON  TRICAKBOXYLIC  ACID  CYCLE  69 

all.  The  examples  discussed  above  indicate  the  impossibility  of  establishing 
such  a  mechanism  without  considering  the  changes  in  malonate  concentra- 
tion or  treating  the  data  quantitatively.  Despite  this  lack  of  positive  evi- 
dence, there  is  certainly  no  doubt  but  that  this  type  of  inhibition  can  occur 
and  may  be  sometimes  very  important.  In  the  oxidation  of  pyruvate  by 
rat  heart  mitochondria,  the  Mg++  concentration  must  fall  below  1  mM 
before  there  is  any  significant  decrease  in  the  rate  (Montgomery  and  Webb, 
1956  b).  We  usually  used  5  mM  Mg++  in  the  medium  so  that  it  would  have 
required  at  least  40  mM  malonate  to  produce  a  detectable  inhibition  by  this 
mechanism.  Malonate  at  50  mM  did  indeed  inhibit  around  50%  but  this 
must  certainly  be  due  to  other  actions  to  a  large  extent.  These  experiments 
were  done  with  the  a-ketoglutarate  oxidase  blocked  by  parapyruvate  so 
that  any  inhibition  of  succinate  oxidation  would  not  be  involved. 

EFFECTS    OF    MALONATE    ON    THE    OPERATION    OF   THE 
TRICARBOXYLIC    ACID    CYCLE 

Malonate  is  usually  assumed  to  produce  its  major  effects  on  cellular  meta- 
bolism and  function  by  disturbing  the  operation  of  the  cycle*  and  reducing 
the  rate  of  formation  of  ATP.  Malonate  has  often  been  used  to  establish  if 
the  cycle  is  operative  in  a  tissue  or  if  a  particular  functional  activity  is 
dependent  on  the  cycle.  It  is  thus  important  to  examine  critically  the 
nature  of  the  cycle  block  and  the  effects  it  may  have  on  the  over-all  oxida- 
tive metabolism.  There  are  two  aspects  that  are  especially  relevant  to  this 
question.  There  is  the  problem  of  the  specificity  of  action  of  malonate  on 
succinate  dehydrogenase  and  this  will  be  considered  later.  In  the  present 
section  we  shall  assume  that  the  only  inhibition  is  on  the  oxidation  of  suc- 
cinate and  discuss  the  problems  relative  to  the  interpretation  of  such 
a  block.  Before  treating  the  actual  results  obtained  with  malonate,  the 
nature  of  the  cycle  and  its  responses  to  inhibition  will  be  outlined. 

Some  General  Principles  of  Cycle  Block 

The  primary  function  of  the  cycle  is  to  incorporate  and  oxidize  acetyl-CoA, 
whether  this  arises  from  pyruvate,  acetate,  fatty  acids,  or  elsewhere,  and 
thus  it  is  particularly  important  to  discuss  the  effects  of  malonate  block  on 
this.  The  situation  is  relatively  clear  in  suspensions  of  isolated  mitochondria, 
in  which  the  concentrations  of  the  cycle  intermediates  are  low  and  the  endo- 
genous respiration  is  generally  negligible.  Pyruvate,  or  other  substances 
giving  rise  to  acetyl-CoA,  may  be  oxidized  through  the  cycle  only  if  some 

*  In  this  chapter  the  term  "cycle"  will  always  refer  to  the  tricarboxylic  acid  cycle 
for  convenience  and  other  cycles  will  be  designated  by  their  special  names. 


70  1.    MALONATE 

cycle  intermediate  (sparker)  is  provided  to  furnish  oxalacetate  to  condense 
with  the  acetyl-CoA.  A  very  small  amount  of  such  a  sparker  may  suffice  to 
initiate  the  entry  of  the  acetyl-CoA  into  the  cycle,  and  the  cycle  will  then 
perpetuate  itself  through  the  continuous  formation  of  oxalacetate,  in  which 
case  pyruvate  will  be  completely  oxidized  to  COg  and  water.  Such  a  system 
should  be  quite  sensitive  to  malonate,  because  an  inhibition  of  the  oxidation 
of  succinate  will  reduce  the  amount  of  oxalacetate  formed  and  consequently 
the  amount  of  acetyl-CoA  entering  the  cycle.  On  the  other  hand,  if  an 
approximately  molar  equivalent  of  fumarate,  malate,  or  oxalacetate  is 
initially  present  with  the  pyruvate,  there  wiU  be  an  adequate  concentration 
of  oxalacetate  to  incorporate  acetyl-CoA  at  a  rapid  rate,  and  the  process  will 
not  depend  on  a  regeneration  of  oxalacetate.  This  system  will  not  be  very 
sensitive  to  malonate,  because  a  block  of  succinate  oxidase  will  not  apprec- 
iably reduce  the  amount  of  oxalacetate  present.  The  first  important  princi- 
ple is,  therefore,  that  the  degree  of  cycle  inhibition  by  malonate  will  depend 
on  the  source  of  oxalacetate. 

When  the  cycle  is  operating  in  a  steady  state,  the  concentrations  of 
intermediates  are  low,  and  oxalacetate  is  formed  just  as  rapidly  as  pyruvate 
is  incorporated,  the  cycle  rate  being  limited  by  the  entry  of  acetyl-CoA, 
This  may  be  the  normal  state  of  the  cycle  (Krebs  and  Lowenstain,  1960)  but 
probably  in  cells,  and  certainly  in  isolated  preparations,  there  are  times  when 
the  cycle  is  not  in  a  steady  state.  There  is  an  initial  rise  in  citrate  concentra- 
tion during  the  oxidation  of  pyruvate  by  heart  mitochondria  in  the  presence 
of  malate  (Montgomery  and  Webb,  1956  a),  indicating  that  the  tricarboxy- 
lates  cannot  be  handled  as  rapidly  as  pyruvate  can  enter  the  cycle  when  the 
supply  of  oxalacetate  is  not  limiting.  The  rate  of  oxygen  uptake  is  initially 
very  high,  falls  to  a  new  level  during  the  first  40  min,  maintains  this  level 
for  2-3  hr,  and  then  suddenly  fails  when  the  pyruvate  is  completely  utilized. 
The  first  phase  occurs  when  oxalacetate  is  readily  available,  and  corresponds 
to  the  accumulation  of  citrate;  the  rate  of  oxygen  uptake  during  this  period 
is  not  an  accurate  measure  of  the  rate  of  operation  of  the  entire  cycle  — 
block  of  succinate  oxidation  may  have  very  little  effect  on  the  oxygen  up- 
take because  relatively  little  of  the  respiration  arises  from  this  region  of  the 
cycle.  The  second  steady-state  phase  should  be  more  sensitive  to  malonate 
because  the  oxidations  of  succinate  and  malate  now  contribute  2  of  the 
total  of  5  oxygen  atoms  taken  up  per  molecule  of  pyruvate.  The  second 
principle  is  thus  that  malonate  inhibition  will  sometimes  depend  on  the  time 
interval  during  which  the  oxygen  uptake  is  measured,  particularly  whether 
it  is  the  initial  rate  or  the  total  oxygen  consumed. 

Succinate  usually  accumulates  during  malonate  inhibition  (see  page  90) 
and  this  will  progressively  reduce  the  degree  of  inhibition  due  to  the  com- 
petitive nature  of  the  inhibition.  Eventually  a  new  steady  state  may  be 
reached  during  which  the  succinate  concentration  remains  constant.  This 


EFFECTS  ON  TRICARBOXYLIC  ACID  CYCLE  71 

will  always  tend  to  lessen  the  effect  of  malonate  and  under  certain  circum- 
stances it  might  effectively  overcome  the  inhibition.  The  third  principle 
is  that  the  degree  of  malonate  inhibition  will  depend  on  the  level  of  suc- 
cinate accumulation  in  the  system  studied. 

Oxalacetate  can  often  be  formed  by  reactions  outside  the  cycle.  Pyruvate 
and  phosphoenolpyruvate  can  be  carboxylated  to  oxalacetate  in  the  presence 
of  oxalacetate  decarboxylases  or  oxalacetokinase  (Bandurski,  1955),  and 
transamination  between  a-ketoglutarate  and  aspartate  may  also  give  rise 
to  oxalacetate.  In  such  cases  the  inhibition  of  pyruvate  oxidation  in  the 
cycle  by  malonate  will  be  reduced  because  the  incorporation  of  pjTuvate 
will  not  be  dependent  only  on  the  regeneration  of  oxalacetate  (Holland  and 
Humphrey,  1953).  Such  reactions  may  occur  in  isolated  mitochondria,  as 
well  as  in  cells,  since  in  heart  mitochondria,  where  pyruvate  alone  is  not 
oxidized  at  all,  the  presence  of  bicarbonate  or  COg  allows  a  substantial  rate 
of  pyruvate  oxidation  (Montgomery  and  Webb,  1956  a),  presumably 
through  the  carboxylation  of  some  of  the  pyruvate  to  oxalacetate.  The 
fourth  principle  is  that  the  degree  of  malonate  inhibition  will  depend  on 
noncycle  sources  of  oxalacetate. 

Alternate  metabolic  pathways  involving  cycle  substrates  or  intermediates 
may  occur  in  some  tissues.  There  are  many  opportunities  for  the  metabolism 
of  pyruvate,  in  addition  to  its  oxidation  through  the  cycle,  and  in  the  pres- 
ence of  malonate  these  pathways  may  become  important.  This  is  particu- 
larly true  in  microorganisms  but  the  ability  to  decarboxylate  pyruvate  to 
acetate  is  common  to  most  species  and  tissues.  Thus,  in  the  presence  of 
high  concentrations  (50  mJf )  of  malonate,  pyruvate  is  quantitatively  trans- 
formed into  acetate  by  rabbit  heart  mitochondria  (Fuld  and  Paul,  1952). 
An  alternate  pathway  for  succinate  that  would  circumvent  a  malonate 
block  is  the  cleavage  of  succinate  (in  the  presence  of  NADH,  CoA,  and 
ATP)  to  2  acetyl-CoA  molecules.  This  succinate-cleaving  enzyme  was  dis- 
covered in  Tetrahymena  (Seaman  and  Naschke,  1955)  but  it  is  also  active 
in  several  rat  tissues  and  in  certain  bacteria.  This  reaction  will  not,  of  course, 
restore  cycle  activity  but  it  can  lead  to  the  formation  of  acetate  or  other 
products  from  acetyl-CoA,  as  well  as  reduce  the  concentration  of  succinate. 
Finally,  the  recently  delineated  glyoxylate  cycle  (Kornberg  and  Krebs, 
1957)  could  bypass  that  region  of  the  cycle  containing  succinate  oxidase, 
malate  being  formed  from  isocitrate  through  the  condensation  of  glyoxylate 
and  acetyl-CoA,  the  over-all  process  being  the  formation  of  succinate  from 

2  PjTuvate  +  3/2  O2    ->    succinate  +  2  CO2  +  HgO 

pyruvate.  This  shunt  would  allow  a  greater  utilization  of  pyruvate  and  a 
greater  oxygen  uptake  in  the  presence  of  malonate  than  would  be  the  case 
with  the  tricarboxylic  acid  cycle  alone.  The  glyoxylate  cycle  has  been 
found  in  many  microorganisms  and  there  is  some  evidence  for  its  occurrence 


72  1.    MALONATE 

in  certain  plants,  but  its  role  in  animal  tissues  is  as  yet  unknown  (Krebs  and 
Lowenstein,  1960).  As  a  result  of  these  considerations,  the  fifth  principle 
of  cycle  block  is  that  the  degree  of  inhibition  by  malonate  will  depend  on 
the  activity  of  various  alternate  pathways  and  shunts;  in  addition,  it  will 
depend  on  what  is  measured,  e.g.,  oxygen  uptake,  CO2  production,  or 
pyruvate  disappearance. 

There  is  no  doubt,  therefore,  tha  the  operation  of  the  cycle  and  any 
ancillary  pathways  will  vary  with  the  experimental  or  physiological  condi- 
tions, and  that  one  must  expect  marked  differences  in  the  behavior  of  the 
cycle  in  different  species  or  tissues.  In  addition  to  the  factors  discussed  a- 
bove,  there  are  several  other  reasons  for  variability  in  response  to  malonate; 
the  different  susceptibilities  of  succinate  dehydrogenase  to  inhibition  (see 
page  49),  the  failure  of  malonate  to  penetrate  readily  into  cells,  and  the  pos- 
sibility that  malonate  can  inhibit  other  enzymes.  The  reliability  of  malonate 
as  an  indicator  of  cycle  activity  in  a  tissue  must  be  evaluated  in  the  light 
of  these  considerations.  Certainly  the  lack  of  an  expected  response  to  mal- 
onate cannot  be  immediately  interpreted  as  indicating  the  absence  of  the 
cycle,  and  the  production  of  a  significant  effect  by  malonate  should  be 
substantiated  by  other  more  direct  evidence  before  the  operation  of  the 
cycle  is  established. 

Inhibition  of  Cycle  Substrate  Oxidation  by  Malonate 

A  summary  of  some  of  the  effects  of  malonate  on  cycle  oxidations  is 
given  in  Table  1-14.  In  many  cases  the  concentration  of  malonate  is  too 
high  to  act  specifically  on  succinate  dehydrogenase,  and  the  results  are  to 
some  extent  meaningless.  It  is  very  difficult  to  make  any  generalizations 
but  malonate  concentrations  above  10  mM  in  subcellular  preparations  must 
be  looked  upon  as  probably  not  completely  specific,  whereas  in  cellular 
systems  it  is  impossible  to  evaluate  the  specificity  because  the  intracellular 
concentration  is  not  known.  In  attempting  to  interpret  the  inhibitions 
observed,  it  is  often  necessary  to  know  the  pathway  of  metabolism  of  the 
substrate  and  how  much  oxygen  is  normally  taken  up  per  molecule  utilized. 
For  example,  when  a-ketoglutarate  is  added  to  a  mitochondrial  suspension, 
it  may  be  oxidized  to  fumarate  (or  malate)  taking  up  2  atoms  of  oxygen, 
or  to  oxalacetate  taking  up  3  atoms  of  oxygen,  or  completely  taking  up  8 
atoms  of  oxygen.  If  succinate  oxidation  is  completely  blocked  by  malonate, 
only  1  atom  of  oxygen  will  be  taken  up.  Thus  the  maximal  inhibitions  in 
these  three  cases  are  50%,  67%,  and  87.5%,  respectively.  Similar  reasoning 
applies  to  each  substrate  and  in  many  cases  the  exact  fate  of  the  substrate 
is  not  known  so  that  it  is  difficult  to  estimate  what  effect  might  be  expected 
from  malonate.  It  is  also  important  in  this  connection,  as  pointed  out  in 
the  previous  section,  to  distinguish  between  inhibition  over  an  initial  short 
period  of  oxidation  and  inhibition  of  the  total  oxygen  uptake. 


EFFECTS  ON  TRICARBOXYLIC  ACID  CYCLE  73 

(a)  Inhibition  of  pyruvate  oxidation.  The  oxidation  or  disappearance  of 
pyruvate  in  cellular  preparations  is  usually  not  depressed  very  much  by  mal- 
onate  at  concentrations  less  than  10  miH ,  whereas  in  mitochondrial  prepara- 
tions the  expected  degree  of  inhibition  is  usually  observed.  This  may  be 
partly  explained  by  poor  penetration  into  the  cells  and  partly  by  the  alter- 
nate pathways  that  may  reduce  the  importance  of  the  cycle.  One  type  of 
correction  that  can  be  applied  for  a  more  accurate  determination  of  cycle 
inhibition  by  malonate  is  that  used  by  Speck  et  al.,  (1946).  In  malarial 
parasitized  erythrocytes,  pyruvate  is  oxidized  without  the  appearance 
of  acetate,  but  in  the  presence  of  malonate,  some  acetate  in  formed.  Cor- 
rection was  made  for  that  pyruvate  that  went  to  acetate,  since  this  fraction 
of  the  pyruvate  utilization  is  not  dependent  on  the  cycle.  The  inhibition  of 
over  all  pyruvate  utilization  was  12%  but  corrected  for  the  acetate  it  was 
31%.  In  the  free  parasites,  the  over  all  inhibition  was  33%  and  the  cor- 
rected inhibition  76%.  Of  course,  pyruvate  here  or  in  other  cells  may  be 
metabolized  in  other  ways,  so  that  the  correction  for  acetate  alone  may 
not  give  the  true  cj^cle  inhibition,  but  at  least  is  provides  a  better  value. 

Malonate  should  inhibit  the  oxidation  of  pyruvate  more  strongly  when 
there  is  a  low  concentration  initially  of  oxalacetate  or  a  substance  forming 
oxalacetate  (see  page  70).  This  was  shown  in  homogenates  of  rat  tissues 
by  Pardee  and  Potter  (1949).  In  each  case  the  inhibition  by  4  mM  malonate 


Tissue 

Substrates 

%  Inhibition 

Heart 

Pyruvate 

91 

Pyruvate  +  oxalacetate 

56 

Kidney 

Pyruvate 

93 

Pyruvate  -|-  oxalacetate 

55 

Brain 

PjTuvate 

74 

Pyruvate  +  oxalacetate 

47 

Liver 

Pyruvate 

28 

Pyruvate  +  oxalacetate 

15 

is  less  when  oxalacetate,  is  present.  Since  the  oxygen  uptake  was  deter- 
mined from  10  to  30  min  after  the  start  of  the  experiments  and  inasmuch  as 
the  concentrations  of  pyruvate  and  oxalacetate  were  3.5  mM,  it  is  unexpect- 
ed that  so  much  inhibition  would  be  exerted  when  the  mixture  is  present. 
This  might  be  due  to  the  decarboxylation  of  sufficient  oxalacetate  so  that 
it  was  less  effective  than  anticipated,  or  even  at  this  low  concentration 
malonate  may  have  been  inhibiting  some  reaction  other  than  the  oxidation 
of  succinate.  In  rat  heart  mitochondria  we  found  that  5  mM  malonate 
inhibits  the  oxidation  of  5  mM  pyruvate  only  about  10%  in  the  presence  of 


74 


1.    MALONATE 


P5 


»0     r-H 

03     ^- 


M  OS 

^     -«     -H 


M 


ft  ^ 

ft  2 

O  ^-' 

5  'O 


o3     _ 


c 

eS 

T--J 

e 

2 

.^ 

'« 

aj 

CC 

02 

-C 

pQ 

-c  i-i  3  M 


=D    & 


Oi         ^  -e 


ft 

ft  -« 

O  C 

CZ2  1-1 


^    ft 

c    S 


<  Q 


a 


3  TP 

C5    - — • 


-5  ^   S  -3 

M    W    CO    CO 


CO 
H 

Q 

o 


M  O    -H    (M    lO    M    O 


O    O 
O    ^ 


oooooooo 


o  o  o  o 


OOOOmOOOOOOOoj 


o 

Q 
O 

K 


Q 

•§ 

5^ 

-§ 

.u 

o 

-< 

s 

«:> 

<u 

=0 

«o 

f^ 

Q^ 

o     5j   t3  ;::; 


O    — '  o 

2   -S   -S    cs 


■^  .!2  "^    <»    "^ 
i=^  ^  :f^  ^'^  .-"^    c 

Oh    ^    ft^    >H    kH    H 


c 
o 

o 
o 

C     S     ft 


o 

b4 

T3 

6C 

-§ 

'E 

1 

ji 

C 

S 

S 

o 
o 

O 

43 

s 

o 

o 

S 

to 

S 

^ 

s 
•^ 

o 

o 

.g 

'« 

73 

'ft 

« 

o 

o 
O 

2 

s 

J, 

Q 

> 

a 

3 

J2 

"o 

<I1 

o 

02 

a, 

o 

s^ 

03 

> 


EFFECTS  ON  TRICARBOXYLIC  ACID  CYCLE  75 


CD 

O 

CO 

05 

CD 

o 

^^ 

2 

05 

^ 

C<) 

Oi 

^ 

ft 
ft 

C 

i 

c 

on 
c3 

ST 

O 
CO 

1—1 

§ 

o 

CO 

CO 

(H 

o 

p-l 

T3 

-D 

C 

CO 

s 

T3 
C 

o 

73 

-D 
ID 

e 

CO 

C5 

2 

c 

"e 

73 

^^ 

x 

c 

a; 

C 

ID 

-c 

c8 
S 

s 

-2 

M 
C 
c3 

s 

c 
o 

CO 

is 

Is 

s 

c 
o 

O 

c 

M 
15 

B 
o 
«> 
c 
o 

Si 

s 
2 

c 
.2 

o 

M 

w 

m 

w 

§ 

fi 

^ 

PM 

s 

Q 

PM 

§ 

o 

H 

^  D 


CO 

o 

05 

^H 

-— ^ 

1 

^ 

^ 

73 

c 
ce 

a 

Ch 

>> 

pq 

<» 

a 

o 

-0 

bj) 

cS 

■1^ 

O 

.1 

§  p^ 


lO-HF-HCOCOCOOO-^OO-^iCOlCOiOCO^H-^COCDC^l^OCOCOt^^H 
05(MI>'<1<  (M  Ttl(NTt<Q0»OO5O5  Tti00t^Tj<Tt<t>C0COCO  f-^ 


lCrti0-*-<*<O'*C0OO-*OlCOTt<Ok0OO©OOlC(Nl0O 


00202        OOccO        O        O        OOO        OOomOmOcc  OO 


oc  o.S^         ■"         a  c 


^  C  S  73 


-c 

^    c 
o 


(B      C      o; 


a.sSo^fi_os  M 

3  3  o  o  c  -^     o 


.a  .a  »  "^ 

cS      eS    -C      M 
-D    42      M     ^4 


<»        ;5        ;^ 


bl 

C 

C 

C 

0^ 

73 

73 

-^ 

^ 

^ 

13 

-*J 

-t^ 

-1-3 

-i-3 

eg 

c6 

c3 

c3 

P5 

P5 

P5 

P5 

o    o  ^    S 

«  a>  o  a 

OPh  SSqooc4p5p5p5p5  Ci  P^O^ 


76 


1.    MALONATE 


M 


«:> 

CO 

lO 

»o 

00 

05 

05 

C5 

05 

^-' 

•-' 

^ 

o 

^ 

-o 
^ 

tH 

4> 

05 

<v 

s 

o 

-0 

c3 

n3 

a 

"S 

TJ 

W 

>> 

T3 
C 

S 

0) 

s 

4) 

cS 

o 

o 

a; 

CO 

o 

^ 

Ph 

P5    S 

§ 

3    S^l 


G? 


-o 

s 

CO 

05 

05 

^ 

CO 

Oi 

^ 

a 

OS 

■^ 

^ 

'o 

•■"I 

CJ 

^ 

" — 

<D 

"n 

0^ 

CD 

73 

c 
c 

05 

o 

m 

a 

-t> 

c 

> 

2 

O 

cS 

CO 

o 

?? 

Ph 

>, 

Ti 

X 

CO 

c 

^ 
« 

c 

T3 
c3 

+3 

4) 

O 

3 

C 

o 

a 

> 

O 

a 

cS 

o 

o 

O 

3 

Ph 

S 

H^l 

O 

Q 

H 


O    -H    fo 
t^    O    O 


fcao'-'^-^'MCiao30fOt^Tt<-^i>0'M03DC505 

fO  1— i'>)COiOCOTt(OOCOt--.t^'MO'— iCCfOCOCO 


0-*0    0(MiOOOC'OOOOOOtJ<0»0    0 

^1         c-i   -H  ^H   '^^   CO   ■*   LO   F^   Tt   Tt<   Tf<         im 


t(<    O    O    O    O 

■ri  o   o  o 

r-l      (M      -^ 


O  O        «}  o 


QOOojO        OOO 


O 


O    ^     o 


o3      c3 


Pi 


>::^  o 


P5        P5        P5 


EFFECTS  ON  TRICARBOXYLIC  ACID  CYCLE  77 


n  -Q 


u5  ii 


jj      cj      3         ^      ci    ^  JS 

^  ffi  PQ     c«  H  o  a 


O    O    O    C-1 


:3^ 

-O 

C5 

^^ 

1 

00 

f-H 

'M 

rt 

lo 

3 

0) 

O 

CO 

ii 

o 

-^ 
5 

o 

CO 

05 

05      >■/} 

c2 

05 

C 
4J 

> 

"o 

S 

0)      (U 

c 

s 

o 

o 

-*-• 

>-. 

o 

b    u 

Cw 

o 

K 

« 

J 

h:j 

w 

M  fc^  Q 

o 

OOOOOa;  O  OOOOOcc        OCOO        Oo 


2    s  ^ 

2  S  I  1  i  -  ^ 


rt     2     cj         o     00   "-^  «*• 


^1        S-2 


2     -5     t* 

S  "S    '^ 

S        ""^ 
ace 

ai     c3     cS 
—     £     S 


S^  =53S"uo  =S  ;:§S  See  rtce!:>:S®£ 


78 


1.    MALONATE 


p:5 


If 

IC   lO   o   ic   o   o 

rf*            — <    'M    -H    (M 

le  S 

o    "^  ^_ 


M  M  <5 


2  2  2 


05     *^ 

o 

CO 

<4J 

"a 

05 

5 
> 

73 

CO 
CO 

eS     * 

s._^ 

o 

.^^ 

C 

ce 

13    ^ 

to     eS 

o  ^ 

o 
o 
o 

o 
o 

c6 

CO      tc 

14  « 

»r5 

t-    Xh 


^    3 


m 

s 

T3 

T3 

C 

C 

eS 

cS 

k( 

fi 

<D 

o 

c 

<  M 


O  O 


"2 

3 

4i 

■1^ 

d 

cS 

tf 

tf 

lO    o 


O    -^    O    C^ 


O   O   O   o   o 
M    O    O    ^H    i^j 


O  O      O  M 


o 

O 

.« 

^ 

-u 

s 

CO     e3 

^ 

-kJ 

,^ 

11 

1-^       C3 

1 

^ 

e 

■^    ft 

-ko    c 

.Vi 

a, 

40 

3  s 

03     3 

Pi  W 

6q 

ft? 

»0     3    00    00    o 
.S    00    ■<#    <N 


o  o  ^ 
o  o 


o  o  o  o  o 


OO     MOOOO        OOOOO 


(^      c      s      .^      ^ 


cS      o 


lU      «      0)    ^ 


»>   ai   s 


S       P  -S 


S 


S     S     o    qH 


"C  -tj  -u  -kj 

5  c^  c^  c3 

^  Im  t-<  Sh 

-"  -ki  +i  4J     .-ti 

^  o  o  o   - 

is 
o 


J>        Q     ^      ^      ^ 


5U    .M    ;— 


5  -H  ic  o 


Q,  CQ  o 


-0 

a 

'a 

o 

=c 

o 

S 
^ 

-u 

s 

■§ 

e 

Of) 

O 

5j 

^ 

o 

o 

Q 

> 

-< 

<1 

Oh 

o 


EFFECTS  ON  TRICARBOXYLIC  ACID  CYCLE  79 


05  _ 


CD 

i  ^  S"  s^         ?;r         o^  f? 


ececg^ji;  "^  ^^C  55 


jn  ^  05       CO  CO    S 

ST  r-        ^  C  ^-^  i-H      ^H        f** 


oscocog,^;,^  rr  3    —  S— ' 


tj&ciSK©         CD'S        o        SS^mS,  -g         rt  j=iJ  .^«3 


CO    IC    W    CD    L'5    O    C<1    CO 


s  s 


CQ 


O         OOOOOO  COCOCDOOOO 


OOOojOO        mO      O      OOOOOO        OOOOO002OOOO 


.2 

Lh 

c 

c 

73 

-a 

a 

a 

c 
0 

0 

s 

.°ci 

0 

CJ 

0 
0 

S 

0 
S 

1 

1 

3 
S3 

'3 

.g 

=0 

1 

ttj 

0 

0 

"C 

5 

0 

« 

0 

A 

0 

c3 

1 

.H 

0 

0) 

04 

0 

e 

« 

« 

bC 

<; 

H 

cS  ^  a: 

E 

P-i 

Q 

0 

< 

-< 

S5 

.2 

^ 

_^ 

0 
"0 

0 
•2 

.2 
"E 

c 
0 

T3 
C 
0 

-C 
c^ 

s 

0 

S 

V 

e 

"0 

'g 

^ 

_&> 

^ 

0 

£ 

s 

"s 

.s 

s 

s 

tH 

0 

(u 

s 

« 

-ts 

00 

=0 

V 

_> 

rt 

S 

S 

S 

0 

Ci 

<u 

<» 

>» 

0 

"0 

"o 

"S 

Eh 

4^ 

> 

V 

i; 

e 

c4 

c3 

< 

a. 

tt. 

Q, 

0 

P5 

Eq    ■-*    ^    ^    ;^    OQ 


^ 


^ 

ce 

c 

s 

S 

s 

0 

s 

0 

.2 

0 

8 

§ 

13 

J3 
0 

S 

^ 

3 
JO 

0 

-§ 

_e 

bi 

s 

"Ss 

5j 

0 
^ 

-«i 

». 

0 

?^ 

_o 

0 

'^ 

0 

"o 

en 

eS 

0 

0 
^ 

"5 

cS 

0 
0 

> 

CQ 

05 

0 

< 

s 


M 


80 


1.    MALONATE 


tf 


05     t— 


^  ""  o   o 


cm  lo 
o 


>o   O 


^  -2  r" 


e  ^'  fi 


*) 


c 
c 

m  o  W 


o 


'A    m     m 


C     S 


ffi 


•;r   -^  yZt 
Cm  cc  W 


O        O  -H 


C5  A 


a 


Oi 


>> 

>. 

0) 

<» 

X 

tc 

X 

03 

cfi 

03 

S  S  W  O 


—        H 


OiCiOI~-l~-lOfOOO-^— * 


s  1 


CO 


O    O    O    I-    O    O    O    O    M    -< 


c   o  o  o  ■ri 


o  o 

LO     O 


OOOOOOOOOO        OooOo: 


O  M 


cS 

^ 

-73 

_^ 

•s 

-c 

e 

o3 

cS 

S 

ft 

ft 

-0 
C 

o 

s 

2 
o 

c 
o 

"o 
o 

'3 

C 
O 

3 

o 

0) 

c 
■ft 

s 

1 

e 

•i 
S 

s 

to 

a 
1 

O 
C 

'3 
1 

ft 
o 

-(J 

ft 

D 

•2 
e 

1 

o 
"3 

s 

2 
't 

s 

H 

o 

fi 
O 

"o 
o 

> 

■fcl 

-c 
c 
o 

"o 
o 

!-. 

Hi 

C 
1 

'ft 

bO 

«i 

_c 

« 

o 

tH 

^ 

e 

£ 

=0 

•S* 

o 

» 

o 

00 

'  r\ 

"ft 

o 

5 
^ 

"e 

o 

3 

rv 

c3 

'3 

^ 

J 

■^ 

H 

CO 

a,  ^ 

^ 

&3 

~p 

O 

o 

S 

O 

o 


tf 


EFFECTS  ON  TRICARBOXYLIC  ACID  CYCLE  81 


-"'■^  fet^  io  ^-^  Z-,  ^         a  2 


_^    J    f-^  C5  5    "^ 


^    T3    ^  M  •  •    ''^ 


-.^    c 


O5-*(NO5C5f0LOTti       o^-'M'^^'00■<^^orJ^oo5(^^■*^0'—   o 


CC    CC    O' 

(M  T^       I  -H    "M    lO    -^    CO    O    O 

-H     CO 


0.0)000       O  Occ  coQOOOO        O       Ox 


eP  i3:C55s  S 


OJ     s     c     c 


P5        05P5WCI,  Cl,q^  >^'^<iici,ll;fi^        p^fe^;^ 


s 

^ 

s 

OJ 

•2 

S 

03 
_0 

s 

^ 

'm 

S 

i 

3 
S 

03 

C 

o 

be 

4^ 

Su 

o     ^**. 

^si 

5s    •« 

A 

CC 


82 


1.    MALONATE 


K 


2  2  S 


w 


<  W 


^  c 

05      O 


o    OJ 
Cm  02 


t-i    a    t- 

03      C 
cS 

o 

"3    o 


<x> 

^ 

05 

^ 

a> 

-(J 

^ 

tH 

M 

T3 

eS 

T3 

C 

> 

O      g      05 


Q  §  PQ 


-H      CO 


oo^oooooo 


13 

C 

o 

.2 

J= 

o 

_c8 

"C 

Q 

'E 

T3 

^^ 

_cS 

a 

T3 

C! 

'r- 

se 

_c3 

'C 

C 
O 

C 

o 

O 

o 
o 

a 
5P 

2 

73 

5 

1 

o 

T3 
C 
O 

OJ 

o 

o 

o 

S 

S 

^ 
« 

o 

O 

■a 

3 

o 
o 

-4J 

-1-^ 

Cm 

1/3 

o 

"g 

k< 

'rj 

03 

o 

g 

a 

03 

?H 

'3 

-2 

05 

c 

.S 

b4 

+3 

O 
73 

chl 

o 

"S- 

14) 

> 

o3 

c 

c3 

<B 

o 

tH 

o 

g 

'^ 

^ 

03 

o 
O 
> 

"ft 

-< 

o 
H 

o3 

Para 
Rat 
Rat 
Hum 

05     05 
^-   CI 


CI  |f^  05  ro 


hJ    Ph 


|1h 


Ph 


Ph 


OQOfOiMOOiMOOO 
O^iCC0l>f0-<*lCiXi 


O    ■*    O    -*    O    Tt    o 
C^l  <M  "M  (M 


OOOOOOOOO      «:'0 


% 


P     fi  jS 


-C  ^ 


!2  ^        ^ 


^  « 


p:;       p5 


P2 


73 

o 

ft 

O     . 


-2   crt 


I  ^  ^ 

^  -^  'm 

C    °  I 

<1^      43  M 

>>    C  m 

°     ^  ^ 

ft  Is 

O      03  13 

CO 


c3 


CO     c 


o     a> 

o    _ 


«    T) 


<u    ;:? 


;^ 


5. 


="  73  ® 

:S  ^  o 

.2  "  ts 

03  TJ  aJ 


.2  ^ 


•—       .,       O     I— I 
C     1^      C    t 

C    >    o  ^ 


s  -^ 


o 

03 

-|J 

■3 

C 

eS 

a 

o 

Lh 

u 

+3 

cS 

c 

o 

s 

8 

o 

43 

fl 

8 

43 

s 

cS 

s 

D 

T3 

O 

C 

-a 

o3 

^  s 


EFFECTS  ON  TRICARBOXYLIC  ACID  CYCLE  83 

5  mM  malate.  However,  when  the  malate  concentration  is  between  0.1  and 
1  mM,  the  inhibition  is  close  to  50%  (Montgomery  and  Webb,  1956  b). 

(6)  Inhibition  of  a-ketoglutarate  oxidation.  The  possible  inhibition  of  a-ke- 
toglutarate  oxidase  by  malonate  is  important  not  only  because  of  the  bearing 
it  has  on  the  effects  of  malonate  on  the  operation  of  the  cycle,  but  also 
because  malonate  has  been  frequently  used  to  block  succinate  oxidation 
in  order  to  study  in  particulate  systems  the  oxidation  of  a-ketoglutarate 
uncomplicated  by  further  oxidations.  This  technique  was  first  proposed  by 
Ochoa  (1944),  who  showed  that  high  concentrations  (25-50  mM)  of  malon- 
ate would  allow  a-ketoglutarate  to  be  oxidized  to  succinate  in  enzyme  pre- 
parations from  cat  heart.  In  four  experiments  with  50  mM  malonate,  0.86 
mole  of  succinate  was  formed  for  every  mole  of  a-ketoglutarate  utilized, 
indicating  that  even  here  an  appreciable  fraction  of  the  succinate  formed 
is  further  oxidized.  No  data  were  given  as  to  whether  these  concentrations 
of  malonate  inhibit  the  utilization  of  a-ketoglutarate.  Slater  and  Holton 
(1954)  used  10  mM  malonate  to  study  the  oxidation  of  a-ketoglutarate 
in  heart  mitochondria,  and  it  was  shown  that  this  concentration  does  not  re- 
duce the  utilization  of  a-ketoglutarate,  although  20-40  mM  does  inhibit  it. 
Malonate  was  also  used  to  investigate  the  formation  of  a-ketoglutarate  from 
citrate  in  Micrococcus  sodonensis  (Perry  and  Evans,  1960),  but  the  rationale 
for  this  is  obscure  since  malonate  by  inhibition  of  succinate  oxidation  would 
not  depress  the  disappearance  of  a-ketoglutarate.  However,  at  the  concen- 
tration of  malonate  used  (75  mM),  it  is  quite  possible  that  the  a-ketoglu- 
tarate oxidase  was  inhibited. 

The  oxygen  uptake  with  a-ketoglutarate  as  the  substrate  in  particulate 
preparations  from  several  sources  has  been  shown  to  be  reduced  by  malonate 
as  expected  if  the  succinate  formed  is  partially  protected  from  oxidation 
(see  Table  1-14).  Malonate  concentrations  around  10  mM  inhibit  40-60% 
in  most  cases.  No  definite  information  on  the  possible  inhibition  of  the 
a-ketoglutarate  oxidase  can  be  obtained  from  such  studies. 

If  the  oxidation  of  a-ketoglutarate  stops  at  fumarate,  the  system  is  a 
two-step  linear  chain  (neglecting  the  other  reactions  involved  in  the  forma- 
tion of  succinate).  An  atom  of  oxygen  is  taken  up  in  each  step: 

1/2  Oa  1/2  O. 

a-Ketoglutarate     ->     succinate     ->■     fumarate 
(1)  (2) 

SO  that  a  complete  and  specific  inhibition  of  reaction  (2)  would  result  in  a 
maximal  inhibition  of  50%  with  respect  to  the  oxygen  uptake.  However,  in 
case  the  first  reaction  is  much  faster  than  the  second,  malonate  would  not 
inhibit  the  initial  rate  of  the  reaction,  even  though  it  inhibited  the  final  total 
oxygen  uptake.  Therefore,  the  inhibition  may  theoretically  vary  from  0  to 


84  1.    MALONATE 

50%,  depending  on  the  relative  rates  of  the  reactions  and  the  period  during 
which  the  oxygen  uptake  is  measured.  If  the  fumarate  is  further  oxidized, 
more  oxygen  will  be  consumed  and  the  inhibition  by  malonate  may  be 
greater  than  50%.  Furthermore,  it  is  essentially  impossible  to  inhibit  the 
succinate  dehydrogenase  completely,  especially  as  succinate  will  accumulate 
and  progressively  overcome  the  inhibition. 

The  use  of  malonate  to  study  a-ketoglutarate  oxidation  in  mitochondria 
involves  the  assumption  that  malonate  does  not  significantly  affect  the  a- 
ketoglutarate  oxidase  directly.  Unfortunately,  no  investigations  of  the  in- 
hibition of  a-ketoglutarate  dehydrogenase  by  malonate  have  been  report- 
ed, and  thus  it  is  difficult  to  compare  the  sensitivities  of  the  two  dehydro- 
genases. Several  studies  have  determined  the  effects  of  malonate  on  the 
disappearance  of  a-ketoglutarate  (a-KG)  during  periods  when  the  oxygen 
uptake  is  depressed,  and  it  is  rather  strange  that  some  effect,  either  positive 
or  negative,  has  always  been  reported.  The  pertinent  data  have  been  sum- 
marized in  the  accompanying  tabulation,  and  it  is  seen  that  an  inhibition 


Preparation 

Malonate 

(mM) 

0 
0 

Inhibition  of 

O2  uptake 

a-KG  utilization 

Streptomyces  coelicolor 

10 

41 

Stim     9 

Pea  seedling  mitochondria 

10 

66 

21 

Blowfly  sarcosomes 

10 

10 

Stim  25 

Heart  mitochondria 

10 

33 

Stim    8 

Rat  brain  homogenate 

3.3 

59 

29 

of  a-ketoglutarate  disappearance  is  observed  in  some  cases  and  a  stimulation 
in  others.  The  oxidation  of  a-ketoglutarate  depends  on  Mg++  and  it  is 
possible  that  the  differences  are  related  to  the  degree  of  Mg++  requirement 
and  the  concentrations  of  Mg++  used  in  the  assay  media.  Price  (1953) 
found  that  even  1  mM  malonate  would  inhibit  a-ketoglutarate  utilization 
in  pea  seedling  mitochondria  and  that  30  raM  had  a  very  marked  effect 
(60%  inhibition).  This  is  not  a  competitive  type  of  inhibition  because 
increasing  the  a-ketoglutarate  concentration  does  not  lower  the  inhibition, 
and  even  increases  it  somewhat.  The  possibility  of  Mg++  reduction  was 
explored  and  it  was  found  that  the  calculated  drop  in  the  Mg++  concen- 
tration could  not  have  been  responsible  for  the  inhibition.  Furthermore,  it 
was  not  possible  to  reverse  the  inhibition  by  increasing  the  Mg^^  concen- 
tration from  1  mM  to  4  mM.  An  inhibition  by  a  Mg-malonate  complex 
was  considered  to  be  the  most  likely  explanation,  but  yet  no  inhibition 
was  seen  when  the  concentration  of  Mg++  in  the  medium  was  reduced  to 
zero,  although  the  enzyme  was  still  partially  complexed  with  Mg+"'".  What- 


EFFECTS  ON  TRICARBOXYLIC  ACID  CYCLE  85 

ever  the  explanation,  it  is  obvious  that  in  such  mitochondria  malonate 
could  not  be  used  to  isolate  a-ketoglutarate  oxidation. 

Another  quantitative  study  on  the  effects  of  malonate  on  a-ketoglutarate 
oxidation  was  made  by  Grafflin  et  al.  (1952),  who  were  attempting  to  find 
a  good  assay  system  for  a-ketoglutarate  oxidase  in  rabbit  kidney  homogen- 
ates.  They  concluded  that  the  use  of  malonate  is  unsatisfactory  and  aban- 
doned this  procedure.  The  difficulty  lies  particularly  in  the  inability, 
except  at  high  malonate  concentrations  (around  30  mM),  to  inhibit  com- 
pletely the  oxidation  of  succinate,  as  determined  from  the  total  oxygen 
uptake  compared  with  the  theoretical  value  for  a  one-step  oxidation  of 
a-ketoglutarate.  Although  no  evidence  on  the  effect  of  malonate  on  the 
a-ketoglutarate  oxidase  was  presented,  it  would  be  surprising  if  concentra- 
tions of  malonate  above  20  mM  had  no  effect.  Lewis  and -Slater  (1954) 
also  remarked  that  even  in  the  presence  of  10  milf  malonate,  the  oxygen 
uptake  from  the  oxidation  of  a-ketoglutarate  greatly  exceeds  the  disap- 
pearance of  a-ketoglutarate,  .lO/.Ja-KG  ratios  generally  being  above  2, 
in  blowfly  sarcosomes.  Also,  the  oxygen  uptake  over  35-45  min  is  depressed 
only  10%  at  this  concentration  of  malonate. 

It  is  difficult  to  understand  why  the  oxidation  of  succinate  is  so  resistant 
to  malonate  under  these  circumstances.  Taking  the  values  of  Kj  that 
have  been  found  in  mammalian  tissue  studies,  malonate  at  10  mM  should 
inhibit  well  over  95%  even  at  succinate  concentrations  that  might  occur 
experimentally.  For  example,  in  beef  heart  preparations,  where  K^  is  about 
0.04  WlM,  10  vnM  malonate  would  inhibit  over  99%  at  a  succinate  concen- 
tration of  2  mM.  It  may  be  that  in  the  intact  system  the  oxidation  of  endo- 
genously  formed  succinate  from  a-ketoglutarate  via  succinyl-CoA  is  ki- 
neticaUy  different  than  the  oxidation  of  exogenous  succinate,  or  that  local 
concentrations  of  succinate  can  reach  much  higher  levels  than  predicted 
on  the  basis  of  over  all  analyses.  One  must  conclude  at  least  at  the  present 
time  that  the  specific  inhibition  of  succinate  oxidation,  even  in  these  rela- 
tively simple  systems,  is  generally  impossible. 

This  problem  has  been  approached  recently  by  Jones  and  Gutfreund 
(1964),  who  experimentally  determined  the  steady-state  concentrations  of 
succinate  in  guinea  pig  liver  mitochondria  during  the  oxidation  of  a-keto- 
glutarate. The  O2  uptake  of  uninhibited  mitochondria  is  40-45%  due  to 
a-ketoglutarate  oxidation,  40-45%  due  to  succinate  oxidation,  and  10-20% 
due  to  other  oxidations.  The  effects  of  malonate  on  the  oxidation  of  exoge- 
nous succinate  were  compared  with  the  effects  on  the  oxidation  of  succin- 
ate-C^*  formed  from  a-ketoglutarate-C^^.  The  rate  of  utilization  of  a-keto- 
glutarate is  not  altered  up  to  8  mM  malonate.  The  steady-state  succinate 
concentration  by  total  analysis  is  0.04  mM,  and  this  is  not  affected  by 
malonate  until  its  concentration  is  higher  than  0.04  mM;  half  the  succinate 
formed  from  a-ketoglutarate   accumulates  with   0.7  mM  malonate,   and 


86  1.    MALONATE 

80%  accumulates  at  8  mM  malonate.  The  Og  uptake  due  to  the  oxidation 
of  succinate  formed  from  a-ketoglutarate  is  reduced  50%  by  0.2  mM  mal- 
onate. The  data  indicate  that  endogenously  formed  succinate  is  at  a  much 
higher  concentration  than  that  determined  by  total  analysis.  The  inhibition 
studies  indicate  the  steady-state  succinate  concentration  in  the  mitochon- 
dria to  be  4.6-13  mM  and  the  isotopic  studies  suggest  concentrations  ex- 
ceeding 4  mM.  Such  high  succinate  concentrations  would  protect  the  suc- 
cinate dehydrogenase  and  result  in  less  malonate  inhibition  than  expected. 
Since  there  appear  to  be  no  permeability  barriers  in  the  mitochondria  to 
succinate,  the  authors  suggested  that  there  is  a  spatial  relation  between 
the  succinate  dehydrogenase  and  the  enzyme  forming  the  succinate.  If 
this  is  true,  it  raises  the  interesting  possibility  that  certain  enzymes  are 
specific  not  only  for  their  substrates  but  also  for  other  enzymes  with 
which  they  interact  to  form  functional  metabolic  complexes. 

(c)  Effects  on  tricarboxylate  utilization.  Malonate  has  variable  affects  on 
the  oxidation  of  citrate  and  isocitrate,  often  inhibiting  rather  well,  but 
sometimes  having  little  effect  or  actually  stimulating  (Table  1-14).  The 
relationship  between  malonate  inhibition  and  isocitrate  oxidation  is  prob- 
ably complex  in  most  instances.  There  seems  to  be  minimal  direct  inhibi- 
tion of  isocitrate  dehydrogenase,  although  data  are  lacking.  Hiilsmann 
(1961)  showed  that  malonate  at  11  mM  reduces  quite  potently  the  utiliza- 
tion of  isocitrate  by  rabbit  heart  mitochondria  when  acetate  or  /?-hydroxy- 
butyrate  is  the  additional  substrate  (these  accelerating  the  utilization  of  iso- 
citrate). The  stimulation  of  isocitrate  utilization  by  these  substrates  was 
claimed  to  be  due  to  the  formation  of  acetoacetyl-CoA,  an  intermediate  in 
fatty  acid  synthesis,  and  this  alters  the  oxidation-reduction  states  of  NAD- 
NADH  and  NADP-NADPH  through  the  transhydrogenase  reaction, 
isocitrate  oxidation  rate  being  dependent  on  the  concentration  of  NADP. 
Kasamaki  et  ol.  (1963)  showed  that  malonate  inhibits  citrate  and  isocitrate 
oxidations  in  Proteus  vulgaris  much  less  when  NADP  is  added.  Chappell 
(1964  a)  suggested  that  the  effects  of  malonate  on  tricarboxylate  utilization 
in  rat  liver  mitochondria  are  due  to  the  block  in  the  formation  of  malate 
from  succinate,  malate  being  necessary  for  the  oxidation  of  isocitrate,  since 
it  provides  a  means  for  reoxidation  of  NADPH  mediated  through  the 
transhydrogenase  and  malate  dehydrogenase.  We  see  here  a  possibly  im- 
portant control  in  the  operation  of  the  cycle  and  how  malonate  can  exert 
effects  indirectly  on  steps  distant  from  the  succinate  oxidation  reaction, 

{d)  Effect  of  cycle  substrate  concentration  on  the  inhibition.  Some  claims 
have  been  made  that  increases  in  the  citrate  concentration  will  tend  to  over- 
come inhibition  of  the  oxygen  uptake  by  malonate,  and  sometimes  this 
has  been  attributed  to  a  competitive  effect,  it  being  assumed  that  the  higher 
concentration  of  substrate  will  give  rise  to  a  higher  succinate  level.  Thus 


EFFECTS  ON  TRICARBOXYLIC  ACID  CYCLE  87 

Laties  (1953)  found  that  increasing  the  citrate  from  1  raM  to  10  mM 
would  reduce  the  inhibition  by  10  mM  malonate  from  around  52%  to  zero  in 
cauliflower  homogenates,  and  Pierpoint  (1959)  reported  that  a  10-fold  rise 
in  citrate  concentration  brings  about  a  decrease  from  62  to  45%  in  the 
inhibition  by  21  mM  malonate  in  tobacco  leaf  mitochondria.  Laties  pointed 
out  that  this  could  not  be  due  to  a  competitive  effect  because  of  an  increased 
production  of  succinate,  since  at  the  malonate  concentration  used  the 
oxidation  of  10  mM  succinate  is  completely  inhibited.  However,  if  what 
was  suggested  in  the  previous  section  regarding  the  difficulty  in  the  inhi- 
bition of  the  oxidation  of  endogenously  produced  succinate  is  valid,  data 
on  the  inhibition  of  exogenous  succinate  may  not  be  applicable.  Laties  felt 
that  the  explanation  might  lie  in  the  interference  between  electron-trans- 
port systems,  so  that  when  citrate  concentration  is  high,  the  contribution 
made  by  succinate  oxidation  to  the  total  respiration  is  less.  Clear-cut  effects 
of  concentration  on  inhibition  have  not  been  observed  with  a-ketoglutarate 
(Grafflin  et  al.,  1952;  Pierpoint,  1959),  pyruvate  (Smyth,  1940),  acetate 
(Jowett  and  Quastel,  1935  c),  or  malate  (Pierpoint,  1959). 

In  experiments  of  this  type,  it  is  well  to  remember  that  the  pattern  of 
oxygen  uptake  may  change  with  the  concentration  of  the  substrate.  The  rela- 

xO  yO 

S     ->     succinate     ->•     P 

tive  amounts  of  oxygen  taken  up  before  and  after  the  succinate  step  may  be 
altered  by  substrate  concentration  due  to  factors  previously  discussed  in 
this  chapter.  Thus  the  ratio  x/y  in  the  above  equation  will  vary  and  hence 
the  effect  of  an  inhibitor  of  succinate  oxidation.  If  the  inhibition  on  succinate 
oxidation  is  i  and  the  ratio  xjy  =  r,  the  over  all  inhibition  on  the  total  oxy- 
gen uptake  will  be  ijil  +  r),  so  that  anything  that  changes  r  will  change  the 
inhibition.  If  high  concentrations  of  the  substrate  produce  an  accumulation 
of  some  intermediate  (as  citrate  rises  when  pyruvate  enters  the  cycle  rapidly), 
r  will  increase,  at  least  over  the  initial  period,  and  the  inhibition  will  be 
less  at  lower  substrate  concentrations. 

(e)  Stimulation  of  cycle  substrate  utilization  by  malonate.  In  a  number  of 
cases  there  is  clearly  a  stimulation  of  the  utilization  of  pyruvate,  acetate, 
or  citrate  by  malonate.  Sometimes  this  is  recognized  in  an  increased  oxygen 
uptake  but  occasionally  the  disappearance  of  the  substrate  is  accelerated 
while  the  oxygen  uptake  is  inhibited.  Often  this  effect  is  very  marked.  In 
the  mycelia  of  Ashbya  gossypii,  oxygen  uptake  due  to  addition  of  acetate  is 
stimulated  48%  by  4  m.M  malonate,  whereas  40  m.M  malonate  inhibits 
53%  (Mickelson  and  Schuler,  1953).  The  increase  in  the  respiration  from 
pyruvate  in  bull  sperm  by  10  mM  malonate  is  almost  as  great  (Lardy  and 
Phillips,  1945).  Table  1-14  cites  a  number  of  other  instances. 


88  1.    MALONATE 

The  mechanisms  for  such  stimulations  must  vary  with  the  particular  sub- 
strate used,  but  it  may  be  useful  to  suggest  some  possible  ways  in  which 
malonate  could  produce  this  apparently  anomalous  effect.  (1)  If  the  prepara- 
tion has  an  active  oxalacetate  decarboxylase,  this  may  reduce  the  concentra- 
tion of  oxalacetate  for  condensation  with  acetyl-CoA.  Malonate  is  able  to 
inhibit  this  enzyme  in  some  cases  (Table  1-12)  (Pardee  and  Potter,  1949), 
in  which  case  oxalacetate  may  be  protected  so  that  cycle  entry  of  acetyl- 
CoA  is  facilitated.  (2)  If  ADP  concentration  is  low  and  ATP  concentration 
high  normally  in  the  preparation,  malonate  by  inhibiting  certain  phases  of 
the  cycle  might  increase  ADP  concentration  and  thus  stimulate  electron 
transport  in  other  oxidative  processes  by  providing  more  phosphate  ac- 
ceptor. (3)  If  there  is  competition  between  the  different  oxidative  reactions 
in  the  cycle  for  some  common  coenzyme  or  cofactor  (e.g.  NAD,  NADP,  or 
Co  A),  inhibition  of  some  oxidations  by  malonate  might  allow  these  factors 
to  be  used  more  readily  by  other  systems.  Such  competition  between  the 
pyruvate  and  a-ketoglutarate  systems  has  been  suggested  in  heart  mito- 
chondria (Montgomery  and  Webb,  1956  a).  (4)  If  malonate  is  metabolized  by 
the  preparation,  it  is  possible  that  a  product  would  accelerate  the  utiliza- 
tion of  some  cycle  intermediate.  In  some  organisms  malonate  can  form  acetyl- 
CoA  and  acetate,  as  well  as  other  products.  (5)  In  intact  cells,  malonate 
might  increase  the  permeability  of  the  cell  membrane  so  that  certain  sub- 
strates, such  as  pyruvate,  citrate,  or  a-ketoglutarate,  could  enter  more 
readily.  (6)  By  chelation  with  inhibitory  metal  ions  that  may  occur  in  the 
preparation,  malonate  might  accelerate  the  rates  of  certain  reactions.  All  of 
these  mechanisms  are  purely  hypothetical,  since  in  no  case  has  the  actual 
mechanism  been  established. 

Intracellular  Concentrations  of  Cycle  Intermediates 

Interpretation  of  intracellular  inhibition  by  malonate  and  other  inhi- 
bitors acting  on  the  cycle  should  ideally  involve  in  many  cases  a  knowledge 
of  the  concentrations  of  certains  intermediates.  Data  collected  for  different 
types  of  cells  are  shown  in  Table  1-15.  These  figures  were  calculated  on 
the  basis  of  intracellular  water  contents.  However,  it  is  likely  that  these 
substances  are  not  distributed  homogeneously  throughout  the  cell  water. 
There  is  evidence  that  some  intermediates  may  occur  within  the  mito- 
chondria at  different  concentrations  than  in  the  surrounding  medium.  Thus 
the  mitochondria/medium  ratios  for  sheep  kidney  mitochondria  under 
certain  conditions  were  found  to  be:  pyruvate  0.84,  fumarate  7.42,  a-keto- 
glutarate 1.0,  citrate  0.83,  and  oxalacetate  0.13  (Bartley  and  Davies, 
1954).  Furthermore,  the  values  given  in  the  table  are  all  abnormal  since 
truly  normal  tissues  were  not  used.  The  rats  were  fasted  for  24  hr  while 
the  suspensions  of  E.  coli  and  yeast  were  metabolizing  acetate  rather  than 
a  more  normal  substrate.  In  normal  rat  tissues  the  concentrations  may  well 


EFFECTS  ON  TRICARBOXYLIC  ACID  CYCLE 


89 


hJ 

•'" 

& 

>-^ 

o 

^ 

^ 

s 

■■ 

w 

r! 

H 

.9 

< 

cS 

O 

H 

fl 

§ 

o 

Z 

o 

Q 

o 


fe) 


w 


m 


C5      CO      — 
■<*     -^     IC 


o   o   o   o 


COCSCOCOJOO^M'M 
OOO-H-<G0     —     -H 

oooooooo 


V 


o   o   o   o 


O'^^OOOOC-'OO 


O     iM     O     O     O     O     O 


ci      IB 


-S    8  I 


P^    h^    O      o    h5 


bn 

rl 

o 
-to 

C 

cS 

4; 

W 

o 

S 

3 

M 

[^ 

§  o 


o 

^ 

H-J 

cfl 

>. 

-o 

-Q 

SJ 

^ 

_o 

rt 

> 

-a 

.2    .2 


«) 

.a 

« 

c 

J2 

o 

"2 

R 

^ 

_         r-  ^ 

^      JZ^  TO 


:^     *i       ^ 


2  ^ 


.o  -^    .,  -^ 


~  o 


"c  -o    g    a  c 


3    "*      D 


p  •«    oj     c 
fS  'S  H     ^ 


'J^ 


tH  ^  ^ 


.2    —      M 


^     S^ 


—      cc     ,0      ~ 


P  §^   I    5 


o  ■— 


73 


3    S   rf  -« 


<U  Q  o  "S 

-§  2  t^  ?^  ^ 

3  ft  3  -43  N 

o  TO  o  oj  ;z; 

13  »  2  <"  ° 

"  i3  ^  -2 

,  ci  ai  a>  rt 

1 .5  :^  1 1 

cS  cS  o  ?  S 

^  ^  =  ^  « 


t-  o 


^    J=    -^     .S     <D      C 


oj      OJ      t-    — 


c 


o    „     o      C     t-    ^ 
S    ~     tc      O     4^ 

o    >    2   O    c    ^ 


•r    cs 


S  ^   t2 


o   Sti 


o    51 


S     g 


Q    " 
o    t 


S  S.  »  s 

CO  a)  2  '^ 

"'-  -c  -5  A 

-^  ^  ^  -J3 


c5      rt 


Q  15  Q 


90  1.    MALONATE 

be  higher,  and  in  yeast  metabolizing  sugar  the  values  may  be  lower.  One 
value  for  oxalacetate  in  normal  rat  liver  is  available  (0.036  mikf)  and  in 
this  case  fasting  for  24  hr  did  not  alter  this  appreciably  (Kalnitsky  and 
Tapley,  1958). 

These  values  do  not  represent  a  thermodynamic  equilibrium  based  on  dif- 
ferences in  free  energy,  but  rather  a  dynamic  or  kinetic  equilibrium,  depend- 
ing mainly  on  the  relative  rates  of  the  cycle  reactions  and  competing  process- 
es. The  higher  concentration  of  fumarate  compared  to  malate  in  rat  tissues 
illustrates  this  because  in  a  thermodynamic  equilibrium  there  would  be 
about  one-fourth  the  concentration  of  malate.  It  is  interesting  that  the  values 
differ  so  widely  from  tissue  to  tissue  and  certainly  this  must  be  one  factor  in 
determining  the  different  responses  to  malonate  or  other  competitive  inhi- 
bitors. The  levels  of  succinate  are  generally  low,  except  in  yeast,  but  as 
pointed  out  above  the  concentrations  within  the  mitochondria  or  at  the  re- 
gion of  the  active  center  of  succinate  dehydrogenase  may  well  be  higher. 
The  effects  of  malonate  on  the  concentrations  of  these  intermediates  in  cells 
will  be  taken  up  in  the  following  section. 

Analyses  of  plant  tissues  have  not  been  presented  here  because  there  is 
some  doubt  as  to  the  significance  of  the  figures.  Most  plants  contain  large 
amounts  of  organic  acids,  including  the  cycle  intermediates.  Beevers  (1952) 
postulated  that  the  cj^cle  in  plants  is  less  readily  blocked  than  in  animal 
tissues  because  of  these  high  concentrations  of  succinate  and  other  cycle 
intermediates.  This  could  well  be  an  important  factor,  but  actually  the 
concentrations  of  these  acids  in  the  plant  cytoplasm  are  not  known  in 
most  cases,  total  analyses  including  the  vacuolar  fluid,  which  is  often  of 
greater  volume  than  the  cytoplasm  and  contains  most  of  the  organic  acids. 
There  is  another  way  by  which  these  plant  acids  could  protect  against 
malonate.  The  presence  of  large  amounts  of  fumarate  or  malate,  or  of  any 
substance  capable  of  forming  oxalacetate,  would  allow  pyruvate  to  be  incor- 
porated into  the  cycle  even  in  the  state  of  complete  block  of  succinate  oxi- 
dation. Examples  of  the  overcoming  of  malonate  inhibition  by  fumarate 
and  malate  will  be  presented  shortly. 

ACCUMULATION    OF   SUCCINATE    DURING    MALONATE 

INHIBITION 

An  effective  inhibition  of  succinate  oxidation  should  lead  to  a  rise  in 
the  concentration  of  succinate  under  conditions  in  which  succinate  can  still 
be  formed.  Such  accumulation  of  succinate  has  been  frequently  observed 
and  some  of  the  more  quantitative  results  are  summarized  in  Table  1-16. 
In  addition  to  the  examples  in  the  table,  accumulation  of  succinate  has 
been  reported  in  the  following  species  and  tissues:  Shigella  (Yee  et  al., 
1958),  Nocardia  (Cartwright  and  Cain,  1959),  Aspergillus  (Shimi  and  Nour 


ACCUMULATION    OF    SUCCINATE  91 

El  Dein,  1962),  tobacco  leaves  (Vickery,  1959;  Vickery  and  Palmer,  1957), 
potato  slices  (Romberger  and  Norton,  1961),  avocado  mitochondria  (Avron 
and  Biale,  1957),  pea  leaf  particulates  (Smillie,  1956),  barley  roots  (Laties, 
1949  b),  Colpidium  (Seaman,  1949),  Trypanosoma  (Bowman  et  al.,  1963), 
carp  liver  mitochondria  (Gumbmann  and  Tappel,  1962  b),  rat  heart  ho- 
mogenates  (Lehninger,  1946  b),  rat  liver  slices  (Elliott  and  Greig,  1937), 
human  heart  slices  (Burdette,  1952),  ascites  carcinoma  cells  (Dajani  et  al., 
1961),  and  in  many  tissues  of  rats  and  rabbits  (Busch  and  Potter,  1952  a; 
Forssman,  1941).  In  the  experiments  leading  to  the  results  in  Table  1-16,  the 
preparations  were  incubated  for  one  to  several  hours  with  malonate  and  the 
succinate  analyzed  at  the  and  of  the  incubation,  so  that  the  rates  of  succinate 
formation  at  any  time  are  difficult  to  evaluate,  and  may  well  have  been 
greater  initially.  The  over  all  succinate  concentrations  may  be  estimated 
from  the  volumes  in  which  the  expriments  were  run  and  in  most  cases  the 
final  succinate  concentrations  range  between  0.5  and  2.5  mM. 

Several  points  are  brought  out  by  the  results  in  Table  1-16.  It  is  seen 
that  succinate  can  be  formed  from  essentially  all  the  cycle  substrates  and 
intermediates  in  the  presence  of  malonate.  Rapid  rates  are  found  when 
oxalacetate  or  some  substance  forming  oxalacetate  is  added  with  pyruvate, 
as  would  be  expected,  because  in  the  absence  of  a  source  of  oxalacetate,  the 
malonate  would  reduce  the  incorporation  of  pyruvate  into  the  cycle  and 
hence  the  rate  of  formation  of  succinate.  It  may  be  noted  in  some  cases 
that,  in  the  absence  of  added  substrates  or  malonate,  some  succinate  ac- 
cumulates (yeast,  Avena  coleoptile,  spinach  leaves,  and  dog  heart),  which 
implies  that  under  the  experimental  conditions  succinate  is  formed  more 
rapidly  than  it  can  be  oxidized.  This  is  somewhat  surprising  because  it  is 
usually  assumed  that  the  activity  of  succinate  oxidase  is  quite  high  in 
most  tissues.  The  possibility  of  the  accumulation  of  sufficient  oxalacetate 
to  inhibit  succinate  dehydrogenase  when  little  acetyl-CoA  is  available 
cannot  be  ignored.  This  phenomenon  is  also  evident  in  the  analyses  for 
succinate  given  in  Table  1-15.  The  interesting  effects  of  malonate  concen- 
tration are  seen  in  two  investigations.  In  spinach  leaves,  the  maximal  suc- 
cinate accumulation  occurs  at  malonate  concentrations  around  or  below 
50  milf;  at  higher  concentrations,  the  malonate  is  apparently  acting  on 
other  enzymes  in  the  cycle  and  reducing  the  rate  of  formation  of  succinate. 
Likewise,  in  brain  minces,  the  high  concentration  of  200  mM  malonate  is 
seen  to  depress  succinate  accumulation. 

Quantitative  conversion  of  cycle  substrates  to  succinate  in  the  presence 
of  malonate  is  generally  not  observed.  In  fact,  in  most  cases  in  which  the 
disappearance  of  substrate  was  determined  simultaneously  with  the  for- 
mation of  succinate,  only  a  small  fraction  appeared  as  succinate.  For  exam- 
ple, Speck  et  al.  (1946)  found  in  malarial  parasitized  erythrocytes  that 
only  22%  of  the  pyruvate  utilized  in  the  presence  of  20  mM  malonate  was 


92 


1.    MALONATE 


:^       ^ 


05 


>. 


m 


o 

03 


0000'i^C<ICO'005-H(MfOTti-HCC'^ 


11 

cS    3 


+ 


<       < 


CO 

JO 

C<1 

1^1 

o 

o 

2- 

2i 

^      cj      0      o 


c     c     ai     OJ 
S     3     o     o 

&H  fe  <:  <: 


CLi     (i, 


Oj  ©       -H        -H 


r-i        rS        cS        C3 
r-  3  C3  C6 

fa  p&  s  s 


ACCUMULATION    OF    SUCCINATE 


93 


02 


1^      7^1      C5      l^      O 

—     (M     -*     O     -^  ~ 


C*^        ^H         ■»!^         3         CO 


M     M     O     Lt     O     O 

r:    --^    't    c    rc    C5 
o    —    ':o    —    -H    tJ< 


+   + 


<!  <j3 


a;     <B    ^ 


Ch     Ph     § 


+  + 

S"  o  3  2  S^  S" 


i'        t^        Z        ~        C        C 


Ch    Ph 


PL, 


o 
-1-2 


^     ^     Ph     fl. 


fa 


94 


1.    MALONATE 


(X 


bO 

o 

C 
c8 


o 

M 

K 

T1 

TJ 

K 


tiJ 


a 


rt     (M     (M     — < 


O    — '  O    — ' 


(^rtO-^-*CO-Ht>C5 


lO    lO    LO    >c 

■ri     "M     5^1     C^l     o 


+  +    +  +    +  + 


+ 


o 

o 

(3, 

.— , 

0 

0 

. 

. 

0 

^^ 

2. 

■" 

-^ 

^ 

"^ 

"-^ 

:;;^ 

~-^ 

:;^ 

M5 

OJ 

« 

« 

0) 

Oi 

0 

OJ 

» 

© 

<u 

(Kl 

^ 

f« 

2 

C3 

* 

03 

03 

CS 

ce 

o3 

(S 

ee 

^ 

> 

^ 

^ 

^ 

> 

r^ 

> 

> 

-tS 

oa 

S 

^ 

^ 

3 

D 

^ 

3 

^ 

te 

a 

k. 

u 

;-i 

b< 

^ 

b 

>-. 

>, 

>- 

>•- 

>. 

2 

>> 

>> 

s 

CM 

Cm 

fl^ 

ei^ 

<^ 

;a^ 

^ 

t>H 

^ 

eu 

0 

Pm 

03 


ACCUMULATION    OF    SUCCINATE 


95 


00    o 
lO    00    ■<* 


O     00     CO 

■*       -H       C^ 


<MOI>'*00'-<  OCa<M 


iC   >c   o    o    o   o 

(M     <M     ifi     »0     O     O 


IC    lO    o    o    o    o 

Cq     <M     lO     iO     O     O 


O 

Q 


W 


H    5 


o   'C 

V0      0) 


0)  «      c  si 


c    ^ 


cS 

c« 

a 

rt 

S 

c 

Si 

3 

d 

<«4 

<4H 

3 

s 

+ 

+ 

Ci 

c«^ 

^^ 

^^ 

, . 

+ 

+ 

lO 

IC 

lO 

<M 

•M 

(M 

O 

0 

0 

::;i 

:3i 

:;::;, 

■3;!, 

:;i 

^ 

ffii 

(D 

i» 

0 

« 

(D 

-f^ 

4J 

*i 

-u 

-ti 

•ii 

a 

eS 

cS 

ci3 

cS 

ce 

> 

> 

> 

> 

> 

> 

? 

S 

3 

3 

3 

3 

Ph   PL|   Pm   Ph   Ph   Pm 


-5 

0 

0 

r5 

0 

0 

rj 

<M 

<M 

— ■ 

^-^ 

<D 

CJ 

<» 

-t^ 

crt 

cS 

crt 

> 

> 

3 

3 

ki 

t- 

>i 

>. 

>. 

Ih 

P4 

(^ 

w 


w 


l:i1 


oj    _tn 


>■,     * 


96  1.    MALONATE 

recoverable  as  succinate.  The  highest  conversion  efficiency  was  observed  by 
Krebs  and  Eggleston  (1940)  in  pigeon  muscle  brei,  where  75-85%  of  the 
pyruvate  utilized  in  the  presence  of  fumarate  and  12.5  mM  malonate  went 
to  succinate.  Complete  conversion  to  succinate  would  not,  of  course,  be  ex- 
pected unless  one  could  inhibit  succinate  oxidation  completely  and  specifi- 
cally, which  in  most  instances  cannot  be  done. 

Factors  Determining  Succinate  Accumulation 

The  effects  of  inhibition  on  the  concentrations  of  intermediates  in  multi- 
enzyme  systems  have  been  treated  in  Chapter  1-7.  Some  of  the  most  impor- 
tant factors  involved  in  malonate  inhibition  will  be  summarized. 

(I)  Degree  of  inhibition  of  succinate  oxidase:  hence,  the  concentration 
of  malonate,  the  affinity  of  the  enzyme  for  malonate,  and  the  ability 
of  malonate  to  penetrate  if  the  preparations  are  cellular. 

(II)  Rate  of  formation  of  snccinate:  this  will  depend  primarily  on  the  availa- 
bility of  cycle  substrates  and  their  concentrations. 

(III)  Action  of  malonate  on  enzymes  other  than  succinate  dehydrogenase: 
such  actions  may  slow  down  the  formation  of  succinate,  as  discussed 
above. 

(IV)  Other  pathways  of  succinate  metabolism:  several  reactions  of  succinyl- 
CoA  or  succinate  are  known  and  these  would  tend  to  prevent  accumula- 
tion. 

(V)  Diffusion  of  succinate  from  cells:  when  only  cells  or  tissues  are  analyzed 
for  succinate,  intracellular  accumulation  will  be  reduced  by  the  loss  of 
succinate  into  the  medium  or  the  blood. 

(VI)  Time  after  addition  of  malonate:  although  this  has  never  been  studied, 
it  is  probable  that  the  succinate  concentration  will  follow  characteristic 
time  courses  in  each  case,  in  some  cases  perhaps  decreasing  after  a  peak 
level  has  been  reached. 

Another  factor,  about  which  nothing  is  known,  is  the  possible  effect  of 
rising  succinate  concentration  on  the  reactions  forming  succinate.  The 
oxidation  of  a-ketoglutarate  forms  succinyl-CoA,  and  succinate  can  arise  in 
at  least  three  different  ways  from  succinyl-CoA. 

P-enzymes 


Q  -Ketoglutarate  — >-  sue  cinyl  -  Co  A *-  sue  einate 

transaeylases 


ACCUMULATION    OF    SUCCINATE  97 

What  effects  succinate  concentration  might  have  on  these  reactions  are 
unknown,  but  some  inhibition  during  succinate  accumulation  is  possible, 
although  a  simple  backing-up  of  the  a-ketoglutarate  -^  succinate  reaction 
would  be  thermodynamically  unlikely. 

The  relative  potential  rates  of  a-ketogluti^rate  oxidase  and  succinate 
oxidase  under  normal  intramitochondrial  conditions  are  not  known,  but  the 
oxidation  of  succinate  is  certainly  one  of  the  most  rapid  reactions  seen  in 
mitochondrial  suspensions.  It  may  be  that  there  is  no  accumulation  of  suc- 
cinate in  the  cycle  under  physiological  conditions,  and  that  the  small 
amounts  of  succinate  found  in  tissues  do  not  truly  indicate  the  situation  in 
the  regions  of  succinate  oxidase.  Succinate  oxidase  is,  perhaps,  the  one 
enzyme  that  has  never  been  considered  as  normally  limiting  the  cycle  rate. 
If  the  maximal  rate  of  succinate  oxidation  is  much  higher  than  the  rate  at 
which  succinate  can  be  formed,  it  would  require  a  fairly  high  inhibition  of 
the  oxidase  before  succinate  accumulates  markedly.  For  example,  if  the  rate 
of  succinate  formation  is  one-tenth  the  rate  at  which  it  can  be  oxidized, 
90%  inhibition  of  the  succinate  oxidase  would  make  the  rates  equivalent, 
and  the  succinate  concentration  would  not  rise  very  much  (probably  not 
more  than  0.02  0.05  n\M).  Under  any  likely  conditions,  calculations  from 
Eqs.  1-7-8  and  1-7-9  make  it  clear  that  succinate  oxidation  must  be  inhibited 
fairly  strongly  to  produce  a  significant  succinate  accumulation.  The  com- 
m.on  assumption  that  succinate  must  accumulate  rather  quantitatively 
when  sufficient  malonate  has  been  added  to  inhibit  succinate  oxidase 
75-90%  is  thus  unjustified.  If  alternate  pathways  for  succinyl-CoA  or  suc- 
cinate exist,  the  accumulation  of  succinate  would  be  even  less  evident. 

Several  instances  of  failure  of  succinate  to  accumulate  during  malonate 
inhibition  have  been  reported.  Hanly  et  al.  (1952)  found  that  in  only  two 
of  six  experiments  with  carrot  root  slices  did  succinate  accumulate,  and 
in  these  the  rise  was  insignificant.  One  might  suspect  a  lack  of  penetration, 
but  15-50  mM  malonate  was  used  at  pH  4;  under  these  conditions,  malonate 
depressed  respiration  strongly.  Weil-Malherbe  (1937)  found  no  succinate 
accumulation  in  guinea  pig  brain  slices  with  malonate  4-40  mM,  respiration 
being  markedly  reduced  at  the  higher  concentrations.  A  depression  of  suc- 
cinate level  in  Streptomyces  due  to  10  vaM  malonate  was  noted  by  Coch- 
rane (1952),  with  either  malate  or  citrate  as  substrate.  Since  the  incubation 
was  16  hr  and  the  pH  5.1-5.4,  it  is  possible  that  a  nonspecific  acid  damage 
from  malonic  acid  penetration  was  responsible  for  the  cycle  depression,  or 
it  might  be  that  in  this  organism  malonate  is  not  specific  at  10  mM,  or, 
as  Cochrane  suggested,  succinate  may  not  be  formed  via  the  cycle.  The  fail- 
ure of  succinate  to  accumulate,  even  when  the  respiration  is  suppressed  by 
malonate,  is  difficult  to  explain  except  on  the  basis  of  actions  other  than 
on  succinate  dehydrogenase. 


98  1.    MALONATE 

Succinate  Accumulation  in  the  Whole  Animal 

Some  of  the  most  interesting  and  suggestive  experiments  on  succinate 
accumulation  resulting  from  malonate  inhibition  have  been  performed  with 
whole  animals  and,  although  such  work  is  often  difficult  to  interpret  in  a 
quantitative  fashion,  the  results  have  demonstrated  that  malonate  can 
partially  block  the  succinate  oxidase  in  various  tissues  of  the  living  animal. 
Such  inhibition  has  obvious  implications  for  developments  in  the  study  of 
drug  actions  and  chemotherapy.  The  first  work  of  this  type  was  done  by 
Krebs  et  al.  (1938),  who  determined  the  urinary  excretion  of  succinate,  cit- 
rate, and  a-ketoglutarate  following  injections  of  cycle  intermediates  and  mal- 
onate into  rats  and  rabbits.  Some  of  their  results  are  shown  in  Table  1-17. 
The  effects  on  citrate  and  a-ketoglutarate  will  be  discussed  later  (see  page 
104).  Although  malonate  increases  the  succinate  excretion  some  5-fold, 
only  2.9%  of  the  injected  fumarate  is  recovered  as  succinate  compared  to 
0.6%  in  the  controls.  The  injection  of  10  millimoles  of  a  substance  into  a 
rabbit  will  lead  to  a  maximal  extracellular  concentration  of  approximately 
30  TdM,  so  that  reasonably  high  concentrations  of  malonate  were  probably 
achieved.  The  effect  of  the  malonate  had  mainly  disappeared  after  24  hr 
due  to  the  excretion  and  destruction  of  the  malonate.  An  almost  10-fold 
increase  in  the  urinary  succinate  was  seen  in  the  more  recent  experiments 
of  Thomas  and  Stalder  (1958),  in  which  3.7  millimoles/kg  of  sodium  mal- 
onate were  fed  to  rats,  the  succinate  over  a  40-hr  period  rising  from  a 
control  value  of  2.35  mg  to  28.0  mg. 

The  blood  concentration  of  succinate  is  increased  in  rabbits  following 
the  injection  of  malonate  (Forssman,  1941).  Intravenous  injection  of  2.8 
millimoles/kg  of  malonate  leads  to  a  slow  rise  in  the  blood  succinate  to 
around  0.20  mM  at  3  hr,  while  injections  of  3.5-5.1  millimoles/kg  give 
levels  as  high  as  0.77  mM.  A  lethal  dose  of  8.25  millimoles/kg  produces 
death  in  35  min  and  at  the  time  of  death  the  succinate  concentration  is 
1.1  mM.  The  lower  doses  produce  no  obvious  effects  on  the  animals.  The 
normal  values  for  blood  succinate  are  about  0.025  milf. 

The  succinate  found  in  the  urine  and  blood  in  these  studies  originated 
mainly  in  the  tissues  of  the  animals.  Are  malonate  inhibition  and  succinate 
accumulation  especially  related  to  a  particular  tissue,  or  do  all  the  tissues 
contribute  to  the  metabolic  disturbance?  Can  differences  in  the  metabolic 
patterns  of  the  various  tissues  be  demonstrated  by  their  responses  to  the 
administration  of  malonate?  How  do  tumors  compare  with  normal  tissues  in 
their  susceptibility  to  malonate?  It  was  to  answer  such  questions  as  these 
that  Busch  and  Potter  (1952  a,  b)  at  the  McArdle  Memorial  Laboratory 
at  Wisconsin  undertook  their  excellent  series  of  studies  on  the  accumulation 
of  succinate  in  various  tissues  of  rats  following  injections  of  malonate. 
Analyses  for  malonate  and  succinate  were  made  by  anion  exchange  chrom- 
atography (Busch  et  al.,  1952)  at  various  times  after  the  subcutaneous 


ACCUMULATION    OF    SUCCINATE 


99 


a 


W 


O 


"l^ 

S 

-^ 

cS 

'o 

b 

2 

en 

J2 

^ 

3 

CO 

's 

« 


to 

a 

S 

ft 

o 

!«! 

c 

V 

> 

-C 

cS 

i:i5 


« 

T3 

^ 

03 

03 

3 

-l-i 

3 

£ 

-D 

4) 

^ 

» 

J2 

-c! 

Is 

H 

c 

00 

cS 

M 

Oi 

Id 

3 

o 

Ch 

e 

73 

■•«& 

C 

■» 

eS 

3 
O 

0) 

P5 

C 

o 

^ 

fe 

3 
o 

e 

,Q 

g 

100 


1.    MALONATE 


injection  of  12  millimoles/kg  of  malonate  (Fig.  1-11).  Maximal  concentra- 
tions of  malonate  are  reached  in  most  tissues  1-2  hr  after  the  injection  and 
the  time  course  of  the  succinate  levels  is  well  correlated  with  that  of  malo- 
nate. The  tissue  succinate  concentration  is  linearly  related  to  the  malonate 
concentration,  except  in  the  Flexner-Jobling  tumor  (Busch  and  Potter, 


Fig.  1-11.  Tissue  levels  of  malonate  and  succinate 
in  rats  injected  with  sodium  malonate   (12  milli- 
moles/kg). (From  Potter  et  al.,  1952). 


1952  a)  (Fig.  1-12).  The  slopes  of  the  lines  give  some  measure  of  the  degree 
of  malonate  effect  in  the  particular  tissue,  but  several  factors  are  involved 
so  that  they  are  not  quantitative  indications  of  succinate  oxidase  inhibition. 
The  urinary  excretion  of  malonate,  succinate,  and  citrate  is  shown  in  Fig. 
1-13.  The  malonate  and  succinate  levels  in  several  normal  tissues  and  tu- 
mors following  24  millimoles/kg  malonate  subcutaneously  are  given  in 
Table  1-18.  If  the  figures  in  the  table  are  multiplied  by  approximately 
0.7,  one  will  obtain  the  millimolar  concentrations  in  the  cell  water,  except 
for  kidney  and  blood,  the  former  tissue  having  extracellular  fluid  high  in 
succinate  and  malonate. 

The  ratios  of  (succinate)/(malonate)  in  the  tissues  give  essentially  the 
slopes  of  the  lines  in  plots  such  as  Fig.  1-12,  except  for  tumors  where  a 


ACCUMULATION    OF    SUCCINATE 


101 


Liver 

/ 

Tumor 

y^  Kidney 

/ 

^^ 

Q\oo^^^^^^^ 

0                                                       10  20 

Malonote  (millimoles/  kg)  - 

Fig.  1-12.  Relationships  between  malonate  and  suc- 
cinate concentrations  in  the  tissues  of  rats  injected 
with  malonate.  (From  Busch  and  Potter,  1952  a). 


Time(hr)- 


FiG.  1-13.  Urinary  excretions  of  malonate,  succinate,  and  citrate 

following  the  injection  of  12  millimoles/kg  sodium  malonate  into 

rats.  (From  Busch  and  Potter,  1952  a). 


102 


1.    MALONATE 


Table  1-18 

Tissue   Succinate   and   Malonate   Concentrations   Following    Injections   of 

Malonate  in  Rats  " 


Tissue 

Succinate 

Malonate 

Ratio 

Control 

After 
malonate 

Control 

After 
malonate 

succinate 

malonate 

Spleen 

0 

12.8 

0 

12.8 

1.00 

Liver 

2.2 

12.0 

2.4 

14.2 

0.85 

Brain 

0 

2.0 

0 

2.8 

0.71 

Thymus 

0 

9.0 

0 

21.0 

0.43 

Kidney- 

0.6 

16.4 

0.3 

44.0 

0.37 

Lung 

0 

5.6 

0 

16.0 

0.35 

Muscle 

1.0 

3.0 

2.0 

8.8 

0.34 

Heart 

0 

5.0 

0 

18.4 

0.27 

Blood 

0 

2.6 

0 

24.6 

0.11 

Tumors 

Flexner 

-Jobling 

carcinoma 

0 

9.2 

0 

17.6 

0.52 

Walker 

256  carcinoma 

0 

8.0 

0 

5.2 

1.54 

Jensen 

sarcoma 

0 

8.8 

0 

8.0 

1.10 

Hepatoma 

0 

6.0 

0 

12.0 

0.50 

Papilloma 

0 

4.4 

0 

4.4 

1.00 

Average 

0 

7.3 

0 

9.4 

0.78 

"  The  figures  are  in  /<equivalents/g  wet  weight  of  tissue.  Malonate  was  injected 
subcutaneously  at  a  dosage  of  12  millimoles/kg,  and  after  1  hr  a  similar  amount  was 
again  injected;  1  hr  following  the  last  injection  the  animals  were  sacrificed.  (From 
Busch  and  Potter,  1952  b.) 


linear  relationship  may  not  be  followed.  These  ratios  indicate  the  amount 
of  succinate  formed  per  unit  concentration  of  malonate  and  do  not  relate 
directly  to  the  degree  of  inhibition  of  succinate  oxidase  but  more  to  the 
ability  of  the  tissue  to  form  succinate  in  the  presence  of  the  inhibition. 
As  discussed  previously,  for  the  same  degree  of  block,  succinate  will  be  form- 
ed at  greatly  different  rates  in  different  tissues,  depending  on  the  succin- 
ate-forming  substrates  available  and  the  activity  of  the  pathways  leading 
to  succinate.  The  concentrations  of  malonate  in  the  tissues  vary  greatly 
and  this  must  be  a  reflection  of  the  differing  permeabilities  of  the  tissues  to 
malonate.  It  may  be  noted  that  the  concentration  in  the  brain  is  quite  low, 


ACCUMULATION    OF   SUCCINATE  103 

a  phenomenon  seen  with  most  ionic  substances,  and  this  must  account  for 
the  poor  accumulation  of  succinate  in  this  organ  and  the  relative  lack  of 
effect  of  malonate  on  central  nervous  system  function. 

The  low  concentration  of  succinate  in  the  blood  is  interesting  since  it 
implies  that  succinate  does  not  leave  the  tissues  readily.  The  conclusion 
that  must  be  reached  is  that  the  rate  at  which  succinate  diffuses  out  of  the 
tissues  into  the  blood  is  slower  than  the  rate  of  renal  excretion  of  the 
succinate.  Substances  that  are  not  resorbed  by  the  renal  tubules  are  excret- 
ed rapidly  and  their  concentrations  in  the  blood  can  be  maintained  at  a 
low  level  despite  a  continuous  influx.  Nevertheless,  it  shows  that  succinate 
leaves  the  tissue  cells  rather  slowly  under  physiological  conditions.  The  slow 
penetration  of  malonate  into  the  tissues  is  suggested  by  the  fact  that  the 
peak  levels  in  the  blood  occur  around  30  min  after  administration  whereas 
the  peak  levels  in  liver  and  tumor  occur  30  min  later  (Fig.  1-11),  the  kid- 
ney concentration  paralleling  the  blood  levels  because  of  the  excretory 
function  of  this  organ. 

The  degree  of  succinate  accumulation  does  not  necessarily  reflect  the 
cycle  activity  in  a  tissue.  For  one  reason,  succinate  can  often  be  formed 
from  other  pathways.  In  certain  tissues  the  amino  acid  content  rises  mark- 
edly (e.g.  +  215%  in  thymus  and  +  160%  in  spleen)  during  malonate 
inhibition,  while  in  others,  especially  the  Flexner-Jobling  carcinoma,  the 
amino  acids  decrease  as  the  succinate  increases.  It  is  likely  in  the  latter 
tissues  that  some  of  the  succinate  is  derived  from  amino  acids,  probably 
mainly  glutamate.  Thus  the  cycle  activity  in  this  tumor  may  be  quite  low 
and  the  normal  accumulation  of  succinate  due  to  other  sources  for  the  suc- 
cinate. The  other  tumors  do  not  show  such  marked  decreases  in  amino  acid 
content  and  this  was  attributed  to  their  greater  necrosis.  It  may  be  recall- 
ed that  the  incorporation  of  acetate  by  Flexner-Jobling  tumor  is  slower 
than  in  most  tissues  and  very  little  labeled  succinate  is  formed  from  la- 
beled acetate  (Busch  and  Potter,  1953),  indicating  a  low  degree  of  cycle 
activity.  We  have  seen  that  this  tumor  also  differs  from  normal  tissues  in 
the  nonlinearity  of  the  plot  of  tissue  succinate  against  tissue  malonate 
(Fig.  1-12).  At  low  levels  of  malonate  the  ratio  (succinate)/(malonate)  is 
near  3  but  at  high  concentrations  diminishes  to  0.5.  This  means  that  as 
the  malonate  concentration  rises,  the  ability  of  the  tumor  to  accumulate 
succinate  decreases.  This  is  the  type  of  curve  expected  if  malonate  at  higher 
concentrations  is  inhibiting  the  reactions  forming  succinate.  If  the  supply 
of  cycle  substrates  in  this  tumor  is  low,  a  relatively  small  block  of  the  cycle 
might  reduce  the  formation  of  succinate  through  the  cycle  to  zero.  The 
slope  approaches  that  of  the  blood,  and  it  is  possible  that  above  the  inflection 
point  the  slow  rise  in  succinate  may  be  due  only  to  the  rise  in  the  blood. 

This  type  of  investigation  could  well  be  applied  to  other  inhibitors 
and  certain  chemotherapeutic  agents.  First,  one  is  able  to  correlate  tissue 


104  1.    MALONATE 

concentrations  of  inhibitor  with  the  metabolic  disturbance  produced. 
Second,  the  inhibition  occurs  under  physiological  conditions,  rather  than 
in  slices  or  minces  or  other  preparations  in  which  the  cell  metabolism  may 
be  very  abnormal.  Last,  one  is  able  to  compare  the  different  tissues  with 
respect  to  their  metabolic  patterns  and  perhaps  determine  some  of  the 
reasons  for  the  selective  actions  of  inhibitors  or  drugs.  These  methods  of 
investigation,  called  "m  vivo  metabolic  blocking  techniques"  by  Busch 
and  Potter,  if  applied  properly,  would  help  to  provide  a  more  rational 
basis  for  development  in  chemotherapy  and  the  selective  depression  of 
tumor  growth. 

ACCUMULATION  OF  CYCLE  SUBSTRATES 
OTHER  THAN   SUCCINATE 

Specific  inhibition  of  succinate  oxidase  would  be  expected  to  lead  to  the 
accumulation  of  succinate  but  of  no  other  cycle  intermediates,  because 
the  free  energy  differences  between  them  are  of  such  magnitude  that  no 
backing-up  from  succinate  would  be  anticipated.  When  other  members  of 
the  cycle  are  found  to  accumulate  in  the  presence  of  malonate,  it  is  generally 
considered  to  be  evidence  that  either  the  action  of  malonate  is  not  specific 
or  that  secondary  reactions  are  proceeding.  Malonate  has  been  shown  many 
times  to  cause  an  accumulation  of  certain  cycle  intermediates,  especially 
citrate  and  a-ketoglutarate,  in  cell  suspensions,  slices,  and  whole  animals. 
The  nature  of  these  effects  will  first  be  summarized  and  then  some  possible 
mechanisms  will  be  considered. 

Accumulation   of  Citrate 

The  administration  of  malonate  to  dogs  (Orten  and  Smith,  1937),  rabbits 
(Krebs  et  al.,  1938),- and  rats  (Busch  and  Potter,  1952  a)  leads  to  an  increased 
urinary  excretion  of  citrate  (Table  1-17  and  Fig.  1-13).  There  is  also  a  rise 
in  plasma  citrate  following  injections  of  malonate  in  rabbits  (Forssman, 
1941)  and  dogs  (Stoppani,  1946).  Tissue  citrate  also  rises  in  mice  injected 
with  10  millimoles/kg  malonate:  kidney  (16  to  20),  heart  (40  to  70),  liver 
(5  to  10),  and  diaphragm  (70  to  225)  (values  in  milligrams  per  kilogram 
wet  weight)  (Chari-Bitron,  1961).  Brain,  however,  shows  no  increase  in  cit- 
rate, perhaps  due  to  the  poor  penetration  of  malonate.  Some  accumula- 
tion of  citrate  has  also  been  observed  in  suspensions  of  Ashbya  gossypii 
mycelia  metabolizing  acetate  and  oxalacetate  (Mickelson  and  Schuler, 
1953),  Schizophyllum  commune  mycelia  metabolizing  pyruvate  and  malate 
(J.  G.  H.  Wessels,  1959),  and  Ehrlich  ascites  tumor  cells  metabolizing  fu- 
marate  (Kvamme,  1958  c)  in  the  presence  of  malonate.  Thus  this  phenom- 
enon is  widespread,  occurring  in  different  types  of  organism  and  under  a 
variety  of  conditions. 


SUBSTRATES   OTHER  THAN   SUCCINATE  105 

On  the  other  hand,  4-10.5  millimoles/kg  of  malonate  injected  intra- 
venously into  rabbits  does  not  increase  plasma  citrate  appreciably  (Forss- 
man  and  Lindsten,  1946),  and  a  number  of  reports  have  indicated  a  depres- 
sion of  citrate  formation  by  malonate.  Eat  brain  and  liver  homogenates  oxi- 
dizing oxalacetate,  or  pyruvate  and  oxalacetate,  form  less  citrate  in  the 
presence  of  10  mM  and  30  milf  malonate,  respectively,  this  being  attributed 
to  an  inhibition  of  oxalacetate  decarboxylase  (Pardee  and  Potter,  1949). 
The  formation  of  citrate  from  acetate  in  yeast  is  inhibited  73%  by  17  mill 
malonate,  while  simultaneously  succinate  accumulates  markedly  (Barron 
and  Ghiretti.  1953).  The  incorporation  of  C^^  from  glucose  into  citrate  in 
potato  tuber  slices  is  also  depressed  71%  by  50  mM  malonate  (Table 
1-19)  (Romberger  and  Norton,  1961).  Although  there  is  an  increase  in 
citrate  in  excised  tobacco  leaves  during  culture  with  malonate,  this  increase 
is  generally  less  than  in  the  controls,  so  that  this  probably  represents  an 
inhibition  of  citrate  formation  (Table  1-20)  (Vickery,  1959;  Vickery  and 
Palmer,  1957).  Finally,  malonate  inhibits  the  formation  of  citrate  from  a 
variety  of  substrates  in  kidney  and  testis  breis,  the  effects  being  surprisingly 
large  for  the  reasonable  concentrations  of  malonate  used  (Table  1-21) 
(Hallman,  1940).  It  is,  therefore,  evident  that  citrate  levels  may  be  affected 
by  malonate  in  a  variety  of  ways,  depending  on  the  malonate  concentration, 
the  type  of  preparation,  and  the  conditions  of  the  experiment.  It  is  not  dif- 
ficult to  explain  the  falls  in  citrate  level  brought  about  by  malonate,  since 
this  could  arise  either  by  a  depression  of  succinate  oxidation  (reducing  the 
rate  of  entry  of  acetyl-CoA  into  the  cycle)  or  inhibitions  of  other  reactions 
(such  as  the  condensation  of  oxalacetate  and  acetyl-CoA),  especially  at  the 
high  malonate  concentrations  often  used.  It  is,  on  the  other  hand,  difficult 
to  interpret  the  accumulation  of  citrate  and  to  this  end  we  must  direct 
our  efforts,  although  only  suggestions  can  be  offered  because  of  the  lack 
of  sufficient  data. 

Citrate  is  being  formed  and  metabolized  continuously  and  thus,  generally 
speaking,  a  rise  in  the  citrate  level  implies  an  inhibition  of  citrate  utili- 
zation or  an  acceleration  of  its  formation,  or  both.  Although  there  is  little 
evidence  for  a  direct  affect  of  malonate  on  isocitrate  utilization,  we  have 
noted  that  an  inhibition  of  succinate  oxidation  can  interfere  with  isocitrate 
oxidation  by  depletion  of  NADP  mediated  by  a  fall  in  malate  concentration 
(Jones  and  Gutfreund,  1964).  This  could  certainly  contribute  to  the  accu- 
mulation of  the  tricarboxylates  in  some  instances. 

The  inhibitions  of  citrate  oxidation  in  Table  1-14  can  be  mostly  explained 
on  the  basis  of  a  block  at  the  succinate  oxidase  step.  On  the  other  hand, 
there  is  certainly  no  evidence  that  the  formation  of  citrate  via  the  cycle 
can  be  stimulated  by  malonate,  most  data  pointing  instead  to  a  depression 
if  there  is  any  effect.  It  would  appear  that  effects  of  malonate  on  the  cycle 
alone  are  not  sufficient  to  explain  an  accumulation  of  citrate.  Other  path- 


106 


1.    MALONATE 


Table  1-19 

Distribution  of  Radioactivity  Following  3-Hr  Incubation  of  Potato  Tuber 
Slices  with  Glucose-u-C^*  "■ 


Component 

Radioactivity  (cpm) 

Control 

Malonate 

%  Change 

Sucrose 

344,000 

306,000 

-   11 

Glucose 

31,900 

26,800 

-   16 

Fructose 

5,950 

5,830 

-     2 

Oligosaccharides 

8,050 

6,000 

-  25 

Weak  acids 

Citrate 

4,350 

1,280 

-   71 

Isocitrate 

235 

757 

+222 

Succinate 

925 

6,300 

+582 

Fumarate 

190 

125 

-  34 

Malate 

8,900 

1,740 

-  80 

Glycolate 

2,750 

1,830 

-  33 

Total 

67,300 

41,500 

-  38 

Acidic  amino  acids 

56,100 

16,700 

-  71 

Neutral  amino  acids 

32,100 

56,200 

+   75 

Basic  amino  acids 

950 

750 

-  21 

Phosphorylated  compounds 

28,000 

10,000 

-   64 

Lipids 

111 

367 

+230 

Respiratory  COj 

22,600 

17,300 

-  23 

"  Fresh  slices  of  potato  tubers  incubated  at  28°  and  pH  5  for  3  hrs  with  uniformly 
labeled  glucose,  in  the  absence  of  and  presence  of  50  xnM  malonate.  (From  Romberger 
and  Norton,   1961.) 


ways  for  the  formation  of  citrate  are  known,  e.g.,  the  citrase  reaction  from 
acetate  and  oxalacetate,  or  through  isocitrate  by  the  isocitrase  reaction 
from  succinate  and  glyoxylate.  However,  the  free  energy  changes  for  these 
reactions  are  such  that  citrate  would  not  accumulate  in  significant  amounts 
even  though  the  substrates  for  its  formation  accumulated.  Also  the  results  in 
Table  1-22  show  that  the  effect  is  not  specific  for  malonate,  but  is  seen  with 
succinate,  malate,  fumarate,  and  other  organic  anions.  The  marked  effect 
of  glutarate,  which  is  a  poor  inhibitor  of  succinate  oxidase,  suggests  that 
the  citrate  accumulation  may  not  be  related  to  the  block  of  this  enzyme. 
It  may  be  noted  that  part  of  the  augmented  citrate  excretion  may  be  attri- 


SUBSTRATES   OTHER  THAN   SUCCINATE 


107 


S" 

GO 

o 

CO 

o 

w 

CO    -^ 

o 

o  o 

TlH      (M 

OO   CO     1 

S^l 

CO 

--^ 

5  ^ 

Ci    o 

<N 

i>   00     1 

^H 

a 

C<1 

^H 

^H 

ID 

1      + 

+  7 

+      1 

+     + 

1 

t-i 
3 

lO 

■* 

•  o 

(M 

t^ 

-*^ 

W^ 

W 

o  00 

CO 

-^  o 

I> 

LO    K)       1 

d 

00 

0 

'3 

I— 1 

o 

lO    CI 

Tf*      OO         1 

(M 

o 

a 

(N 

I-H 

I-H 

^H 

a 

1    + 

+  1 

+ 

+     + 

_\_ 

CO 

o 

^~, 

lO 

Ti 

IC    ■*' 

1— ( 

Oi    o 

CO    o 

<N    ^       1 

1 

1 

'^^ 

00    o 

-* 

CO               1 

(M 

I 

1 

bJD 
C 

Ksi    O     ft 

+  + 

-M 

+  + 

+ 

+    + 

_J_ 

(-1 

(M 

C5 

»o 

»o 

o 

"^ 

(C 

-*-:> 

00  d 

1 

lO    ^ 

CO    CO 

o    O      1 

(M 

1 

1 

k5* 

I 

CO 

IM 

<s 

1 

1 

^ 

^ 

+  + 

1    1 

+    + 

^  + 

_^ 

"o     ff>      • 

CO 

^ 

Ci 

CO 

S    t-  -Q 

a; 

-t^  ,o    s 

o  c- 

1 

-H       CO 

lO    fM 

o    O      1 

0<l 

1 

1 

"  ^    " 

1 

O    CO 

S 

1 

I 

5-S-S 

(M      -H 

,3 

"^^ 

t^ 

Tt< 

^-1 

CO 

CO 

w 

O   ci 

CO 

Tt<  cc 

^    O 

Oi    CO    0 

^ 

'X 

-^, 

CO 

CO     C5 

Tin    rt 

Tt<      0 

5 

ft 

i-H 

I-H 

I-H 

^ 

+ 

+  1 

+      1 

+  +  1 

1 

CO 

(^, 

s 

S" 

lO 

'M 

LO 

CO 

<3) 

w 

C<J     C2 

•* 

O       -H 

O      -H 

LO     C5     0 

,_) 

r- 

0 

o 

CO 

T)<      O 

CO      I-H 

lO    CO 

00 

_c 

1— 1 

3 

I-H 

1-H       I-H 

»-H 

1  + 

+  1 

+  1 

+  +  1 

1 

0) 

-2 

00 

0 

3 

1 

o 

CO 

OO 

CO 

CO 

+s 

W 

(M    00 

,—, 

-H      ^ 

CO       -H 

'>^   ^   <Z> 

^H 

0 

0 

"3 

(N 

(M 

t-   — < 

LO     OO 

Tti 

o 

Is 

ft 

C<1 

I-H      I-H 

s 

1      + 

+  1 

+  1 

+  +    1 

J_ 

_H 

t^ 

s 

^ 

O 

CI 

CO 

(M 

Tj 

w 

O      Tt< 

C5 

r-  c<) 

oi  C5 

t-  0  d 

I-H 

0 

0 

-* 

a>  -H 

CO  c^ 

00 

« 

ft 

I-H 

i-H 

f— ( 

1        + 

+        1 

+    1 

+  +    1 

1 

8  ^  e 

?M 

CO 

0 

CO 

CO 

^  o  W 
bi!  o    P. 

Tj<    IC 

o 

CO    Ci 

CO    CO 

10     1    0 

F-H 

1 

1 

-M     >0 

c^ 

(M       1 

1 

1 

+    + 

^ 

+    +    +    +    +              1 

1 

t- 

05 

t^ 

i-H 

<M 

CO 

Oj 

<x, 

-^ 

Ci    o 

1 

lo   CO 

O'    tJH 

0     1    0 

I-H 

1 

I 

1 

^   t^ 

>o 

2  1 

1 

1 

+  1 

1   1 

+  + 

^     1 

1 

10 

Oi 

"o    a>    -^ 

'* 

»o 

CO 

u    u   Xi 

<B 

-►^    ,0     3 

o  o 

1 

CO     GO 

JO    -* 

0       1      CO 

I-H 

1 

1 

G    ^-'     o 

(N 

1 

r^    lO 

CO  ->* 

cS       i 

1 

1 

a-s  -s 

(M    rt 

CO 

^ 

N 

, 

J3 

'^ 

^— V 

Ef     tJD 

C 

CO 

eg 
Ci 

_8i 

5  s 

c 

^ ^ 

~a^ 

C      (D 

^      2      M 

'm 

QJ         -H 

G 

o 

a 

a 

,G 

« 
C 

i5   ^  =« 

:G     M_g 

"a- 

s 

O 

13 

'5c  ^ 

^     -1 

-2  J? 

g    tl     G 

G     3 
^    in 
_g     0 

ce  ~- 0 

G    o~ 

-9     G 

o3 

a 

>    M    ft 

s 

o 

a  Is 

s  ■^  2 

s 

"tS  7 

"3    S 

"  <1  P 

H 

-S 

b  o 

03    S    Ph 

M 

UH        <!' 

k?H         C^ 

1-4  O 


P^ 


£    CI 


108 


1.    MALONATE 


H 

:z 

H 

P5 

H 

P4 

p:^ 

Oi 

Q 

-^ 

s 

1^ 

o 

IJI 

P3 

m 

h 

w 


O 


^ 

O 

a 

"S 

o 

S 

(N 

>. 

Xi 

a 

_o 

'-ij 

^ 

^ 

2 

s 

c 

1— 1 

o 

CO         -< 


O 


.a 

'-3 


-H         i>         00         ^ 

CO  ^  -H  ^ 


1 

> 

o 

CI 

o 

3 

» 

'o 

-i: 

w 

o 
3 

5 

^ 

flH 

c 

M 

E^ 

t3 


SUBSTRATES  OTHER  THAN  SUCCINATE 


109 


Table  1-22 

Urinary  Excretion  of  Citrate  Following  Administration  of  the  Sodium  Salts 

OF  Various  Weak  Acids 


Substance 


Urinary  citrate  "  after  dose ''  of : 


4.35 


26 


26 


39 


None 

NaCl 

NaHCOg 

Malonate 

Succinate 

Pyruvate 

a-Ketoglutarate 

Fumarate 

Malate 

Oxalacetate 

Glutarate 

Adipate 

Maleate 

Acetate 

Citrate 

Aconitate 


0.47 

16 

0.50 

— 

0.90 

44 

14.76 

583 

14.80 

418 



336 

17.10 

275 

18.50 

287 

— 

216 

— 

1150 

— 

200 

61.30 

210 

1.53 

— 

15.2 

— 

413 


2.4 
9-25 
54.2 
29.5 
10.2 
12.3 
25.8 
13.1 

6.7 
44.3 

3.7 
11.4 


26.3 


10-17 

72.4 

39.4 

42.3 

36.6 

20.1 

45.7 

16.2 

65.8 

17.6 

45.0 

34.5 

52.5 

42.0 


"  The  second  column  shows  the  24-hr  urinary  citrate  (mg/kg)  in  dogs.  From  Orten 
and  Smith,   (1937.) 

The  third  column  shows  the  urinary  concentration  of  citrate  (mg%)  in  rats. 
(From  Simola  and  Kosunen,  1938.) 

The  fourth  and  fifth  columns  show  the  24-hr  urinary  citrate  (mg)  in  rats.  (From 
Krusius,  1940.) 

*  Dose    in    miUimoles    per   kilogram. 


buted  to  alkalosis.  Crawford  (1963)  confirmed  that  the  injection  of  10  mil- 
limoles/kg  of  malonate  in  rats  causes  a  marked  rise  in  urinary  citrate 
(1.7  ->  47  /^moles/kg/hr),  but  found  that  succinate,  malate,  and  sodium 
bicarbonate  also  have  this  effect.  Serum  citrate  simultaneously  rises  very 
moderately.  It  was  concluded  that  all  the  effects  are  due  to  an  alkalosis 
induced  by  the  administration  of  sodium.  However,  this  cannot  account  for 
all  of  the  actions  of  malonate,  nor  could  it  be  responsible  for  the  citrate  ac- 
cumulation in  cell  suspensions  and  isolated  preparations. 

Another  possibility  is  that  the  citrate  arises  from  the  substances  that 
are  administered.  This  was  favored  by  Orten  and  Smith  (1937),  but  it  was 


110  1.    MALONATE 

difficult  then  with  incomplete  knowledge  of  the  cycle  and  related  pathways 
to  understand  how  such  conversions  could  take  place;  it  still  is.  That  is, 
if  there  is  no  significant  impairment  of  the  utilization  of  citrate  by  mal- 
onate,  it  is  difficult  to  conceive  of  a  pathway  for  the  formation  of  citrate  from 
malonate  that  would  be  so  rapid  as  to  lead  to  a  large  rise  in  the  citrate  level. 
A  possibility  that  seems  not  to  have  been  considered  is  that  the  substance 
determined  as  citrate  may  not  have  been  citrate  but  a  related  tricarboxylic 
acid  or  some  other  compound  giving  a  positive  test.  Although  Hallman 
(1940)  examined  the  specificity  of  the  determination,  there  are  many 
substances  that  have  not  been  tested.  For  example,  it  is  easy  to  formulate 
reactions  in  which  malonyl-CoA  could  react  with  various  carbonyl  sub- 
stances, such  as  glyoxylate  or  pyruvate,  to  form  tricarboxylate  anions  which 
might  be  oxidized  to  pentabromacetone  in  the  citrate  test  and  be  mistaken 
for  citrate.  Certain  dicarboxylates.  such  as  itaconate,  also  are  determined  in 
this  test.  Such  substances  may  not  be  readily  metabolized  and  hence  would 
accumulate  much  more  readily  than  citrate.  Although  this  posibility  may 
seem  far-fetched,  it  would  be  well  to  make  certain  that  it  actually  is  citrate 
that  is  accumulating  during  the  action  of  malonate.  It  would  be  necessary 
to  convert  only  a  small  fraction  of  the  administered  malonate  to  such  a 
compound,  since  a  dose  of  26  millimoles/kg  (see  column  3  of  Table  1-22) 
would  theoretically  give  rise  to  almost  1  g  of  a  substance  with  a  molecular 
weight  near  that  of  citrate,  whereas  actually  only  around  one-twentieth 
of  this  was  determined  as  citrate.  Of  course,  such  estimations  depend  on  the 
degree  of  sensitivity  of  the  test  to  the  compound.  If  there  is  any  validity  in 
this  suggestion,  it  may  be  that  depressions  of  citrate  levels  may  occur  in 
those  preparations  or  tissues  where  such  reactions  of  malonate  or  its  meta- 
bolic products  do  not  occur,  i.e.,  where  the  response  to  malonate  is  the  one 
expected  on  the  basis  of  its  inhibition  of  the  functioning  of  the  cycle. 

Accumulation  of  a-Ketoglutarate 

Very  large  increases  in  urinary  a-ketoglutarate  following  the  administra- 
tion of  malonate  to  rabbits  and  rats  were  reported  by  Krebs  et  al.  (1938) 
(Table  1-17)  and  this  has  been  confirmed  by  El  Hawary  (1955),  who  found 
a  3.7-fold  increase  in  serum  a-ketogluterate  30  min  after  the  intraperitoneal 
injection  of  20  millimoles/kg  malonate.  As  in  the  case  of  citrate,  Krusius 
(1940)  found  that  a-ketoglutarate  excretion  is  increased  not  only  by  mal- 
onate but  by  many  organic  anions:  malonate  (46.4),  maleate  (42.0),  malate 
(17.7),  succinate  (17.7),  /5-hydroxybutyrate  (15.7),  acetate  (14.3),  pyruvate 
(12.1),  fumarate  (9.3),  and  sodium  bicarbonate  (0.6-7.8)  (the  figures  give 
urinary  excretion  in  milligrams/day).  He  concluded  that  essentially  all  the 
substances  that  increase  citrate  also  raise  the  a-ketoglutarate  excretion. 
However,  glutarate  is  a  notable  exception,  for  it  potently  augments  citrate 
formation  but  has  no  effect  on  a-ketoglutarate  excretion.  This  would  seem 


SUBSTRATES   OTHER  THAN  SUCCINATE  111 

to  disprove  the  theory  that  glutarate  is  active  because  it  undergoes  /?-oxi- 
dative  decarboxylation  to  malonate.  Little  has  been  done  with  isolated 
preparations,  but  in  three  cases  accumulation  of  a-ketogkitarate  has  been 
observed  in  the  presence  of  malonate:  in  locust  sarcosomes  from  malate 
(Rees,  1954),  in  suspensions  of  ascites  cells  from  fumarate  (Kvamme, 
1958  c),  and  in  ascites  cells  and  rat  heart  mitochondria  from  glutamate  and 
malate  (Borst,  1962),  in  all  instances  the  malonate  concentration  being 
rather  high  (20-50  mM). 

The  mechanism  of  such  accumulation  is  poorly  understood.  We  have  seen 
that  in  some  tissues  the  a-ketoglutarate  oxidase  can  be  inhibited  by  malo- 
nate, especially  at  concentrations  above  10  mM,  so  that  the  results  can  be 
partially  explained  in  this  way  (at  least  for  the  ascites  cells  and  locust 
particulate  preparations).  The  moderate  increases  in  a-ketoglutarate  excre- 
tion brought  about  by  the  cycle  intermediates  and  substrates  might  well  be 
due  to  a  greater  rate  of  formation  with  unchanged  utilization  rate.  However, 
the  possibility  of  a  formation  of  c-ketoglutarate  from  glutamate  cannot  be 
ignored.  It  will  be  recalled  that  in  potato  slices  the  succinate  formed  in  the 
presence  of  malonate  comes  partly  from  such  a  source  (Romberger  and 
Norton,  1961).  Permeability  effects  causing  a  leakage  of  a-ketoglutarate 
and  other  anions  from  the  tissues  must  also  be  considered.  El  Hawary  (1955) 
found  that  several  inhibitors  (arsenite,  maleate,  iodoacetate,  alloxan,  and 
fluoride)  increase  the  serum  a-ketoglutarate  levels  in  rats,  and  simulta- 
neously raise  pyruvate  levels.  It  may  well  be  that  any  severe  metabolic 
disturbance  causes  a  release  of  cycle  substrates  from  the  tissues  and  an 
increased  excretion,  as  well  as  secondary  changes  such  as  hyperglycemia. 

Effects  on  the  Levels  of  Other  Cycle  Substrates 

The  tissue  concentrations  of  all  the  cycle  substrates  are  probably  altered 
by  malonate  and  it  is  sufficient  here  to  mention  the  results  with  potato  tuber 
slices  (Table  1-19)  and  tobacco  leaves  (Table  1-20).  The  reduction  in  the 
incorporation  of  C^**  from  glucose  into  malate  in  the  former  and  the  marked 
falls  in  malate  level  in  the  latter  would  be  anticipated  from  a  block  of  suc- 
cinate oxidation.  Fumarate  and  oxalacetate  levels  probably  are  changed 
similarly.  In  this  connection,  one  wishes  that  more  information  were  avail- 
able on  the  factors  that  control  the  tissue  pools  of  cycle  intermediates,  since 
it  is  evident  that  these  substances  do  not  occur  only  in  the  mitochondria 
in  kinetic  equilibria  depending  on  the  relative  rates  of  the  cycle  reactions, 
but  must  also  be  present  in  cellular  compartments.  The  transfer  of  the 
substances  between  these  compartments  and  the  active  cycle  regions  must 
depend  on  processes  that  could  be  affected  by  inhibitors.  Such  compart- 
ments are  well  known  in  plant  cells  but  it  is  probable  that  similar  situations 
are  applicable  to  animal  cells. 


112  1.    MALONATE 

Sequential  Inhibition  with  Malonate  and  Fluoroacetate  on  Citrate  Levels 

Before  leaving  the  subject  of  accumulation  of  cycle  intermediates,  a 
few  words  must  be  said  on  the  effects  of  malonate  on  the  increases  in  citrate 
brought  about  by  fluoroacetate.  This  was  discussed  in  Volume  I  (page  502) 
and  the  results  obtained  by  Potter  (1951)  presented  (Fig.  1-10-5).  The  prin- 
ciple of  the  experiments  is  simply  that  fluoroacetate  blocks  the  utilization 
of  citrate  so  that  the  tissue  citrate  levels  rise  at  different  rates  and  to  dif- 
ferent degrees.  If  malonate  is  administered  to  the  animals  prior  to  the 
fluoroacetate,  the  accumulation  of  citrate  may  be  modified.  These  studies 
thus  provide  information  on  the  effects  of  malonate  on  the  rates  of  forma- 
tion of  citrate  and  supplement  the  results  discussed  above. 

The  tissues  differ  greatly  in  their  response  to  malonate  in  the  presence  of 
fluoroacetate.  In  thymus,  for  example,  malonate  blocks  the  formation  of 
citrate  completely  and  no  citrate  accumulation  at  all  occurs.  The  kidney 
behaves  similarly  but  some  citrate  begins  to  accumulate  an  hour  after 
the  fluoroacetate  is  injected.  In  spleen  and  brain  the  inhibition  of  citrate 
formation  is  around  50%.  Heart  responds  quite  differently  from  the  other 
tissues.  Here  malonate  actually  increases  the  citrate  formation  somewhat. 
Potter  believed  these  results  to  indicate  that  in  heart  there  are  pathways 
other  than  the  cycle  for  the  synthesis  of  oxalacetate,  and  suggested  the  con- 
version of  ketone  bodies  (malonate  induces  ketonemia)  to  citrate  by  a 
pathway  involving  oxalacetate,  the  ketone  bodies  arising  from  fatty  acids. 
Liver  was  not  studied  but  if  similar  reactions  occur  in  this  tissue,  they 
could,. at  least  in  part,  account  for  the  increases  in  serum  citrate  and  urinary 
excretion  of  citrate.  In  any  event,  these  experiments  illustrate  very  well 
the  inherent  differences  between  tissues  with  respect  to  their  metabolic 
pathways.  The  stimulation  of  citrate  formation  in  the  heart  in  the  presence 
of  fluoroacetate  was  confirmed  by  Fawaz  and  Fawaz  (1954),  whereas  in  the 
kidney  only  a  depression  was  observed.  The  use  of  fluoroacetate  to  block 
the  utilization  of  citrate  is  a  useful  technique  by  which  to  study  the  effects 
of  malonate  uncomplicated  by  possible  effects  on  the  rate  of  disappearance 
of  citrate. 


ANTAGONISM    OF    MALONATE    INHIBITION 
WITH    FUMARATE 

The  overcoming  of  an  inhibition  by  the  addition  of  an  intermediate  nor- 
mally arising  distal  to  the  site  of  the  block  is  often  excellent  evidence 
for  the  locus  of  action  of  the  inhibitor  and  for  the  specificity  of  the  inhi- 
bition. For  this  reason,  fumarate  has  frequently  been  used  in  malonate- 
inhibited  preparations  and  a  reversal  of  the  inhibition  taken  as  proof  for 
the  specific  action  of  malonate  on  succinate  oxidase.  The  first  acceptable 


ANTAGONISM    WITH    FUMARATE  113 

data  for  fumarate  reversal  were  reported  by  Quastel  and  Wheatley  (1935), 
who  showed  that  the  inhibition  of  acetoacetate  utilization  by  rat  liver  in 
the  presence  of  malonate  is  mainly  abolished  when  fumarate  is  added, 
and  they  used  this  as  evidence  that  malonate  does  not  act  directly  on  the  en- 
zyme involved  in  the  breakdown  of  acetoacetate.  The  use  of  fumarate  im- 
mediately became  popular  and  many  studies  since  1935  have  included  its 
addition  for  the  purpose  of  demonstrating  that  the  observed  effect  of  mal- 
onate is  indeed  due  to  its  block  of  succinate  oxidation. 

Most  of  the  tests  for  fumarate  reversal,  it  must  be  admitted,  have  not 
been  done  properly  and  the  results  have  been  evaluated  uncritically,  so  that 
little  of  value  has  been  demonstrated.  There  are  several  points  that  must 
be  considered  in  the  planning  and  interpretation  of  such  experiments. 

(a)  An  increase  in  the  oxygen  uptake  upon  addition  of  fumarate  to  a 
malonate-inhibited  preparation  is  not,  by  itself,  very  meaningful.  Let  us  take 
a  typical  experiment  similar  to  many  reported  in  the  literature.  The  nor- 
mal Qq  of  a  tissue  preparation  is  10  and  this  is  decreased  to  4.5  by  mal- 
onate. When  malonate  and  fumarate  are  added  together,  the  Qq  is  8.7. 
It  has  been  generally  stated  in  such  cases  that  fumarate  is  capable  of 
antagonizing  the  malonate  inhibition.  Actually,  an  increase  in  the  respira- 
tion could  have  been  brought  about  by  the  addition  of  any  substrate  that 
is  oxidized  by  the  preparation,  including  substrates  completely  unrelated 
to  the  cycle.  One  can  say  that  fumarate  has  overcome  the  malonate  inhibi- 
tion but  the  results  are  of  no  particular  significance  with  respect  to  malonate 
specificity  or  the  site  of  the  inhibition.  What  one  has  shown  is  that  fumarate 
can  be  oxidized  in  the  presence  of  malonate  and  this  need  not  imply  that 
the  operation  of  the  cycle  has  been  even  partially  restored.  If  fumarate 
had  been  added  to  the  uninhibited  preparation  and  a  Qq  of  14.2  found,  the 
following  might  have  been  concluded:  malonate  inhibits  the  endogenous 
respiration  55%  and  the  respiration  in  the  presence  of  fumarate  either  13% 
(with  respect  to  the  endogenous  control)  or  39%  (with  respect  to  the  rate 
with  fumarate  alone  present),  in  both  cases  the  inhibition  being  less  than  that 
for  the  endogenous  respiration.  Actually,  it  has  been  shown  that  the  oxi- 
dation of  fumarate  is  unaffected  by  malonate,  since  the  same  absolute 
rise  in  the  Qq^  is  obtained  from  fumarate  in  the  uninhibited  and  inhibited 
preparations.  If  the  oxygen  uptake  resulting  from  the  addition  of  fumarate 
results  mainly  from  the  oxidation  of  malate,  the  results  would  have  little 
bearing  on  the  mechanism  or  selectivity  of  malonate  inhibition. 

(6)  The  addition  of  fumarate  would  not  in  any  case  restore  the  complete 
cycle,  since  oxidation  of  succinate  would  still  be  depressed  and  less  energy 
would  be  available  from  this  step.  It  is  possible  in  some  instances  that  the 
energy  from  succinate  oxidation,  rather  than  the  over-all  energy  production 
from  all  oxidations,  is  important,  and  this  would  not  be  restored  by  fu- 
marate. 


114  1.    MALONATE 

(c)  The  accumulation  of  succinate  due  to  malonate  inhibition  would,  of 
course,  not  be  reversed  by  fumarate;  instead,  it  is  usually  increased.  If  some 
response  to  malonate  is  dependent  on  the  rise  in  succinate  concentration 
(e.g.,  a  direct  effect  of  succinate  on  other  enzymes  or  some  cell  function,  or 
the  increased  formation  of  some  substances  derived  from  the  succinate), 
this  would  not  be  antagonized  by  fumarate. 

(d)  The  response  to  fumarate  will  often  depend  on  what  is  being  measured. 
One  example  will  be  used  here  to  illustrate  this  and  others  will  be  men- 
tioned later.  Malonate  inhibits  the  oxidation  of  trilaurin  and  octanoate 
in  liver  and  kidney  slices,  and  also  reduces  the  amount  of  C^^Og  formed  from 
labeled  substrates  (Geyer  et  al.,  1950  a).  It  was  found  that  fumarate  is  very 
ineffective  in  counteracting  the  inhibition  of  C^^O,  production,  and  this 
might  be  attributed  to  a  direct  effect  of  malonate  on  fatty  acid  oxidations. 
However,  the  results  can  be  explained  on  the  basis  of  an  inhibition  of  suc- 
cinate oxidase.  Much  of  the  C^*  taken  into  the  cycle  from  acetyl-CoA 
would  accumulate  in  succinate  and  fumarate  would  have  no  effect  on  this. 
For  the  full  release  of  CO2,  the  operation  of  the  entire  cycle  is  required,  and 
thus  fumarate  would  increase  the  C^^Og  formation  only  moderately  in  the 
presence  of  malonate.  If  the  oxygen  uptake  had  been  determined,  fumarate 
could  well  have  shown  a  complete  reversal  of  the  malonate  inhibition. 

(e)  Absence  of  a  reversal  by  fumarate,  or  the  failure  to  achieve  a  complete 
reversal,  can  be  due  to  a  variety  of  causes.  In  an  experiment,  such  as  the  one 
shown  in  the  following  tabulation,  done  on  Aplysia  muscle  slices: 


Additions  Qq^ 


Endogenous 

0.33 

Succinate 

1.80 

Malonate 

0.33 

Succinate  -f  malonate 

0.96 

Succinate  +  malonate  +  fumarate 

0.90 

although  it  was  stated  that  fumarate  was  unable  to  relieve  the  malonate 
inhibition  (Ghiretti  et  al.,  1959),  it  may  simply  be  that  fumarate  would  not 
have  been  oxidized  if  added  to  the  uninhibited  tissue.  Certainly  the  direct 
inhibition  of  succinate  oxidase  would  not  be  expected  to  be  overcome  by 
fumarate,  but  in  any  case  a  control  with  fumarate  alone  must  be  run  to 
give  any  significance  to  the  results.  It  is  even  possible  for  fumarate  to  in- 
crease the  inhibition  produced  by  malonate.  This  was  observed  for  the  respi- 
ration of  Helix  hepatopancreas  (see  accompanying  tabulation)  (Rees,  1953). 
Here  the  endogenous  respiration  apparently  is  little  dependent  on  the  cycle, 
whereas  upon  the  addition  of  fumarate,  through  the  formation  of  oxalace- 


ANTAGONISM    WITH    FUMARATE  115 

tate  and  its  partial  decarboxylation  to  pyruvate,  the  cycle  presumably 
becomes  activated.  The  high  inhibition  of  fumarate  oxidation  by  malonate 
cannot,  of  course,  be  attributed  solely  to  an  inhibition  of  succinate  oxidase, 

Additions  Og  uptake       %   Inhibition 


Endogenous 

30 

Malonate 

25 

Fumarate 

167 

Malonate  +  fumarate 

47 

16.7 
71.9 


and  an  effect  on  oxalacetate  decarboxylase  or  some  other  enzyme  at  the 
relatively  high  malonate  concentration  (33  mM)  must  be  assumed.  It  is 
desired  to  point  out  that  the  effects  with  fumarate  are  unrelated  to  the 
inhibition  of  the  endogenous  respiration  observed  with  malonate  and 
throw  no  light  on  the  mechanism  of  the  inhibition. 

Let  us  now  turn  to  some  experiments  in  which  the  addition  of  fumarate 
provides  results  indicative  of  the  mechanism  of  malonate  inhibition.  The 
most  significant  studies  usually  have  involved  the  determination  of  the 
utilization  or  formation  of  a  particular  substance.  In  the  original  work 
of  Quastel  and  Wheatley  (1935)  mentioned  above,  the  disappearance  of 
acetoacetate  in  rat  liver  slices  was  determined,  and  fumarate  was  found  to 
increase  its  utilization  in  the  presence  of  malonate,  whereas  in  the  absence 
of  malonate  it  had  essentially  no  effect.  Krebs  and  Eggleston  (1940)  likewise 
showed  that  fumarate  would  increase  the  utilization  of  pyruvate  in  pigeon 
muscle  in  the  presence  of  malonate,  at  high  malonate  concentrations  the 
amount  of  pyruvate  utilized  being  equivalent  to  the  fumarate  added.  The 
results  obtained  by  Stare  et  at.  (1941)  with  pigeon  muscle  show  definitely 
that  the  block  of  pyruvate  utilization  produced  by  malonate  is  effectively 
overcome  by  fumarate   (see  accompanying  tabulation).   Results  such  as 


Additions  PjTuvate  utilized 

Endogenous  16 . 3 

Malonate  7 . 2 

Fumarate  21.4 

Malonate  +  fumarate  23.3 

these  demonstrate  that  fumarate  can  counteract  the  effect  of  malonate  on 
a  specific  process  and,  from  the  nature  of  the  experiments,  it  is  likely  that 
the  cycle  block  is  being  overcome  in  a  sense,  that  is,  that  oxalacetate  is 
being  made  available  for  condensation  with  acetyl-CoA.  They  also  provide 


116  1.    MALONATE 

some  information  on  the  site  and  specificity  of  the  malonate  inhibition. 
Quite  different  results  on  acetate  utilization  by  yeast  were  obtained  by 
Stoppani  et  at.  (1958  b),  who  found  that  fumarate  is  completely  unable  to 
reverse  the  inhibition  by  malonate  (see  accompanying  tabulation).  These 

Additions  Acetate  utilized 


Acetate  30 
Acetate  +  malonate  4.4 

Acetate  +  fumarate  27 
Acetate  +  malonate  +  fumarate  3.8 


data  would  nuply  that  the  inhibition  of  acetate  utilization  by  malonate  is 
not  mediated  through  a  block  of  succinate  oxidation  but  by  another  jnecha- 
nism.  It  is  difficult  to  explain  this  by  a  failure  of  fumarate  to  penetrate 
into  the  cells  since  malonate  seems  to  enter  readily.  It  was  suggested  that 
malonate  might  interfere  with  acetate  activation  by  depleting  the  system 
of  coenzyme  A,  due  to  the  formation  of  relatively  stable  malonyl-CoA. 
Whatever  the  mechanism,  these  results  point  to  an  action  of  malonate 
other  than  on  succinate  oxidase. 

An  interesting  illustration  of  the  useful  information  that  may  be  ob- 
tained from  the  use  of  fumarate  is  given  in  the  inhibition  of  urea  formation 
by  malonate.  The  formation  of  both  arginine  and  urea  from  citrulline  and 
glutamate  in  liver  homogenates  is  potently  inhibited  by  malonate  (Cohen 
and  Hayano,  1946;  Krebs  and  Eggleston,  1948).  Fumarate  is  able  to  counter- 
act this  block  completely.  These  results  were  difficult  to  understand  ini- 
tially, but  it  is  now  known  that  transamination  must  occur  between  gluta- 
mate and  oxalacetate  to  form  aspartate,  which  reacts  with  the  citrulline 
to  form  arginosuccinate,  from  which  arginine  and  urea  are  derived.  The 
effect  of  malonate  is  to  reduce  the  supply  of  oxalacetate  for  transamination 
and  it  is  clear  why  fumarate  will  abolish  this  inhibition  (Ratner,  1955). 
Krebs  and  Eggleston  (1948)  observed  the  formation  of  3-5  molecules  of 
urea  for  each  molecule  of  fumarate  added.  This  may  be  explained  by  the 
fact  that  the  formation  of  arginine  from  arginosuccinate  involves  the  release 
of  fumarate,  which  can  again  go  to  oxalacetate. 

The  demonstration  that  fumarate  will  counteract  the  inhibitory  action 
of  malonate  on  some  tissue  function  is  indicative  of  a  primary  block  of  the 
succinate  oxidase,  but  even  complete  reversal  does  not  prove  a  specific  ac- 
tion of  malonate.  One  example  would  be  the  inhibition  of  Br~  uptake  in 
barley  roots  by  malonate  (Machlis,  1944).  Malonate  (10  mM)  inhibits  the 
uptake  57%  but  if  fumarate  is  present  the  uptake  is  35%  above  the  control. 
On  the  surface  this  would  imply  an  effective  reversal  of  the  malonate  inhi- 


SPECIFICITY  OF  MALONATE  INHIBITION  IN  THE  CYCLE  117 

bition,  but  actually  fumarate  alone  increases  the  uptake  to  71%  above  the 
control.  Thus  malonate  inhibits  a  significant  amount  even  in  the  presence 
of  fumarate,  pointing  to  an  action  other  than  on  succinate  oxidation. 
Another  example  would  be  the  malonate  inhibition  of  cell  division  oiArhacia 
eggs  (Barnett,  1953).  Cleavage  is  inhibited  almost  completely  by  60  milf 
malonate  and  an  equimolar  concentration  of  fumarate  abolishes  this.  It  is 
likely  that  fumarate  overcomes  the  cycle  block  but  one  cannot  conclude 
that  the  action  of  malonate  is  specific  on  succinate  oxidase.  Malonate,  can 
also  inhibit  to  some  extent  steps  in  the  utilization  of  fumarate,  but  due  to 
the  high  concentration  of  fumarate  enough  oxalacetate  is  formed  to  allow 
cleavage.  Indeed,  succinate  atGOmJf  also  abolishes  the  inhibition,  indicating 
that  as  a  result  of  the  competitive  nature  of  the  inhibition  enough  succinate 
has  broken  through  the  block  to  restore  cleavage.  It  must  be  remembered 
that  a  complete  reversal  of  a  metabolic  block  is  not  always  necessary  for  a 
cell  function  to  proceed  normally. 

SPECIFICITY    OF    MALONATE    INHIBITION    IN    THE    CYCLE 

At  this  point  we  may  summarize  some  of  the  conclusions  with  respect  to 
the  specificity  of  action  of  malonate  on  the  succinate  dehydrogenase.  Possi- 
ble effects  of  malonate  outside  the  cycle  will  be  discussed  in  a  later  section. 
We  have  seen  that  malonate  can  inhibit  enzymes  other  than  succinate  dehy- 
drogenase rather  potently  (Table  1-12),  that  the  oxidations  of  certain  cycle 
substrates  are  suppressed  more  than  predicted  on  the  basis  of  a  selective 
action  on  succinate  oxidation  (Table  1-14).  that  the  accumulation  patterns 
of  cycle  intermediates  are  sometimes  distorted  by  malonate  in  ways  implying 
inhibition  at  more  than  one  site,  and  that  fumarate  is  seldom  able  to  reverse 
the  actions  of  malonate  completely.  Of  particular  significance  are  the  clear 
demonstrations  of  the  inhibition  of  reactions  related  to  the  entry  of  acetyl- 
CoA  into  the  cycle,  particularly  those  of  Pardee  and  Potter  (1949)  pointing 
to  an  inhibition  of  the  condensation  reaction,  those  of  Stoppani  et  al. 
(1958  b)  reporting  a  marked  inhibition  of  acetate  utilization  apparently 
unrelated  to  an  inhibition  of  succinate  oxidase,  and  our  own  results  on  rat 
heart  mitochondria  where  malonate  inhibits  p^Tuvate  oxidation  in  the 
presence  of  malate  and  with  the  a-ketoglutarate  oxidase  completely  blocked. 
Another  susceptible  site  is  a-ketoglutarate  oxidation,  especially  in  view 
of  the  clear  proof  by  Price  (1953)  that  specific  inhibition  can  not  even  be 
obtained  in  the  simple  system, 

a-Ketoglutarate      -^     succinate      ->■     fumarate 

There  is  thus  a  large  amount  of  evidence  that  malonate  can  at  certain 
concentrations  in  various  conditions  inhibit  other  reactions  than  succinate 
oxidation. 


118  1.    MALONATE 

It  would  be  convenient  if  one  could  specify  a  malonate  concentration, 
or  range  of  concentrations,  which  would  most  likely  be  specific,  but  there 
are  too  many  factors  involved  to  do  this  with  any  confidence.  Some  of  the 
factors  may  be  listed:  (a)  the  species,  tissue,  or  preparation  used,  (b)  the 
enzymes  or  metabolic  pathways  involved  in  what  is  measured,  (c)  the 
conditions  of  the  experiment,  e.g.,  the  pH  or  the  Mg++  concentration,  (d) 
the  degree  to  which  succinate  can  accumulate  and  antagonize  the  inhibi- 
tion, (e)  the  effective  concentration  of  malonate  within  cells,  and  (f)  the 
possibility  of  nonenzyme  effects  on  cell  membranes  or  other  structures. 
Specificity  for  an  inhibitor  is  not  a  constant  to  which  can  be  given  a  value 
for  all  cases,  but  a  characteristic  that  must  be  evaluated  for  each  experi- 
ment. Aside  from  the  direct  actions  of  malonate  on  enzymes,  there  are  the 
problems  of  metal  cation  depletion,  the  possible  effects  of  Na+  or  K+  added 
with  the  malonate,  and  the  inactivation  of  the  coenzyme  A  in  some  pre- 
parations through  the  formation  of  malonyl-CoA. 

Bearing  in  mind  these  difficulties,  a  few  general  remarks  may  be  made. 
It  is  probable  that  malonate  usually  does  not  inhibit  any  cycle  enzyme  more 
strongly  than  succinate  dehydrogenase,  so  that  the  major  effect  will  be  re- 
lated to  the  inhibition  at  this  site,  but  it  will  be  recalled  that  in  certain 
species  the  succinate  dehydrogenase  is  not  very  susceptible  to  malonate.  It 
is  impossible  to  achieve  a  nearly  complete  block  of  succinate  oxidation 
without  affecting  other  cycle  reactions;  if  one  wishes  a  specific  effect,  one 
must  be  satisfied  with  a  moderate  inhibition  of  succinate  oxidation.  If  a 
single  malonate  concentration  for  general  usefulness  had  to  be  chosen,  5  mM 
might  be  taken  provisionally.  Although  in  some  cases  rather  incomplete  inhi- 
bition will  be  obtained,  this  concentration  will  probably  not  inhibit  other 
cycle  reactions  significantly.  This  applies  to  noncellular  preparations  where 
penetration  is  not  a  factor  but  otherwise  higher  external  concentrations 
may  have  to  be  used.  In  any  case,  good  evidence  for  a  specific  action  should 
be  obtained  under  the  conditions  of  the  investigation,  and  reliance  should 
not  be  based  on  generalities. 

EFFECTS   OF   MALONATE 
ON   OXIDATIVE  PHOSPHORYLATION 

An  uncoupling  action  on  oxidative  phosphorylation  has  been  claimed  for 
malonate  several  times  and  it  is  important  to  determine  if  this  is  actually 
so.  This  is  quite  difficult  because  pertinent  or  reliable  data  are  generally 
lacking.  There  are  four  important  ways  in  which  malonate  could  alter  the 
P  :  0  ratio:  (1)  a  direct  effect  on  the  coupling  between  oxidation  and  phos- 
phorylation, (2)  an  alteration  of  the  pattern  of  substrate  oxidation,  since 
different  substrates  may  have  different  P  :  0  ratios,  (3)  a  differential  inhi- 
bition of  electron  transport  pathways  for  a  single  substrate  but  with  differ- 


EFFECTS    ON    OXIDATIVE    PHOSPHORYLATION  119 

ent  P  :  0  ratios,  and  (4)  an  effect  on  the  hydrolysis  of  ATP  or  other  high 
energy  substances.  Only  the  first  mechanism  should  be  considered  as  true 
uncoupling,  although  it  is  often  difficult  to  determine  the  exact  mechanism. 
Included  in  the  first  mechanism  would  be  the  chelation  of  malonate  with 
Mg++  or  Mn++,  since  these  cations  are  beheved  to  be  cofactors  in  phospho- 
rylation. 

Claims  for  an  uncoupling  action  will  be  discussed  first.  Lehninger  (1951) 
stated  that  malonate  uncouples  oxidative  phosphorylation  associated  with 
the  oxidation  of  /5-hydroxybutyrate  by  rat  liver  mitochondria  but  no  data 
were  given.  In  a  previous  report  (Lehninger,  1949)  7.5  mM  malonate  was 
shown  to  inhibit  oxygen  uptake  48%,  the  formation  of  acetoacetate  from 
/?-hydroxybutyrate  64.7%,  and  phosphorylation  37.3%  (as  determined  by 
the  incorporation  of  P''-  into  the  ester  fraction).  Since  phosphorylation  is 
inhibited  less  than  oxidation,  no  uncoupling  is  evident,  and  indeed  the  P:0 
ratio  should  increase.  Berger  and  Harman  (1954)  claimed  that  malonate 
inhibits  phosphorylation  associated  with  the  one-step  oxidation  of  a- 
ketoglutarate  and  completely  suppresses  phosphorylation  during  the  oxida- 
tion of  L-glutamate  by  muscle  mitochondria.  However,  the  malonate  con- 
centration is  not  given  and  the  absence  of  data  prevents  evaluation  of  the 
results.  An  inhibition  of  phosphorylation  does  not  necessarily  mean  an 
uncoupling  action.  Malonate  at  30  milf  drops  the  P:0  ratio  from  1.6  to  0.5 
in  the  oxidation  of  choline  by  rat  liver  mitochondria  (Rothschild  et  al., 
1954).  Although  no  control  with  malonate  alone  was  reported,  it  would 
appear  that  this  is  the  most  valid  instance  of  uncoupling  by  malonate. 
The  malonate  concentration  was  high  and  a  reduction  of  Mg++  (total  con- 
centration was  5.7  mM)  must  be  considered.  Finally,  the  phosphorylation 
associated  with  succinate  oxidation  in  lupine  mitochondria  was  shown  to 
be  strongly  depressed  by  10  mM  malonate  (Conn  and  Young,  1957),  but 
it  may  be  observed  that  the  oxygen  uptake  was  inhibited  even  more  (Ta- 
ble 1-23),  so  that  no  uncoupling  occurred. 

All  of  the  reports  in  which  P:0  ratios  were  calculated  are  summarized 
in  Table  1-23  and  in  all  cases,  except  for  E.  coli,  Carcmvs  maenas,  and  the 
oxidation  of  choline,  it  is  seen  that  the  P:0  ratio  is  actually  increased  by 
malonate.  However,  most  of  these  only  illustrate  mechanism  (2)  above, 
because  in  the  oxidation  of  a-ketoglutarate  a  rise  in  the  P:0  ratio  would 
be  expected  upon  blocking  succinate  oxidase  due  to  the  fact  that  P:0  for 
succinate  oxidation  is  2,  whereas  for  the  one-step  oxidation  of  a-ketoglu- 
tarate to  succinate  usually  it  is  experimentally  between  3  and  4.  All  one  can 
say  from  such  data  is  that  there  is  no  evidence  for  an  uncoupling  action  by 
malonate.  Copenhaver  and  Lardy  (1952)  used  3-20  mM  malonate  in  all 
their  media  in  the  study  of  the  phosphorylation  associated  with  a-keto- 
glutarate oxidation  and  obtained  high  P:0  ratios,  again  providing  evidence 
against  any  uncoupling  activity.  It  was  shown  by  Slater  and  Holton  in 


120 


1.    MALONATE 


05 


pq 


\4 


m 


M 


X 


CO   -^     t^  o     00  o 

do     d  -H     d  ^ 


lO    CO      C3    CO      o    >c 
—I    CO       — I    ■*      iM    (M 


•M     IC    C^      C2    CO 

^  rt  ^    d  -H 


^ 


+ 


Cm 


1^ 


W 


o 


^ 


^ 


H 


fe] 


s 

c 

c 

03 

t>n 

Xi 

S 

O 

C 

t<n 

P 

J 

s 

tin 

~ 

§^^ 

^ 


p* 

o 

<5 

p. 

M 

+3 

1 

03 

►-H 

EFFECTS    ON    OXIDATIVE    PHOSPHORYLATION  121 


'-~.  lO 


Oi 

ft 

I— ( 

ft 

— 

o3 

^^ 

H 

-t^ 

t3 

cS 

w. 

c3 

T3 

C 

fi 

c3 

cS 

Q 


t4 


o  -^  ao  o;^ft-c_go 


GO 
05 

CO 

'M 
iM 

CO 

C5 

^   00 

-H      CO 

00 

o 

CO 

cc 

00 

CO 

o  ^ 

o 

§ 

CO 

CO 

00 

CO 

CJ 

•M 

-H 

o 

o 

S^ 

CO 

CO 

fc  ec 

CO 

CO 

CO 

CO 

CO 

CO    C^l 

-H 

o 

rt     C^l 

o 

o 

o 

o 

I   I   I 


+  I        I   I   I        I        I        I        I        I        I       + 


a  a  as  csccOs 


?^ 

?i< 

^^ 

;>■. 

o 

rV. 

> 

^ 

cs 

5^ 

x> 

>i) 

o^o^^d       PM  pm        §        0^        Ph        p5        m 


122  1.    MALONATE 

rat  heart  (1954)  that  malonate  increases  the  P:0  ratio  with  a-ketoglutarate 
as  the  substrate  as  the  malonate  concentration  is  increased  up  to  20  mM; 
at  40  mM  the  P:0  ratio  decreases  somewhat,  so  that  at  this  high  concen- 
tration a  small  degree  of  uncoupling  may  occur.  Azzone  and  Carafoli  (1960) 
found  the  same  in  pigeon  muscle.  Evidence  against  uncoupling  by  malonate 
intracellularly  is  provided  by  the  study  on  ascites  carcinoma  cells  by 
Greaser  and  Scholefield  (1960).  Comparison  of  the  changes  brought  about 
by  malonate  (20  mM)  in  the  respiration  and  the  sum  of  the  concentrations 
of  ADP  and  ATP  led  to  the  estimation  of  a  15%  increase  in  the  P:0  ratio, 
whereas  the  classic  uncoupler,  2,4-dinitrophenol,  depressed  the  P:0  ratio 
70%.  Malonate  may  actually  stimulate  the  esterification  of  phosphate.  In 
addition  to  the  two  examples  in  Table  1-23,  Jackson  et  al.  (1962)  reported 
an  8%  elevation  in  phosphate  uptake  by  0.1  mM  malonate  in  barley  root 
mitochondria  oxidizing  succinate,  and  Rosa  and  Zalik  (1963)  found  such  a 
stimulation  in  pea  seedling  mitochondria,  which  is  maximal  around  0.01 
mM  malonate,  the  oxygen  uptake  not  being  altered.  Above  this  concentra- 
tion, both  phosphorylation  and  oxidation  are  depressed  in  a  parallel  fashion; 
respiration  is  hence  always  depressed  somewhat  more  than  is  phosphate 
esterification,  and  no  uncoupling  is  seen  at  any  malonate  concentration. 
The  transfer  of  phosphate  from  phosphorylated  coenzyme  Q  to  ADP  to 
form  ATP  in  mitochondrial  preparations  is  not  altered  by  even  10  milf 
malonate  (Gruber  et  al.,  1963),  whereas  dinitrophenol  inhibits  this  readily. 
It  would  therefore  appear  to  be  legitimate  to  conclude  that  malonate 
does  not  exhibit  uncoupling  activity  except  possibly  at  high  concentrations 
(above  30  mM).  The  evidence  for  uncoupling  at  concentrations  commonly 
used  is  considered  to  be  inadequate  and  outweighed  by  the  mass  of  indirect 
evidence  that  P:0  ratios  are  not  reduced  in  the  oxidations  of  citrate,  a- 
ketoglutarate,  and  succinate. 


EFFECTS    OF    MALONATE    ON    GLUCOSE    METABOLISM 

The  actions  of  malonate  on  carbohydrate,  lipid,  amino  acid,  porphyrin, 
and  other  types  of  metabolism  will  now  be  considered,  after  which  it  will  be 
possible  to  evaluate  the  specificity  of  malonate  more  broadly.  The  effects 
on  glucose  metabolism  are  important  but  difficult  to  analyze.  In  the  first 
place,  the  interrelationships  between  the  cycle  and  the  glycolytic  pathways 
are  complex  and  secondary  effects  on  glucose  utilization  must  be  expected. 
In  the  second  place,  much  of  the  work  on  the  alteration  of  glucose  utiliza- 
tion by  malonate  has  not  been  adequate  for  the  determination  of  mecha- 
nisms nor  do  the  results  even  provide  useful  information  in  many  cases,  and 
for  this  reason  only  certain  reports  will  be  discussed.  There  are  three  basic 
ways  by  which  malonate,  might  alter  glucose  oxidation.  (1)  Malonate  will 
usually  cause  a  depression  of  the  oxygen  uptake  related  to  glucose  metabo- 


EFFECTS  OF  MALONATE  ON  GLUCOSE  METABOLISM  123 

lism  because  in  the  total  oxidation  of  glucose  to  CO,  and  water,  five-sixths 
of  the  oxygen  is  taken  up  in  cycle  reactions.  This  effect  would  be  due  to  a 
block  of  the  cycle  and  would  be  roughly  equivalent  to  the  inhibition  of 
of  pyruvate  oxidation  via  the  cycle.  (2)  Malonate  may  cause  an  increased 
uptake  and  utilization  of  glucose  as  a  result  of  inhibition  of  oxidative  reac- 
tions in  the  cycle,  the  mechanism  being  essentially  that  of  the  Pasteur  reac- 
tion. (3)  Malonate  may  have  direct  actions  on  the  cellular  uptake  of  glucose 
or  on  the  glycolytic  pathways.  Evidence  for  each  of  these  mechanisms  will 
be  presented,  and  then  a  more  detailed  analysis  of  the  possible  shifts  in 
metabolic  patterns  brought  about  by  malonate  will  be  given. 

Effects  on  Oxygen  Uptake  Associated  with  Glucose  Oxidation 

There  are  many  reports  on  the  effects  of  malonate  on  respiration  with 
glucose  as  the  substrate,  but  in  few  have  the  data  provided  a  correction  for 
an  effect  on  the  endogenous  respiration.  Indeed,  an  endogenous  correction 
to  determine  the  oxygen  uptake  associated  only  with  glucose  oxidation 
is  particularly  unreliable,  even  when  the  data  are  available.  This  is  because 
of  the  well-known  effect  of  glucose  in  suppressing  endogenous  respiration  and 
mitochondrial  oxidations  (Crabtree  effect).  Thus  the  nonglucose  respiration 
may  be  significantly  changed  when  glucose  is  added  and,  perhaps,  completely 
suppressed  in  some  cases. 

Let  us  assume  that  the  oxygen  uptake  measured  derives  only  from  glucose 
and  that  malonate  acts  only  on  the  cycle.  What  inhibitions  could  one  theo- 
retically expect?  The  inhibition  will  depend,  other  than  on  the  degree  of 
cycle  block,  on  the  final  products  of  glucose  oxidation  before  and  after  the 
addition  of  malonate.  The  following  equations  give  the  molar  oxygen  up- 
takes for  the  oxidation  of  glucose  to  varying  degrees  of  completeness: 

To  CO2  and  water 

CeHi^Oe  +  6  O2  ^  6  CO,  +  6  H,0 

To  pyruvate 

CgHi^Oe  +  O2  ->  2  CH3COCOOH  +  2  H2O 

To  acetate 

CeHiaOe  +  2  0,  ->  2  CH3COOH  -f-  2  CO,  +  2  H^O 

To  succinate 

CfiHi^Oe  +  5/2  O2  ->  HOOCCH2CH2COOH  +  2  CO2  +  3  H2O 

CsHiaOg  +  2  HOOCCH2COCOOH  +  4  0,  ->  2  HOOCCH2CH2COOH  +  6  CO2  +  4  H^O 

To  lactate 

CsHi^Oe  +  2  H  -).  2  CH3CHOHCOOH 

The  second  equation  for  the  formation  of  succinate  might  express  the  situa- 
tion in  which  there  is  a  source  of  oxalacetate  other  than  from  pyruvate 


124  1.    MALONATE 

carboxylation,  whereas  the  first  equation  assumes  no  external  oxalacetate 
source.  If  the  inhibition  on  succinate  oxidase  is  complete  and  glucose  is 
oxidized  completely  in  the  control,  we  may  calculate  the  inhibitions  for  the 
situations  where  in  the  presence  of  malonate  the  glucose  is  transformed  into 
the  various  oxidation  products:  to  succinate  with  external  oxalacetate 
source  (33%),  to  succinate  in  a  closed  system  (58%),  to  acetate  (67%), 
to  pyruvate  (83%),  and  to  lactate  (100%).  Of  course,  experimentally  it  is 
probably  impossible  to  achieve  a  complete  cycle  block  with  malonate,  so 
that  the  inhibitions  will  usually  be  less  than  the  theoretical.  These  calcula- 
tions also  assume  that  the  rate  of  glucose  disappearance  is  not  altered  by 
the  inhibition.  The  degree  of  inhibition,  therefore,  even  under  these  simple 
conditions,  will  vary  a  good  deal,  depending  on  the  terminal  product  of 
glucose  metabolism  and  thus  on  the  enzymes  and  pathways  occurring  in 
the  cells  under  consideration.  Actually,  it  is  likely  that  various  substances 
will  accumulate  during  glucose  metabolism,  and  not  only  those  given  above 
but  others  into  which  these  substances  can  be  metabolized. 

A  strong  inhibition  of  the  oxygen  uptake  related  to  glucose  metabolism 
probably  indicates  the  participation  of  the  cycle  in  the  oxidation,  at  least 
when  the  malonate  concentration  is  not  unreasonably  high.  Thus,  in  the 
instances  shown  in  the  tabulation  and  selected  at  random  of  inhibition  of 


Preparation 

Malonate 
(mif) 

% 

Inhibitior 

1                 Reference 

Sarcina  lutea 

10 

50 

Dawes  and  Holms  (1958) 

Yeast 

12.5 

73 

Krebs  et  al.  (1952) 

Bull  sperm 

10 

57 

Lardy  and  Phillips  (1943  a) 

Malarial  parasitized 

RBC 

20 

69 

Speck  et  al.  (1946) 

Brain  mince 

17 

100 

Huszak  (1940) 

/iL+-Stimulated  brain  slices 

10 

62 

Kimura  and  Niwa  (1953) 

glucose  respiration,  one  would  conclude  that  the  role  of  the  cycle  is  impor- 
tant, but  no  information  on  the  exact  mechanism  or  on  the  fate  of  glucose 
in  the  presence  of  malonate  may  be  obtained.  However,  in  the  cases  of  brain 
mince  and  yeast,  fumarate  is  unable  to  overcome  the  inhibition,  so  that 
some  doubt  is  introduced  even  into  this  interpretation.  The  inhibition  of 
the  glucose  oxygen  uptake  in  parasitized  erythrocytes  is  very  similar  to 
that  for  the  oxygen  uptake  associated  with  pyruvate  oxidation,  as  expected 
if  the  cycle  is  the  terminal  pathway  for  glucose  metabolism.  On  the  other 
hand,  there  have  been  many  reports  that  malonate,  even  at  high  concentra- 
tions, does  not  inhibit  the  glucose  respiration  at  all,  and  in  these  instances 
little  can  be  concluded  because  of  the  possibility  that  malonate  does  not 
penetrate  and  that  an  adaptation  of  the  glucose  utilization  occurs.  The 


EFFECTS  OF  MALONATE  ON  GLUCOSE  METABOLISM  125 

endogenous  respiration  of  Ehrlich  ascites  cells  is  inhibited  around  35%  by 
30  vaM  malonate,  but  when  glucose  is  present  there  is  very  little  effect  of 
malonate  (Seelich  and  Letnansky,  1960).  Put  in  another  way,  in  the  pres- 
ence of  malonate  the  addition  of  glucose  increases  the  0,  uptake  somewhat 
instead  of  depressing  it  as  it  does  in  uninhibited  cells.  Another  possible 
reason  for  a  failure  of  malonate  to  depress  glucose  respiration  is  the  oc- 
currence of  oxidative  pathways  unassociated  with  the  cycle.  Caldariomyces 
fumago  has  no  hexokinase  and  the  usual  Embden-Meyerhof  glycolytic 
pathw^ay  is  absent;  the  oxidation  of  glucose  occurs  more  directly  through 
glucose  and  gluconate  oxidases  (Ramachandran  and  Gottlieb,  1963).  The 
respiration  in  this  organism  is  not  affected  by  malonate  at  10  mM. 

Effects   on    Glucose    Utilization 

A  depression  of  cycle  oxidations  by  malonate  would  be  expected  to  cause 
an  increased  utilization  of  glucose,  as  do  hypoxic  conditions,  in  cells  capable 
of  exhibiting  a  Pasteur  reaction.  This  is  a  manifestation  of  one  interrela- 
tionship between  the  cycle  and  the  glycolytic  pathway,  and  is  partially 
mediated  through  changes  in  the  concentrations  of  inorganic  phosphate 
and  phosphate  acceptors,  such  as  ADP.  For  such  a  response  to  occur,  the 
utilization  of  glucose  must  have  been  initially  limited  by  the  coupled 
phosphorylation  at  the  3-phosphoglyceraldehyde  oxidation  step.  Another 
mechanism  for  the  influence  of  cycle  activity  on  glucose  utilization  involves 
the  membrane  transport  of  glucose  and  its  phosphorylation.  The  inward 
transport  of  glucose  under  certain  conditions  may  limit  glucose  utilization, 
and  it  has  been  shown  that  anoxia  accelerates  this  transport  in  rat  heart 
(Morgan  et  al.,  1961  a,  b).  It  is  quite  possible  that  malonate  by  its  depres- 
sion of  cycle  activity,  can  alter  these  transport  processes.  A  third  mechanism 
might  involve  the  rate  of  oxidation  of  NADH.  The  addition  of  pyruvate  to 
ascites  cells  stimulates  the  formation  of  C^^02  from  labeled  glucose  (Wenner 
and  Paigen,  1961)  Initially  the  rate  of  pyruvate  oxidation  is  limited  by  the 
rate  of  NADH  oxidation  and  the  exogenous  pyruvate  acts  as  an  electron 
acceptor,  1  mole  of  lactate  appearing  for  each  mole  of  pyruvate  oxidized 
through  the  cycle.  The  accumulation  of  pyruvate  as  a  result  of  the  cycle 
block  by  malonate  could  initiate  this  dismutation  so  that  more  glucose 
would  be  utilized  than  otherwise.  Indeed,  lactate  formation  is  often  increased 
during  malonate  inhibition.  In  all  these  ways,  and  perhaps  others,  a  cycle 
block  might  affect  glucose  uptake  and  utilization. 

A  stimulation  of  glycolysis  by  malonate  was  first  observed  by  Kutscher 
in  Heidelberg  (Kutscher  and  Sarreither,  1940;  Kutscher  and  Hasenfuss, 
1940).  Malonate  was  injected  into  guinea  pigs,  the  muscle  removed  later, 
and  the  formation  of  lactate  determined  in  a  brei.  In  some  cases,  glucose, 
succinate,  or  fumarate  was  also  injected.  Malonate  accelerates  lactate  for- 
mation and  this  is  overcome  by  both  succinate  (see  accompanying  tabula- 


126  1.    MALONATE 

tion)  and  fumarate.  It  was  also  shown  that  the  rate  of  glycogen  disappear- 
ance in  muscle  brei  is  stimulated  some  35%  by  malonate.  These  responses 
were  equated  with  the  Pasteur  reaction.  Similar  effects  have  been  found  in 


Injection 

Lactate  formation 

Control 

2.95 

Malonate 

9.2 

Glucose 

5.1 

Glucose  +  malonate 

9.4 

Succinate  +  malonate 

-1.0 

Glucose  +  succinate  +  malonate 

-0.1 

other  tissues  more  recently.  The  rates  of  glucose  utilization  and  of  lactate 
formation  in  brain  slices  are  stimulated  13%  and  44%,  respectively,  by 
10  mM  malonate  (Takagaki  et  at.,  1958),  an  effect  much  like  that  produced 
by  azide.  Simultaneously  less  glucose  is  oxidized,  the  increased  utilization 
being  diverted  to  lactate.  Ehrlich  ascites  tumor  cells  exhibit  a  more  rapid 
rate  of  glycolysis  when  the  oxygen  tension  is  reduced  and  a  similar  response 
is  seen  with  malonate  (Kvamme,  1957,  1958  d).  When  fumarate  is  added 
to  the  malonate-blocked  cells,  glucose  utilization  and  lactate  formation  are 
suppressed.  Strictly  speaking,  it  is  fumarate  that  gives  rise  to  a  Pasteur 
reaction  in  the  presence  of  malonate,  just  as  the  addition  of  oxygen  does 
in  preparations  previously  anaerobic.  Kvamme  obtained  data  which  led  him 
to  conclude  that  this  effect  is  mediated  through  changes  in  the  concentra- 
tions of  inorganic  phosphate  and  phosphate  esters.  The  relationships 
between  glycolysis  and  mitochondrial  oxidations  are  particularly  well  seen 
in  the  reconstructed  systems  of  Aisenberg  et  al.  (1957).  A  supernatant  frac- 
tion from  rat  liver  forms  lactate  from  glucose  and  the  addition  of  a  mito- 
chondrial suspension  suppresses  this  markedly.  Malonate  partially  prevents 
this  suppression;  that  is,  added  to  the  complete  system,  lactate  formation  is 
increased.  This  stimulation  of  glycolysis  occurs  despite  the  fact  that  mal- 
onate inhibits  the  glycolytic  rate  15.7%  in  the  supernatant  fraction.  This  is 
a  good  illustration  of  how  malonate  can  produce  different  effects  on  glucose 
metabolism,  depending  on  the  conditions  and  the  factors  controlling  gly- 
colysis. Stimulation  of  aerobic  glycolysis  by  malonate,  and  a  variety  of 
other  inhibitors,  is  particularly  well  seen  in  thymocytes;  malonate  ac- 
celerates lactate  formation  sigmoidally  from  3  mM  to  100  mM  (Araki 
and  Myers,  1963). 

Malonate  may  have  no  effect  on  glucose  uptake,  or  may  inhibit  it,  in 
other  tissues.  Chick  chorioallantoic  membrane  infected  with  influenza 
virus  exhibits  a  47.5%  reduction  in  the  endogenous  respiration  in  the  pre- 
sence of  6  mM  malonate,  and  the  final  virus  titer  drops  to  zero,  but  the 


EFFECTS  OF  MALONATE  ON  GLUCOSE  METABOLISM  127 

uptake  of  glucose  is  unaffected  (Ackermann,  1951).  An  example  of  a  marked 
inhibition  was  reported  for  guinea  pig  brain  slices,  glucose  utilization  being 
reduced  by  10  mM  malonate  in  both  normal  and  K+-stimulated  slices  (see 
accompanying  tabulation)  (Tsukada  and  Takagaki,  1955).   It  may  be  ob- 


KCl 

(70  mM) 

Malonate 

(10  mM) 

Glucose 
utilization 
(//mole/g) 

Lactate 
formation 
(//mole/g) 

O2  uptake 
(//mole/g) 

35.9 

34.9 

54.1 

— 

+ 

4.71 

28.0 

61.5 

+ 

— 

112 

99.0 

104 

+ 

+ 

63.7 

117 

44.9 

served  that  the  amount  of  lactate  formed  per  glucose  consumed  is  increased 
by  malonate,  but  some  of  the  lactate  must  be  derived  from  other  sources 
than  glucose.  It  is  difficult  to  understand  the  discrepancy  between  these 
results  and  those  obtained  later  on  the  same  tissue  (Takagaki  et  al.,  1958), 
where  a  slight  stimulation  of  glucose  utilization  by  10  mM  malonate  was 
reported.  There  are  so  many  different  results  obtained  relative  to  the  utili- 
zation of  glucose  that  it  often  appears  each  organism  or  tissue  exhibits  a 
characteristic  pattern  of  response.  Malonate  at  40  mM  has  relatively  little 
effect  on  the  uptake  of  glucose  by  brain  and  kidney  slices,  depressing  it 
slightly  in  the  former  and  perhaps  accelerating  it  in  the  latter,  but  reduces 
the  C^^Og  formed  from  uniformly  labeled  glucose  79%  and  52%,  respective- 
ly, the  O2  uptake  being  suppressed  comparably  (Cremer,  1962).  Since  much 
less  of  the  glucose  goes  to  amino  acids  and  proteins  in  the  presence  of  mal- 
onate (Cremer,  1964),  it  is  likely  that  here  there  is  an  accumulation  of 
certain  cycle  intermediates,  such  as  succinate,  and  of  lactate.  Succinate  is 
normally  formed  and  released  into  the  medium  by  Trypanosoma  cruzi. 
It  is  formed  aerobically  through  both  the  glycolytic  pathway  and  the  cycle, 
and  by  CO2  fixation;  some  must  be  metabolized  through  the  cycle  since 
malonate  elevates  the  succinate  level  even  further  (Bowman  et  al.,  1963). 
The  uptake  of  glucose  is  increased  almost  20%  by  malonate  but  there  is  no 
change  in  C^^Oo  formation,  most  of  the  excess  glucose  probably  appearing 
as  succinate,  acetate,  and  related  anions. 

Direct  Effects  on  the  Glycolytic  Pathways 

There  is  no  evidence  that  any  enzyme  of  the  Embden-Meyerhof  glycolytic 
pathway  is  significantly  inhibited  by  malonate  at  concentrations  below  20 
mM  (Table  1-12),  although  some  of  the  enzymes  have  never  been  investigat- 
ed. Several  of  the  enzymes  in  this  pathway  require  Mg++,  or  a  related  cation. 


128  1.    MALONATE 

and  under  certain  conditions  malonate  could  inhibit  through  the  chelation 
of  these  ions.  A  possibly  sensitive  enzyme  is  lactate  dehydrogenase,  which 
functions  in  anaerobic  glycolysis,  although  insufficient  quantitative  work 
has  been  done,  and  it  is  likely  that  the  enzyme  from  different  sources  would 
be  inhibited  to  different  degrees  by  malonate.  It  is  unfortunate  that  so 
little  work  on  the  effects  of  malonate  on  anaerobic  glycolysis  has  been  done, 
since  studies  under  aerobic  conditions  are  always  complicated  by  the  sec- 
ondary reactions  discussed  above.  In  a  number  of  cases,  an  inhibition  of 
aerobic  glycolysis  has  been  observed,  i.e.,  a  decreased  formation  of  lactate 
in  air  or  95%  oxygen,  and  these  inhibitions  are  usually  small.  In  bull  sperm 
(Lardy  and  Phillips,  1943  b),  10  mM  malonate  inhibits  6.7%;  in  beef 
thyroid  homogenates  (Weiss,  1951),  33  mM  malonate  inhibits  8%;  and  in 
rat  liver  supernate  (Aisenbergef  oi.,  1957),  25  mM  malonate  inhibits  15.7%. 
Although  rises  in  lactate  can  be  explained  on  an  indirect  basis,  a  decrease 
in  the  rate  of  lactate  formation  must  usually  be  attributed  to  some  inhibi- 
tion along  the  glycolytic  pathway,  since  it  is  not  very  likely  that  malonate 
would  shift  the  fate  of  pyruvate  from  lactate  to  the  cycle.  Greater  inhibi- 
tions have  been  observed:  in  mouse  brain,  40  mM  malonate  inhibits  lactate 
formation  57%  and  in  mouse  liver  mitochondria  100%,  this  being  taken  as 
evidence  of  some  direct  effect  on  glycolysis  (du  Buy  and  Hesselbach,  1956). 
This  concentration  is,  of  course,  rather  high  and  could  have  depleted  the 
Mg++  from  either  the  3-phosphoglyceraldehyde  dehydrogenase  or  enolase 
systems,  or  could  have  inhibited  lactate  dehydrogenase.  Since  the  substrate 
in  these  cases  was  3-phosphoglyceraldehyde,  an  action  earlier  in  the  pathway 
is  impossible.  There  are  also  miscellaneous  reports  which  might  be  inter- 
preted as  indicating  an  inhibition  of  glycolytic  pathways,  for  example  the 
results  of  Greville  (1936)  on  rat  brain,  in  which  20  mM  malonate  inhibits 
glucose  oxidation  around  50%  and  pyruvate  oxidation  only  around  15%. 
Only  one  instance  of  a  direct  test  on  anaerobic  glycolysis  has  come  to  my 
attention,  that  of  Covin  (1961),  who  found  that  5  mM  malonate  inhibits 
lactate  formation  in  rat  ventricle  slices,  but  the  rate  of  lactate  formation  in 
this  tissue  is  so  slow,  that  Covin  expressed  some  doubts  as  to  the  reliability 
of  the  measurements.  One  must  conclude  from  the  incomplete  data,  that 
there  is  some  evidence  for  a  minor  inhibition  of  glycolysis  by  malonate, 
especially  at  the  higher  concentrations. 

The  problem  of  the  direct  inhibition  of  glycolysis  by  malonate  has  been 
studied  particularly  well  by  Eva  and  George  Fawaz  at  the  American  Univer- 
sity of  Beirut.  They  had  observed  that  30  mM  malonate  almost  completely 
blocks  the  cycle,  leads  to  the  accumulation  of  succinate,  and  yet  does  not 
depress  the  dog  heart  significantly  (Fawaz  et  ol.,  1958).  However,  60  mM 
malonate  causes  rapid  reduction  in  cardiac  frequency  and  contractile 
failure,  although  no  more  succinate  accumulates  than  with  30  mM.  This  de- 
pression of  cardiac  function  must  be  related  to  an  action  other  than  on  the 


EFFECTS  OF  MALONATE  ON  GLUCOSE  METABOLISM  129 

cycle.  The  formation  of  lactate  from  glucose  in  extracts  of  rat  muscle  is 
inhibited  over  90%  by  60  mM  malonate;  simultaneously  there  is  a  decrease 
in  inorganic  phosphate,  creatine-P,  and  readily  hydrolyzable  phosphate, 
accompanied  by  an  increase  in  nonhydrolyzable  phosphate  (Fawaz  and 
Fawaz,  1962).  It  was  concluded  that  the  acid-resistant  phosphate  fraction 
occurring  in  the  presence  of  malonate  must  be  made  up  of  glycolytic  inter- 
mediates, and  it  was  then  shown  by  analyses  of  the  incubated  extracts  at 
various  times  that  there  is  accumulation  of  3-P-glycerate  and  glycerol- 1-P 
particularly,  with  smaller  contributions  from  phosphoenolpyruvate  and 
2-P-glycerate  (see  accompanying  tabulation).  Furthermore,  the   addition 


Incubation 

Accumulation  of  intermediates  (mg  ^ 

P/100  g 

tissue) 

time  (min) 

P-enolpyruvate 

2-P-glycerate 

3-P-glycerate 

Glycerol- 1-P 

6 

6.12 

1.84 

19 

.00 

22 

.30 

30 

8.60 

2.45 

29 

.10 

30. 

,80 

120 

10.90 

3.18 

45 

.60 

33 

,40 

of  these  intermediates  to  a  malonate-treated  extract  resulted  in  the  ap- 
pearance of  a  major  fraction  as  3-P-glycerate,  whereas  in  control  incuba- 
tions they  break  down  to  inorganic  phosphate  and  pyruvate  or  lactate. 
This  might  imply  a  block  at  the  pyruvate  kinase  and  some  reversal  of  the 
glycolytic  pathway  which  cannot  proceed  beyond  3-P-glycerate  due  to  the 
lack  of  ATP.  The  addition  of  pyruvate  to  an  inhibited  extract  leads  mainly 
to  the  formation  of  lactate.  Pyruvate  kinase  from  rabbit  muscle  is  inhibited 
86%  by  60  mM  malonate  and  this  correlates  quite  well  with  the  results 
seen  in  the  extracts.  Malonate  at  30  mM  only  partly  inhibits  glycolysis, 
causes  less  accumulation  of  the  phosphorylated  intermediates,  and  inhibits 
pyruvate  kinase  67%.  The  relative  lack  of  affect  of  30  mM  malonate  on 
cardiac  function  is  probably  due  to  a  combination  of  two  factors:  the  intra- 
cellular malonate  concentration  is  undoubtedly  less  than  30  mM,  and  it 
is  likely  that  glycolysis  in  the  heart  can  be  depressed  to  a  certain  degree 
before  failure  occurs,  the  myocardium  having  other  sources  of  energy 
available.*  Glycolysis  in  dog  muscle  extracts  proceeds  somewhat  differently 
than  in  rat  muscle  extracts  (e.g.  accumulation  of  hexose  phosphates  occurs) 
and  the  response  to  malonate  is  consequently  different  (Fawaz  et  at., 
1963).  High  concentrations  of  malonate  cause  the  accumulation  of  the  same 

*  Since  the  heart  is  generally  considered  to  obtain  much  of  its  energy  from  fatty 
acid  oxidation  under  certain  circumstances,  it  would  be  important  to  know  the  effects 
of  these  high  concentrations  of  malonate  on  this  pathway.  However,  with  the  cycle 
inhibited,  the  generation  of  energy  from  the  oxidation  of  fatty  acids  should  be  reduced. 


130  1.    MALONATE 

intermediates  in  dog  muscle  as  described  above  for  rat  muscle,  but  in  ad- 
dition there  is  an  inhibition  of  the  accumulation  of  fructose-l,6-diP,  so 
that  some  inhibition  of  a  proximal  step  in  glycolysis  seems  likely.  Of  the 
four  enzymes  involved  previous  to  fructose-l,6-diP,  only  phosphogluco- 
mutase  is  sensitive  to  malonate,  40%  inhibition  being  produced  by  60  mM 
malonate  and  76%  inhibition  by  120  mM  malonate.  Thus  a  secondary  site 
for  the  inhibition  of  glycolysis  is  likely.  Other  mechanisms  may  be  involved 
in  intact  muscle  cells. 

Effects   on   the    Distribution    of  C"   from    Labeled    Glucose 

If  glucose  is  metabolized  exclusively  by  the  Embden-Meyerhof  glycolytic 
pathway  and  no  initial  decarboxylation  of  glucose  or  the  hexose  phosphates 
occurs  (as  it  would  if  the  pentose-P  pathway  were  operative),  malonate 
should  depress  the  formation  of  C^'^Og  from  glucose-1-C^*  and  glucose-6-C^* 
equally.  However,  if  the  pentose-P  pathway  is  important,  malonate  should 
decrease  the  formation  of  C^^Og  from  glucose-6-C^*  relatively  more  than  from 
glucose-1-C^*,  and  hence  increase  the  C-l/C-6  ratio.  The  difference  between 
the  C^'^Oo  formed  from  these  precursors  is  often  taken  as  a  measure  of  the 
activity  of  the  pentose-P  pathway;  this  may  not  be  strictly  true  because 
every  hexosephosphate  which  is  decarboxylated  may  not  pass  through  the 
pentose-P  pathway  completely,  but  it  is  certainly  the  best  evidence  for 
the  relative  importance  of  these  two  pathways. 

Experiments  of  this  type  were  performed  with  slices  of  various  tissues 
from  the  rat  (van  Vals  et  al.,  1956).  The  specific  activities  of  the  C^^Og 
formed  from  C-1  and  C-6-labeled  glucose  are  essentially  the  same  in  the 
controls,  indicating  the  pentose-P  pathway  to  be  unimportant  (see  ac- 
companying tabulation).  Malonate  increases  the  C-l/C-6  ratios,  which  was 


C-l/C-6  ratio 
Tissue 


Control        Malonate  30  mM 


Heart 

0.95 

2.38 

Brain 

0.93 

2.18 

Kidnev 

0.99 

1.92 

Diaphragm 

1.02 

1.32 

taken  as  evidence  for  a  malonate-induced  appearance  of  the  pentose-P 
pathway,  although  it  would  account  for  only  a  small  fraction  of  the  glucose 
oxidized.  In  the  lung  and  various  mouse  tumors,  in  which  the  pentose-P 
pathway  is  operative  normally  (C-l/C-6  rations  between  1.4  and  5.6), 
malonate  further  augments  the  importance  of  the  pentose-P  pathway  as 


EFFECTS  OF  MALOXATE  ON  GLUCOSE  METABOLISM  131 

determined  by  C-l/C-6  ratios,  which  are  increased  to  vakies  between  4 
and  125.  A  similar  situation  has  been  encountered  in  sheep  thyroid  shoes 
(the  results  of  three  experiments  are  averaged  in  the  accompanying  tabu- 
lation) (Dumont,  1961).  The  formation  of  C^^Oofrom  glucose-6-C^*  is  inhibit- 

Ci*0,  Control       Malonate  100  mM 


From  glucose-1-C^^  (cpm) 

10.7 

9.72 

From  glucose-6-Ci*  (cpm) 

4.1 

0.47 

C-l/C-6 

2.58 

20.5 

(C-1)  — (C-6) 

6.6 

9.25 

ed  strongly,  whereas  that  from  ghicose-1-C^^  is  scarcely  affected.  This,  of 
course,  indicates  an  almost  complete  block  of  the  cycle,  which  is  not  sur- 
prising at  this  high  malonate  concentration,  but  it  also  suggests  a  greater 
participation  of  the  pentose-P  pathway  in  the  presence  of  malonate.  These 
results  are  perhaps  more  understandable  in  the  light  of  the  glycolytic 
inhibitions  by  high  malonate  concentrations  discussed  in  the  previous 
section.  Quite  different  results  were  obtained  in  electrically  stimulated  rat 
ventricle  strips   (see   accompanying  tabulation)    in   which   the   pentose-P 

Ci^O,  Control       Malonate  5.6  mM 


From  glucose- 1-C^*  (cpm) 

0.318 

0.271 

From  glucose-6-C'''  (cpm) 

0.285 

0.264 

C-l/C-6 

1.12 

1.03 

(C-1)  -  (C-6) 

0.033 

0.007 

pathway  is  presumably  not  important,  malonate  at  this  concentration 
having  little  effect  on  glucose  utilization  (Rice  and  Berman,  1961).  It  is 
possible  that  higher  concentrations  of  malonate  would  produce  changes 
such  as  observed  with  other  tissues.  However,  this  concentration  of  mal- 
onate is  quite  effective  in  modifying  ventricular  function. 

An  indirect  mechanism  for  the  acceleration  of  the  pentose-P  pathway  by 
malonate  may  involve  the  levels  of  NADP  and  ATP  in  the  tissues.  The  oxida- 
tive decarboxylation  of  glucose-6-P  to  initiate  this  pathway  requires  NADP, 
the  concentration  of  which  may  be  changed  due  to  the  action  of  malonate  on 
the  cycle.  Also  the  phosphorylation  of  fructose-6-P  in  the  Embden-Meyerhof 
pathway  requires  ATP,  the  level  of  which  may  be  reduced  by  high  concen- 
trations of  malonate.  However,  little  is  known  about  the  control  of  the 
pentose-P  pathway  and  further  experiments  are  needed  to  elucidate  its 


132  1.    MALONATE 

role  during  malonate  inhibition.  Nothing  is  known  of  the  possible  direct 
effects  of  malonate  on  the  pentose-P  pathway.  The  production  of  C^^Og 
from  ribose-l-C^*  is  inhibited  strongly  by  malonate  in  heart  homogenates 
(Jolley  et  al.,  1958)  and  it  is  believed  that  ribose  is  metabolized  in  this 
pathway,  but  the  inhibition  may  reflect  an  action  on  the  cycle.  It  is  inter- 
esting that  fluoroacetate  does  not  inhibit  C^^Og  formation  very  potently, 
except  at  very  high  concentrations,  so  that  some  direct  effect  on  ribose  meta- 
bolism is  possible.  The  results  of  the  experiments  discussed  above  would 
argue  against  this.  D-Xylose  and  D-ribose-5-P  are  oxidized  through  se- 
doheptulose-7-P  in  extracts  of  Pseudomonas  hydrophila  and  this  is  not  af- 
fected by  malonate  at  20  mikf  (Stone  and  Hochster,  1956). 

The  only  investigation  of  the  effects  of  malonate  on  the  general  distri- 
bution of  C^^from  labeled  glucose  is  that  of  Romberger  and  Norton  (1961)  in 
potato  tuber  slices  incubated  with  uniformly  labeled  glucose  for  3  hr  (Table 
1-19).  The  situation  in  this  tissue  is  complex  inasmuch  as  fresh  slices  do  not 
metabolize  glucose  appreciably,  whereas  36-hr-old  slices  oxidize  it  quite 
rapidly.  In  the  aged  tissue,  CO2  production  is  inhibited  92%  by  50  mM  mal- 
onate at  pH  5,  while  the  formation  of  CO2  in  fresh  tissue  is  stimulated  28%. 
In  the  fresh  tissue,  they  suggest  that  CO2  is  formed  mainly  in  the  pentose-P 
pathway  and  little  through  the  Embden-Meyerhof  sequence,  but  glycolysis 
contributes  more  and  more  with  time,  so  that  the  marked  inhibition  by 
malonate  in  aged  tissue  is  not  surprising.  The  synthesis  of  sucrose  accounts 
for  over  half  of  the  labeling  from  glucose-u-C^*  and  this  is  inhibited  only 
11%  by  malonate.  After  3-hr  incubation,  however,  one  can  deduce  little 
about  the  initial  attack  on  glucose.  It  may  be  mentioned  in  this  connection 
that  in  Acetobacter  xylmum,  where  the  sole  product  of  glucose  assimilation 
is  cellulose,  malonate  at  10  mM  does  not  inhibit  the  formation  of  cellulose. 
Laties  (1964)  has  investigated  this  problem  in  detail  and  found  that  dif- 
ferent methods  of  aging  result  in  metabolically  different  potato  slices, 
in  that  some  exhibit  a  malonate-sensitive  and  some  a  malonate-resistant 
respiration.  In  the  malonate-sensitive  slices  the  formation  of  C^^Og  from 
labeled  glucose  is  almost  completely  abolished  by  malonate,  whereas  in  the 
malonate-resistant  slices  there  is  little  effect  by  malonate  on  the  production 
of  C^^Og.  There  is  a  similar  correlation  with  respect  to  the  effects  of  malonate 
on  glucose  uptake.  Since  dinitrophenol  does  not  interfere  with  glucose 
uptake,  one  can  eliminate  depression  of  ATP  formation  by  malonate  as 
responsible  for  the  inhibition  of  the  uptake  in  sensitive  slices.  Laties  con- 
sidered the  possibility  that  increased  citrate  levels  might  inhibit  phos- 
phofructokinase,  but  the  mechanism  is  not  yet  well  understood. 

Effects  of  Tissue  Age  on  the  Response  of  Glucose  Metabolism  to  Malonate 

We  have  seen  above  that  aging  of  potato  slices  increases  their  sensitivity 
to  malonate  with  respect  to  glucose  oxidation.  It  might  be  expected  that 


EFFECTS  OF  MALONATE  ON  GLUCOSE  METABOLISM  133 

the  metabolic  characteristics  of  tissues,  would  change  with  age  and  that 
this  would  be  reflected  in  difl"erent  susceptiblies  to  inhibitors.  The  altered 
response  to  inhibitors  could  provide  some  information  on  the  nature  of 
metabolic  aging.  The  results  obtained  on  rat  brain  are,  however,  discordant. 
Tyler  (1942),  using  a  minced  preparation,  found  that  malonate  inhibition 
of  the  respiration  in  the  presence  of  glucose  increases  up  to  a  rat  age  of  10 
days,  after  which  it  remains  at  the  adult  level  (see  accompanying  tabula- 
tion). Although  the  control  respiration  rises,  the  malonate-resistant  fraction 


Rat  age  (days) 

Control 

Malonate  10 

mi¥ 

"o  Inhibition 

1 

656 

547 

16.9 

2 

582 

507 

12.9 

6 

814 

590 

27.6 

8 

712 

559 

34.3 

10 

1080 

530 

50.9 

16 

1500 

810 

46.0 

31 

1900 

960 

49.5 

Adult 

1723 

864 

50.0 

of  the  respiration  remains  constant  up  to  10  days,  i.e.,  the  increased  respi- 
ration is  all  due  to  the  development  of  activation  of  a  system  sensitive  to 
malonate,  presumably  the  cycle.  On  the  other  hand,  Muir  et  al.  (1959) 
reported  that  the  glucose  respiration  of  adult  rat  brain  slices  is  less  sensitive 
to  10  inM  malonate  (30%  inhibition)  than  the  respiration  of  tissue  from 
young  animals  of  1-3  days  (70%).  In  this  case,  the  young  brain  respires 
almost  2.5  times  as  rapidly  as  adult  brain.  The  differences  in  these  obser- 
vations may  be  related  to  the  preparations  used  (mince  or  slice).  There  are 
several  reasons  why  malonate  sensitivity  would  change  with  age,  for  exam- 
ple, an  alteration  of  cycle  activity,  the  development  or  loss  of  pathways 
other  than  the  cycle  for  the  metabolism  of  acetyl-CoA,  or  a  change  in  the 
ability  to  demonstrate  a  Pasteur  effect.  The  effect  of  age  should  also  be 
studied  on  the  electrically  stimulated  or  K+-stimulated  respiration  of  the 
brain,  since  it  is  more  sensitive  to  malonate  and  the  results  might  have 
more  physiological  pertinence. 

The  respiratory  rates  and  patterns  of  fungus  spores  are  altered  during 
the  initiation  of  germination  and  the  subsequent  development  of  the  germ 
tube  (Gottlieb,  1964).  During  the  incubation  of  the  spores  of  Penicillium 
oxalicum  and  Ustilago  maydis  and  the  progress  of  germination,  the  respira- 
tion in  the  presence  of  glucose  rises  markedly  and  this  is  accompanied  by  an 
increasing  sensitivity  to  malonate  (see  accompanying  tabulation)  (Caltrider 
and  Gottlieb,  1963).  It  is  somewhat  difficult  to  determine  if  this  implies 


134 


1.    MALONATE 


an  increase  in  cycle  activity  since  the  concentration  of  malonate  was  100 
mM  and  more  than  the  cycle  might  be  inhibited. 


Incubation  (hr) 

Respiration 

0/ 

/o 

Inhibition 

0 

1.3 

0 

6 

6.0 

16 

9 

11.1 

22 

12 

23.5 

32 

Effects  on   Resting  and   Stimulated   Glucose   Metabolism   in   Brain 

The  effects  of  malonate  on  a  tissue  may  depend  on  the  activity  of  the 
tissue  as  well  as  the  age.  Stimulation  of  a  tissue  such  as  brain  alters  the 
metabolic  pattern  and  this  is  exhibited  in  altered  responses  to  inhibitors. 
The  results  obtained  by  Heald  (1953)  on  guinea  pig  cerebral  cortex  slices 
are  shown  in  Fig.  1-14.  The  resting  respiration  and  aerobic  glycolysis  are 


(  Malonote  ! 


10       100 
mM 


Fig.  1-14.  Effects  of  malonate  on  guinea  pig  brain 
slices  with  glucose  as  the  substrate.  Electrical  stimula- 
tion through  grids.  The  respiration  and  aerobic  glycol- 
ysis in  //moles/g  wet  weight/hr.  (From  Heald,  1953.) 


EFFECTS   OF   MALONATE   ON  LIPID  METABOLISM  135 

little  affected  by  malonate  up  to  10  mM,  whereas  the  electrically  stimulated 
respiration  is  readily  depressed  (down  to  the  resting  level  at  10  mM  mal- 
onate) and  the  stimulated  lactate  formation  markedly  increased.  This  means 
that  the  glucose  metabolism  appearing  upon  stimulation  is  quite  sensitive 
to  malonate  and  perhaps  involves  a  greater  participation  of  the  cycle. 
These  results  were  confirmed  by  Kimura  and  Niwa  (1953)  in  guinea  pig 
brain  stimulated  by  K+,  and  a  stimulation  of  lactate  formation  by  10  mM 
malonate  was  observed  by  Tsukada  and  Takagaki  (1955).  An  abolition  of 
the  inhibition  of  respiration  upon  addition  of  fumarate  occurs  (Takagaki  et 
at.,  1958).  Rat  brain  slices  behave  similarly,  the  resting  respiration  in  the 
presence  of  glucose  being  unaffected  by  malonate  up  to  0.8  mM,  while  the 
stimulated  respiration  is  readily  suppressed  (Wallgren,  1960).  A  malonate 
concentration  as  low  as  0.2  mM  inhibits  the  stimulated  respiration  15%. 
Pyruvate  utilization  and  the  associated  oxygen  uptake  are  also  inhibited 
more  strongly  in  stimulated  slices  than  in  resting  slices  (Takagaki  et  al., 
1958).  The  C^^Og  from  labeled  pyruvate  is  formed  about  twice  as  rapidly  in 
high  K+  medium  compared  to  the  controls  (Kini  and  Quastel,  1959),  and 
this  is  inhibited  more  strongly  by  malonate  in  the  K+-stimulated  slices, 
while  the  stimulated  respiration  is  depressed  to  the  endogenous  level. 

These  results  taken  together  clearly  indicate  a  dependency  of  malonate 
inhibition  on  the  metabolic  activity  of  brain  tissue,  whether  altered  by 
electrical  stimulation  or  K+.  A  Pasteur  effect  is  observed  and  it  is  possible 
that  the  inhibition  by  malonate  would  have  been  greater  if  it  had  not 
induced  a  greater  utilization  of  glucose.  The  data  do  not  necessarily  imply 
a  specific  activation  of  the  cycle;  a  greater  uptake  or  utilization  of  glucose 
would  impose  a  greater  load  on  the  cycle,  and  this  might  be  inhibited  more 
readily.  Whatever  the  explanation  for  these  effects,  such  results  have 
important  bearing  on  the  actions  of  malonate  on  intact  and  functioning 
nervous  tissue. 


EFFECTS   OF   MALONATE   ON    LIPID   METABOLISM 

The  major  pathway  for  fatty  acid  oxidation  is  a  helical  degradation  into 
acetyl-CoA,  which  normally  enters  the  cycle  by  condensation  with  oxal- 
acetate.  Each  turn  of  the  helix,  releasing  one  acetyl-CoA,  takes  up  2  atoms 
of  oxygen,  and  the  complete  oxidation  of  acetyl-CoA  through  the  cycle 
takes  up  4  more  atoms  of  oxygen.  Thus,  approximately  two-thirds  of  the 
oxygen  uptake  due  to  fatty  acid  oxidation  occurs  in  the  cycle,*  and  one 
would  expect  malonate  to  depress  this  fraction  in  proportion  to  the  cycle 

*  The  term  "cycle."  as  before,  will  indicate  the  tricarboxylate  cycle  only,  and  the 
pathway  of  degradation  of  fatty  acids  to  acetyl-CoA  and  other  terminal  products 
will  be  designated  the  "helix"  for  convenience. 


136  1.    MALONATE 

block  it  induces.  The  situation  is  vey  similar  to  that  of  ghicose  oxidation  in 
this  respect.  Generally  speaking,  malonate  could  act  on  either  the  cycle, 
or  the  helix,  or  both.  Despite  the  extensive  work  that  has  been  done  on  the 
effects  of  malonate  on  fatty  acid  oxidation,  direct  information  on  the  actions 
on  the  helix  is  lacking,  since  the  five  reactions  involved  in  each  turn  of  the 
helix  and  the  enzymes  associated  with  these  have  not  all  been  tested  for 
susceptibility  to  malonate,  nor  has  the  operation  of  the  helix  dissociated 
from  the  cycle  been  studied.  Our  evidence  on  this  point  must  be  indirect. 

Before  considering  this  evidence,  let  us  outline  some  of  the  possibilities 
for  mechanisms  of  helix  inhibition.  Just  as  in  the  oxidation  of  glucose,  ATP 
is  required  for  the  initiation  of  the  helix  reactions  and  Mg++  is  a  necessary 
cofactor  (e.g.,  for  the  fatty  acid  thiokinase),  so  that  malonate  might  depress 
the  operation  of  the  helix  by  depleting  the  system  of  either  of  these  sub- 
stances. The  extent  of  such  an  inhibition  will  depend  on  the  availablity  of 
ATP  or  the  presence  of  systems  generating  it,  and  on  the  concentration  of 
Mg++.  Possibly  a  more  important  factor  is  the  requirement  for  coenzyme  A. 
Malonate  could  deplete  conzyme  A  by  at  least  two  mechanisms.  The  for- 
mation of  a  relatively  stable  malonyl-CoA  would  remove  some  of  the 
coenzyme  A  from  participating  in  the  helix.  A  block  of  the  cycle  would 
impede  the  entrance  of  acetyl-CoA  into  the  cycle  and  the  regeneration  of 
coenzyme  A  will  depend  on  the  enzymes  present  for  the  metabolism  of 
acetyl-CoA.  The  usual  pathways  for  acetyl-CoA  are  (1)  a  simple  splitting 
to  form  acetate,  (2)  a  transfer  of  the  coenzyme  A  to  another  acid,  and  (3) 
a  condensation  of  two  acetyl-CoA's  to  form  acetoacetyl-CoA  and  eventually 
acetoacetate.  As  in  the  oxidation  of  pyruvate  through  the  cycle,  the  fate 
of  acetyl-CoA  will  depend  also  on  the  presence  of  reactions  forming  oxal- 
acetate  by  pathways  other  than  the  cycle.  The  rate  of  fatty  acid  oxidation 
can  thus  be  limited  by  the  rate  of  regeneration  of  coenzyme  A.  These 
considerations  lead  one  to  ])redict  that  the  effects  of  malonate  on  fatty 
acid  oxidation  would  be  variable  and  dependant  on  the  metabolic  charac- 
teristics of  the  tissue  studied  and  the  conditions  of  the  experiment.  This 
prediction  is  borne  out. 

There  is  a  good  deal  of  evidence  that  malonate  in  concentrations  up  to 
20  raM  does  not  directly  inhibit  the  reactions  of  the  helix.  Although  an 
inhibition  of  the  oxygen  uptake  or  the  CO2  production  during  fatty  acid  oxi- 
dation is  not  indicative  of  the  site  of  inhibition  when  the  helix  and  the 
cycle  are  operating  together,  the  absence  of  inhibition  implies  a  lack  of 
action  on  the  helix.  Malonate  at  16.8  mM  has  no  effect  on  the  C^^Oa  arising 
from  palmitate-1-C^*  in  soluble  extracts  of  peanut  cotyledons  (Castelfranco 
et  al.,  1955),  nor  does  1  roM  malonate  have  an  effect  on  the  oxygen  uptake 
associated  with  palmitate  oxidation  in  peanut  microsomes  (Humphreys 
et  al.,  1954).  The  anaerobic  dehydrogenation  of  C4-C18  fatty  acids  in  liver 
homogenates   with   methylene   blue  as  an   acceptor  is  not  inhibited  by 


EFFECTS   OF   MALONATE   ON   LIPID   METABOLISM  137 

10  raM  malonate  (Blakley,  1952).  The  rate  of  oxidation  of  decanoate  by 
Serratia  marcescens  is  also  not  affected  by  10  mM  malonate,  although 
20  m.M  inhibits  somewhat  (Waltman  and  Rittenberg,  1954).  Geyer  and 
Cunningham  (1950)  stated  that  their  data  indicated  no  inhibition  directly  of 
octanoate  oxidation  in  liver  by  5  mM  malonate  (this  work  will  be  discussed 
in  greater  detail  later). 

On  the  other  hand,  Lehninger  and  Kennedy  (1948)  reported  that  10  mil/ 
malonate  stronglj'  inhibits  octanoate  oxidation  in  particulate  suspensions, 
from  rat  liver.  Not  only  is  the  respiration  from  the  oxidation  inhibited 
but  the  utilization  of  octanoate  is  almost  completely  suppressed.  The  ad- 
dition of  malate  or  oxalacetate  reduces  the  inhibition  only  partially,  the 
utilization  of  octanoate  still  being  inhibited  around  70%.  It  may  be  noted 
that  the  total  Mg+"'"  concentration  in  these  experiments  was  0.25  mM, 
which  is  quite  low,  so  that  malonate  could  have  inhibited  by  depletion  of 
this  cofactor.  An  interesting  point  is  that  the  strain  of  rats  used  is  very 
important,  since  10  mM  malonate  inhibits  octanoate  oxidation  25%  in 
preparations  from  livers  of  Sprague-Dawley  rats,  but  in  preparations  made 
from  a  heterogeneous  stock  colony  2  mM  malonate  inhibits  50-75%. 
Such  differences  in  strain  behavior  may  explain  some  of  the  discrepancies  in 
the  reports  on  malonate  inhibition.  Weinhouse  et  al.  (1949)  found  that 
20  mM  malonate  almost  completely  blocks  the  oxidation  of  octanoate  in  rat 
liver  slices  and  that  fumarate  only  partially  overcomes  this,  suggesting  to 
them  that  malonate  must  have  some  action  other  than  on  succinate  oxidase. 
In  several  instances  malonate  has  been  found  to  inhibit  the  oxygen  uptake 
from  fatty  acid  oxidation  very  potently.  Butyrate  respiration  in  peanut 
mitochondria  is  inhibited  75%  by  6  mM  malonate  and  the  formation  of 
C^^Og  from  butyrate-l-C^*  is  depressed  even  more  strongly  (Stumpf  and 
Barber,  1956).  Malonate  at  10  mM  inhibits  the  oxidation  of  octanoate  by 
carp  liver  mitochondria  80%  (Brown  and  Tappel,  1959).  In  suspensions  of 
particulates  from  desert  locust  thorax,  butyrate  oxidation  is  inhibited  70% 
by  malonate  at  concentrations  as  low  as  1  mM  and  maximally  85%  (Meyer 
et  al.,  1960).  The  oxidation  of  octanoate- 1-C^*  and  myristate-1-C^*  by  sub- 
cellular particles  from  the  lateral  line  of  the  rainbow  trout,  as  measured 
by  the  C^^Oj  released,  is  reduced  95%  by  10  mM  malonate  (Bilinski  and 
Jonas,  1964).  One  of  the  most  sensitive  systems  is  found  in  the  oxidation  of 
linolenate  in  liver  mitochondria  of  vitamin  E-deficient  chicks,  0.25  mM 
malonate  inhibiting  40%  (Kimura  and  Kummerow,  1963).  These  examples 
must  be  interpreted  as  indicating  either  a  direct  or  an  indirect  inhibition 
of  the  helix  by  malonate.  Finally,  it  was  stated  by  Mudge  (1951),  on  the  basis 
of  unpublished  experiments,  that  malonate  inhibits  fatty  acid  oxidation 
more  strongly  than  succinate  dehydrogenase  in  kidney  particulate  prepa- 
rations. The  possibility  of  effects  on  the  helix  must  therefore  be  entertained 
on  the  basis  of  our  present  knowledge. 


138 


1.    MALONATE 


Effects  on  the  Formation  of  Acetoacetate  and  Other  Ketones 

It  was  known  before  1912  that  acetate,  oxalate,  and  maleate  can  either 
be  metabolized  to  acetoacetate  or  so  alter  metabolism  that  acetoacetate  ac- 
cumulates, and  for  this  reason  Momose  (1914)  in  Berlin  studied  the  effects 
of  malonate  perfused  through  starved  dog  livers  at  a  concentration  of 
approximately  13  mM.  He  found  that  acetone  appears  and  detected  a  sub- 
stance which  he  only  later,  after  returning  to  Japan  (Momose,  1925),  proved 
was  acetoacetate.  However,  he  postulated  that  malonate  -^  acetate  — >■ 
acetoacetate  -^  acetone,  which  was  not  unreasonable  considering  the  inhi- 
bitory action  of  malonate  was  unknown.  The  appearance  of  acetone  in  the 
urine  of  rabbits  fed  malonate  or  rats  injected  subcutaneously  with  malonate 
was  observed  by  Huszak  (1935),  and  simultaneously  Annau  (1935)  dem- 
onstrated that  malonate  causes  the  formation  of  acetone  in  slices  and 
breis  of  rabbit  kidney.  Acetoacetate  has  been  shown  to  accumulate  in 
tissues  as  a  response  to  malonate  (see  accompanying  tabulation).  Since 
acetoacetate  and  acetone  are  the  most  important  ketonic  substances  ap- 
pearing in  the  tissues,  these  results  clearlj^  show  that  malonate  is  ketogenic. 


Preparation 

Substrate 

Malonate 

Reference 

Whole  rabbits  (blood) 

Endogenous 

1.6g/kg 

Handler  (1945) 

Whole  rats  (blood) 

Endogenous 

O.S 

!g/kg 

Mookerjea  and  Sadhu  (1955) 

Rat  liver  slices 

Acetate 

40 

mM 

Jowett  and  Quastel  (1935  c) 

Butyrate 

40 

mM 

Jowett  and  Quastel  (1935  c) 

Guinea  pig  liver  slices 

Propionate 

40 

mM 

Jowett  and  Quastel  (1935  c) 

Butyrate 

40 

mM 

Jowett  and  Quastel  (1935  c) 

Rat  liver  slices 

Endogenous 

10 

mM 

Edson  (1936) 

Rat  liver  slices 

Fatty  acids 

5 

mM 

Geyer  and  Cunningham  (1950) 

Rat  liver  suspension 

Fatty  acids 

10 

mM 

Lehninger  (1946  a) 

Rat  liver  homogenates 

Pyruvate 

4 

mM 

Recknagel  and  Potter  (1951) 

Rabbit  liver  mitochondria 

Fatty  acids 

15 

mM 

Cheldelin  and  Beinert  (1952) 

Jensen  rat  sarcoma  mince 

Glucose 

6.7  mM 

Boyland  and  Boyland  (1936) 

Peanut  mitochondria 

Butyrate 

6 

mM 

Stumpf  and  Barber  (1956) 

Acetoacetate  is  an  important  substance  in  intermediary  metabolism  and 
the  pathways  for  its  formation  and  utilization  are  often  complex.  The  con- 
centration of  acetoacetate  will  depend  on  the  relative  rates  of  its  formation 
and  utilization.  The  accumulation  of  acetoacetate  in  the  presence  of  mal- 
onate could  result  from  either  an  acceleration  of  its  formation  or  an  inhibi- 
tion of  its  utilization,  or  both.  The  earliest  concept  that  malonate  itself 
gives  rise  to  the  acetoacetate  was  soon  abandoned,  and  several  investigators 
assumed  that  malonate  interferes  with  the  disposal  of  acetoacetate,  while 


EFFECTS  OF  MALONATE   ON  LIPID   METABOLISM 


139 


recently  more  emphasis  has  been  placed  on  the  diversion  of  carbohydrate 
and  fatty  acid  metabolism  to  acetoacetate  by  malonate.  We  may  summarize 
some  of  the  major  pathways  of  acetoacetate  before  discussing  the  mecha- 
nisms for  the  action  of  malonate  (Ac-CoA  =  acetyl-CoA,  and  AcAc-CoA  = 
acetoacetyl-CoA).  Reaction  (1)  for  the  formation  of  Ac  Ac-Co  A  from  aceto- 


Fatty  acids  ( 
Pyruvate  ij 


Butyrate 


-»-  Ac -Co A 


AcAc  —  CoA 


-»-  Ac  Ac  — CoA 


Even-numbered 
fatty  acids 

•-Butyryl  —CoA *-  Ac  Ac  — CoA 

Succinate      +       Ac  Ac— CoA 

/3-Hydroxybutyrate 

Amino  acids  (phenylalanine, 
tyrosine,  leucine,  and 
isoleucine) 


Acetone  +  COj 
.AcAc-CoA  (1) 

I  acetoacetate  I ' *-AcAc— CoA  (2) 

Sterols 
/3-Hydroxybutyrate 


acetate  is  catalyzed  by  an  activating  enzyme  in  the  presence  of  CoA  and 
ATP,  while  reaction  (2)  is  catalyzed  by  a  CoA  transferase  in  the  presence 
of  succinyl-CoA.  All  of  these  reactions  do  not  occur  in  a  single  tissue  and 
the  response  to  malonate  depends  in  part  on  which  reactions  are  possible 
in  any  case. 

A  block  of  the  cycle  restricts  the  entrance  of  acetyl-CoA,  derived  from 
pyruvate  and  fatty  acids,  into  the  cycle,  unless  there  is  an  adequate  synthe- 
sis of  oxalacetate  from  a  noncycle  source,  which  is  seldom  the  case.  If  the 
acetyl-CoA  accumulates,  coenzyme  A  soon  becomes  tied  up  and  the  oxida- 
tion of  pyruvate  and  fatty  acids  would  cease.  However,  in  most  tissues  2 
molecules  of  acetyl-CoA  condense  to  form  acetoacetate  and  coenzyme  A  is 
regenerated;  in  other  situations,  hydrolysis  to  acetate  may  occur.  Malonate 
may  thus  divert  acetyl-CoA  from  the  cycle  to  acetoacetate.  Quantitative 
conversion  to  acetoacetate  has  been  observed  (Recknagel  and  Potter,  1951). 
Another  reaction  possibly  favoring  acetoacetate  formation  during  malonate 
inhibition  results  from  the  accumulation  of  succinate,  which  can  now  react 
more  readily  with  acetoacetyl-CoA  in  a  transfer  of  coenzyme  A.  The  effec- 
tiveness of  such  a  mechanism  depends  on  the  continued  formation  of  suc- 
cinate and,  hence,  usually  on  a  noncycle  source  of  oxalacetate.  Acetoacetyl- 
CoA  is  also  formed  as  the  terminal  product  of  the  helical  oxidation  of  even- 
numbered  fatty  acids.  These  relationships  are  illustrated  in  Fig.  1-15 
where  a  block  of  succinate  oxidation  induces  accumulation  of  acetoacetate 
by  accelerating  its  formation  through  two  mechanisms.  The  other  pathways 
for  the  formation  of  acetoacetate  are  probably  less  important  in  most  tis- 
sues and  would  not  be  accelerated  by  malonate.  Accumulation  of  aceto- 
acetate implies  that  its  utilization  must  not  be  too  rapid.  Liver  is  notable  in 
this  respect  because  it  lacks  enzymes  to  metabolize  acetoacetate  rapidly, 


140  1.    MALONATE 

especially  the  activating  system  for  the  formation  of  acetoacetyl-CoA,  and 
possesses  an  active  deacylase  to  split  acetoacetyl-CoA.  Therefore,  acetoace- 
tate  accumulation  is  most  readily  observed  in  liver  and  most  investi- 
gations have  been  on  this  tissue.  The  urinary  acetoacetate  found  after  the 
administration  of  malonate  is  probably  derived  mainly  from  liver.  In  heart, 
on  the  other  hand,  the  enzyme  balance  is  such  as  to  favor  the  rapid  me- 
tabolism of  acetoacetate  and  it  does  not  accumulate.  Acetate  rather  than 
acetoacetate  accumulates  in  some  cells,  for  example  in  heart  mitochondrial 
suspensions  metabolizing  pyruvate  in  the  presence  of  8.8  mM  malonate 
(Fuld  and  Paul,  1952). 


(Malonate) 
a  -Ketoglutorate      ^  Succmyl-  Co  A Succinate  -X"^  Fumorcte  — ^  —^Citrate 


Fatty    acids  Pyruvate 


Fig.   l-l.'j.  Diagram  of  some  pathways  involved  in  the  effects 
of  malonate  on  the  metabolism  of  acetoacetate. 

One  would  predict  that  fumarate  should  counteract  the  ketogenic  activi- 
ty of  malonate  because,  by  supplying  oxalacetate,  acetyl-CoA  will  again  be 
able  to  enter  the  cycle.  However,  it  may  be  noted  that  fumarate  may  lead 
to  an  even  greater  accumulation  of  succinate  and  if  the  formation  of  aceto- 
acetate by  the  transfer  of  coenzyme  A  from  acetoacetyl-CoA  to  succinate  is 
important,  fumarate  will  only  augment  the  malonate  effect.  Administration 
of  fumarate  with  malonate  to  rabbits  abolishes  the  appearance  of  acetone 
that  arises  with  malonate  alone  (Huszak,  1935).  Addition  of  fumarate  to 
malonate-inhibited  minces  of  rat  sarcoma  likewise  prevents  the  accumula- 
tion of  acetone  bodies  (Boyland  and  Boyland,  1936).  However,  fumarate 
has  very  little  effect  on  the  appearance  of  acetoacetate  in  rat  liver  slices 
inhibited  by  malonate  (Edson,  1936),  and  this  might  indicate  a  mechanism 
for  malonate  action  other  than  the  inhibition  of  succinate  oxidation,  or  the 
importance  of  the  coenzyme  A  transfer  reaction. 

It  is  now  easy  to  see  how  malonate  can  reduce  the  oxygen  uptake  and 
the  CO2  production  from  fatty  acid  oxidation  without  necessarily  decreasing 
the  utilization  of  the  fatty  acids.  A  fraction  that  would  normally  be  com- 
pletely oxidized  is  diverted  into  acetoacetate  (or  acetate,  acetone,  and  other 
products).  One  of  the  best  indications  that  malonate  does  not  inhibit  the 
helix  directly  is  the  fact  that  the  C^Oo  appearing  in  the  end  products  from 
labeled  fatty  acid  is  not  reduced  by  malonate.  To  illustrate  this  it  will  be  con- 
venient to  turn  to  the  excellent  studies  of  Geyer  and  his  group  at  Harvard. 


EFFECTS    OF   MALONATE    ON    LIPID    METABOLISM  141 

The  basic  procedure  in  these  investigations  was  to  incubate  carboxyl- 
labeled  fatty  acids  with  rat  liver  and  kidney  slices,  and  determine  the  dis- 
tribution of  C^*  in  acetoacetate  and  COg.  Malonate  at  5  mM  depresses  the 
formation  of  C^^Og  from  octanoate-C^*  00^  around  60%  and  fumarate  is 
able  to  overcome  this  inhibition  only  partially  (Geyer  et  al.,  1950  a).  Fu- 
marate and  other  cycle  intermediates  increase  the  total  COg  formed  but 
have  little  effect  on  the  C^^Oa-  This  was  explained  by  the  accumulation  of 
some  of  the  C^*  as  succinate,  this  not  being  relieved  by  fumarate,  and  we 
have  previously  cited  this  as  an  example  of  the  importance  of  considering 
what  is  measured  in  demonstrating  a  reversal  by  fumarate. 

Where  does  the  C^*  go  that  does  not  appear  as  C^^Oo  in  inhibited  slices? 
They  found  that  in  the  presence  of  malonate  much  of  the  C^*  appears  in 
acetoacetate  (Table  1-24)  (Geyer  and  Cunningham,  1950).  The  ratio  AcAc- 
QujQuQ^  is  near  1.21  in  the  controls  and  is  increased  to  around  4.51  by 
malonate,  averaging  the  results  from  the  five  fatty  acids  used.  It  may 
also  be  noted  that  malonate  generally  increases  the  total  C"  recovered,  even 
though  succinate  was  not  determined,  showing  that  malonate  does  not  in- 
hibit the  fatty  acid  oxidation  directly.  Later  they  determined  both  the  car- 
boxyl  and  carbonyl  C^*  in  acetoacetate  and  the  more  complete  results  are 
summarized  in  Table  1-25,  where  I  have  taken  the  liberty  of  averaging 
the  data  for  all  the  fatty  acids  used,  inasmuch  as  the  effects  are  always 
in  the  same  direction  although  differences  between  the  different  fatty  acids 
are  evident.  These  results  show  clearly  the  diversion  of  fatty  acid  metabo- 
lism into  acetoacetate  by  malonate.  Weinhouse  et  al.  (1949)  reported  that  in 
rat  liver  slices  10  mM  malonate  inhibits  COo  formation  and  no  acetoacetate 
appears,  which  was  so  contradictory  to  the  results  obtained  by  Geyer  that 
the  latter  studied  malonate  in  concentrations  up  to  20  mM,  but  found  only 
that  even  more  acetoacetate  accumulates.  Also  they  tested  three  different 
strains  of  rat  and  the  results  were  the  same.  The  reason  for  this  discrepancy 
could  not  be  explained. 

The  differential  labeling  in  the  carboxyl  and  carbonyl  groups  of  aceto- 
acetate is  difficult  to  explain.  If  acetoacetate  arises  by  a  condensation  of 
acetyl-CoA  units,  the  labeling  in  these  positions  should  be  uniform.  However, 
the  ratio  is  seldom  unity  as  may  be  seen  in  the  results  summarized  by  Chai- 
koff  and  Brown  (1954).  In  the  work  of  Geyer  with  rat  liver  slices,  the  ratio 
CHgC^^  0 — / — CHoC^*  00"  is  less  than  1  in  the  controls  and  increases  with 
the  length  of  the  fatty  acid  chain.  Malonate  increases  this  ratio,  that  is,  it 
increases  relatively  the  labeling  in  the  carbonyl  group.  Chaikoff  and  Brown 
have  given  a  detailed  analysis  of  the  possible  factors  determining  this  ratio, 
and  the  explanation  is  based  on  the  existence  of  two  types  of  2-carbon 
fragment  formed  from  fatty  acids,  one  designated  the  CH3CO —  fragment 
and  the  other  the  — CH2CO —  fragment.  These  fragments  are  assumed 
to  arise  from  different  portions  of  the  fatty  acid  chain  and  only  the  — CHg 


142 


1.    MALONATE 


Q 

T 

o 

>H 

<; 

H- 

< 

fo 

O 

Q 

W 

6 

iJ 

w 

K 

< 

J 

ij 

>H 

X 

o 

m 

P3 

•rf 

< 

6 
1 

O 

o 

<1 

53 

o 

o 


o 


o 


Q 

(B 

-P 

-kJ 

tS 

cS 

Cj 

O 

_« 

^ 

> 

-2 

a) 

o 

c 

+3 

O 

c 

o 


fa 


EFFECTS   OF  MALONATE   ON   LIPID   METABOLISM 


143 


Table  1-25 

Effects  of  5  mM  Malonate  on  the  Oxidation  of  Fatty  Acids  in  Rat  Liver 

Slices  " 


Control 

Malonate 

Change 

C^Og  produced 

10,694 

4,536 

-  6,158 

AcAc-carbonyl-C*  formed 

3,572 

8,488 

+  4,916 

AcAc-carboxyl-Ci^  formed 

7,268 

13,438 

+  6,170 

CHaCi^O— /— CH^Ci^OO- 

0.47 

0.63 

Total  AcAc-Ci*  formed 

10,840 

21,926 

+  11,086 

C'^Oa  +  AcAc-C"  formed 

21,534 

26,462 

+  4,928 

AcAc-CiVC'^Oii 

0.83 

4.55 

"  Slices  incubated  1  hr  with  carboxyl-labeled  fatty  acids  shown  in  Table  1-24  at 
38°  and  pH  7.1.  The  data  from  the  five  fatty  acids  were  averaged  to  indicate  the 
general  effects  of  malonate.  (From  Geyer  et  al.,  1950  b.) 


CO—  fragments  are  believed  to  enter  the  cycle.  The  CHgC^*  o_/— CHaC^^ 
00"  ratio  in  acetoacetate  will  depend  on  the  rates  of  production  and  utili- 
zation of  these  two  fragments.  As  pointed  out,  these  two  types  of  2-carbon 
fragment  may  be  only  convenient  designations  for  two  reactive  forms  of 
acetyl-CoA.  Malonate  is  assumed  to  increase  the  formation  of  acetoacetate 
by  condensation  of  randomized  fragments  of  the  — CHaC^^O —  type,  so 
that  the  ratio  rises.  The  extra  acetoacetate  formed  over  the  control  when 
malonate  is  present  does  indeed  exhibit  a  ratio  of  unity  for  hexanoate  and 
octanoate  oxidation  (Geyer  et  al.,  1950  b).  If  malonate  does  this  by  blocking 
the  cycle,  a  preferential  accumulation  of  — CH2CO —  fragments  would  occur, 
and  a  greater  proportion  of  the  acetoacetate  would  be  formed  from  them. 
The  effect  of  malonate  on  acetoacetate  accumulation  will  depend  on  the 
pathway  of  fatty  acid  oxidation  in  the  uninhibited  tissue  and,  hence,  on  the 
experimental  conditions.  For  example,  Witter  et  al.  (1950)  found  that  3  mM 
malonate  inhibits  acetoacetate  formation  from  hexanoate  4%  and  10  mM 
malonate  inhibits  9%  in  suspensions  of  washed  particles  from  rat  liver. 
However,  hexanoate  is  quantitatively  converted  to  acetoacetate  in  the  con- 
trols, presumably  because  no  cycle  intermediates,  are  present  to  form  oxal- 
acetate  for  condensation  of  the  acetyl-CoA  units.  Under  such  circumstances 
malonate  would  not  be  expected  to  increase  acetoacetate  and  the  small 
inhibitions  must  be  attributed  to  actions  directly  on  the  helix.  The  rela- 
tionship between  acetoacetate  formation  in  malonate-inhibited  systems 
and  the  presence  of  cycle  intermediates  was  illustrated  and  discussed  by 
Cheldelin  and  Beinert  (1952). 


144  1.    MALONATE 

We  have  seen  that  the  accumulation  of  acetoacetate  in  the  presence  of 
malonate  can  be  attributed  to  an  increased  rate  of  formation  of  the  aceto- 
acetate. Is  the  accumulation  due  entirely  to  this  or  can  malonate  also  inhibit 
the  utilization  of  acetoacetate  in  some  tissues?  The  rise  in  the  acetoacetate 
in  liver  slices  in  the  presence  of  malonate  was  believed  by  Jowett  and 
Quastel  (1935  c)  to  be  due  to  the  inhibition  of  the  decomposition  of  aceto- 
acetate, since  at  that  time  the  pathways  for  the  formation  of  acetoacetate 
were  not  understood.  However,  Quastel  and  Wheatley  (1935)  soon  provided 
evidence  that  malonate  can  interfere  with  the  disappearance  of  acetoacetate 
in  rat  liver  and  kidney  slices.  In  kidney  slices,  malonate  at  8  mM  inhibits 
around  42%  and  at  16  mM  64%,  and  in  liver  slices  an  inhibition  of  74% 
was  observed  with  40  mM  malonate.  Fumarate  is  able  to  counteract  this 
inhibition  partially  and  it  was  concluded  that  acetoacetate  oxidation  must 
be  coupled  with  other  oxidations  inhibited  by  malonate.  Very  similar  results 
were  reported  by  Edson  and  Leloir  (1936);  indeed,  20  mM  malonate  inhibits 
disappearance  of  acetoacetate  in  rat  kidney  slices  93%  and  it  was  stated, 
"Malonate  is  a  powerful  and  relatively  specific  inhibitor  of  respiration  and 
of  aerobic  disappearance  of  acetoacetic  acid  in  kidney."  Both  Handler 
(1945)  and  Mookerjea  and  Sadhu  (1955)  in  their  work  with  whole  animals, 
favored  the  concept  that  malonate  interfered  with  acetoacetate  metabolism 
accounting  for  the  rises  in  blood  acetoacetate.  Inasmuch  as  several  different 
pathways  are  open  to  acetoacetate  and  these  vary  with  the  tissue  used,  it  is 
difficult  to  interpret  accurately  these  results.  In  some  tissues,  acetoacetate 
can  be  split  into  acetyl-CoA  fragments  that  enter  the  cycle  and  here  malo- 
nate might  inhibit  by  blocking  the  cycle  and  the  formation  of  oxalacetate, 
which  is,  of  course,  esentially  the  same  mechanism  adduced  to  explain  the 
increased  formation  of  acetoacetate.  There  is  evidence  that  malonate  does 
not  inhibit  the  reduction  of  acetoacetate  to  /?-hydroxybutyrate  (Edson  and 
Leloir,  1936),  nor  does  it  seem  to  interfere  with  the  formation  of  sterols  from 
acetoacetate  (Mookerjea  and  Sadhu,  1955).  It  is  probably  best  in  the 
present  state  of  our  knowledge  to  attribute  the  accumulation  of  acetoacetate 
in  the  presence  of  malonate  primarily  to  a  diversion  of  2-carbon  units  away 
from  oxidation  through  the  cycle,  without  eliminating  the  possibility  that 
malonate  may  interfere  in  other  pathways  for  the  utilization  of  acetoacetate. 

Effects  on  Propionate  Metabolism 

Propionate  arises  terminally  from  the  /^-oxidation  of  odd-numbered  fatty 
acids  and  in  certain  tissues,  such  as  the  liver,  can  be  oxidized  completely 
through  the  cycle.  However,  the  oxidation  of  propionate  differs  from  that 
of  other  fatty  acids.  The  following  sequence  of  reactions  has  been  suggested: 

CO, 

+ 

.  ATP  ATP  B,j 

Propionate      — >      propionyl-CoA      — >      methylmalonyl-CoA      — >      succinyl-CoA 


EFFECTS    OF   MALONATE   ON   LIPID   METABOLISM  145 

The  over-all  reaction  is  the  carboxylation  of  propionate  to  succinate.  Other 
pathways  occur  in  bacteria,  e.g. 

Propionate  ->  propionyl-CoA  ->  acrylyl-CoA  ->  lactyl-CoA  ->■  pyruvate  ->  COg  +  H2O 

Such  a  sequence  may  also  operate  in  animal  tissues,  since  lactate  was  iden- 
tified chromatographically  after  incubation  of  mouse  liver  slices  with  pro- 
pionate (Daus  et  al.,  1952).  If  the  principal  pathway  of  propionate  is  via 
succinate,  malonate  would  be  expected  to  inhibit  its  oxidation  readily  but, 
if  acetyl-CoA  is  formed,  the  inhibition  will  vary  with  the  conditions  as 
discussed  for  the  effects  of  malonate  on  pyruvate  oxidation. 

Malonate  has  been  shown  to  inhibit  strongly  the  C^^Oo  formation  from 
labeled  propionate  in  mouse  liver  slices  (Daus  et  al.,  1952),  rat  liver  slices 
(Katz  and  Chaikoflf,  1955),  suspensions  of  rabbit  liver  particles  (Wolfe, 
1955),  and  peanut  mitochondria  (Giovanelli  and  Stumpf,  1958),  as  anticipat- 
ed. In  the  rabbit  liver  particulate  preparation,  10  mM  malonate  suppresses 
the  formation  of  C^^O,  from  both  propionate-l-C^*  and  propionate-2-C^^ 
almost  completely,  and  at  the  same  time  leads  to  the  accumulation  of  suc- 
cinate, and  in  rat  liver  slices  malonate  also  causes  succinate  accumulation. 
From  these  data  alone  it  is  impossible  to  say  whether  the  succinate  arises 
directly  from  propionate  or  is  formed  via  the  cycle,  but  the  marked  inhibi- 
tion of  C^^O,  formation  would  indicate  the  former.  This  is  substantiated 
by  the  demonstration  of  labeled  methylmalonate  in  the  rat  liver  slices. 
Another  possible  site  for  malonate  inhibition  is  suggested  by  the  work  of 
Flavin  et  al.  (1955)  on  rat  tissues.  The  intercon version  of  methylmalonate 
and  succinate  was  shown  to  be  inhibited  completely  by  5  mM  malonate  and 
thus  malonate  leads  to  the  accumulation  of  methylmalonate  during  propion- 
ate metabolism.  However,  it  is  not  known  if  malonate  can  inhibit  the 
methylmalonyl-CoA  isomerase,  which  catah^zes  the  interconversion  in  the 
normal  pathway,  or  if  malonate  only  inhibits  the  formation  of  methyl- 
malonyl-CoA  from  methylmalonate.  The  latter  is  reasonable  because  mal- 
onate could  compete  with  methylmalonate  for  the  active  site  on  the  enzyme. 

In  peanut  mitochondria  the  situation  may  well  be  different.  Malonate  at 
6  mM  inhibits  the  formation  of  C^^Oa  from  propionate- 1-C^*  41%  (Giovanelli 
and  Stumpf,  1958).  It  was  felt  that  this  inhibition  is  not  as  much  as  would 
be  expected  if  the  pathway  from  propionate  leads  to  succinate.  Further- 
more, fluoride,  which  inhibits  the  carboxylation  of  propionyl-CoA,  does 
not  depress  the  C^^Oa  significantly.  The  pathway  through  methylmalonyl- 
CoA  to  succinate  may  not  be  operative  here,  and  the  following  pathway  was 
proposed: 

Propionate    ->•    propionyl-CoA    ->   acrylyl-CoA    ->   ^-hydroxypropionyl-CoA    ->• 
/?-hydroxypropionate   ->   malonic  semialdehyde   ->   malonyl-CoA  ->   acetyl-CoA 

The  COo  formed  in  the  last  step  derives  from  the  carboxyl  group  of  propion- 


146  1.    MALONATE 

ate  so  that  an  inhibition  of  C^^02  formation  from  proprionate-l-C^*  would 
imply  an  action  of  malonate  somewhere  along  this  pathway.  It  is  possible 
that  malonate,  or  malonyl-CoA  formed  from  it,  could  compete  with  the 
malonyl-CoA  from  propionate  and  in  this  way  reduces  the  formation  of 
0^*02-  It  is  known  that  the  addition  of  methylmalonate  simultaneously 
with  propionate  depresses  the  propionate  utilization  strongly  (Feller  and 
Feist,  1957).  Lactating  rat  mammary  gland  preparations  convert  propion- 
ate to  fatty  acids  in  part,  the  principal  pathway  being  direct  condensation 
with  malonyl-CoA  to  form  the  odd-chain  fatty  acids.  Malonate  at  10  vnM 
inhibits  this  incorporation  50-65%,  whatever  the  position  of  C^^  in  pro- 
pionate, and  simultaneously  C^^Og  is  depressed  around  30%  from  proprion- 
ate-l-C^*  and  propionate-2-C^^,  and  nearly  50%  from  priopionate-3-C^^ 
(Cady  et  al.,  1963).  This  was  interpreted  not  as  a  direct  action  on  the  pro- 
pionate pathway  but  as  a  reduction  of  ATP  or  NAD(P)H,  these  being  nec- 
essary for  fatty  acid  synthesis,  as  a  consequence  of  inhibition  of  the  cycle. 
In  Rhodospirillum  rubrum,  both  the  oxidation  (Clayton  et  al.,  1957) 
and  the  photosynthetic  dissimilation  (Clayton,  1957)  of  propionate  are 
inhibited  by  malonate  to  approximately  the  same  extent  as  are  the  similar 
reactions  of  succinate,  and  this  was  given  as  evidence  to  support  the  meta- 
bolism of  propionate  to  succinate  in  these  organisms. 

Effects  on  Synthesis  of  Fatty  Acids 

There  are  at  least  three  systems  for  the  synthesis  of  fatty  acids;  one  is  the 
reversal  of  the  /^-oxidation  in  the  helix  and  the  other  two  involve  the  for- 
mation of  malonyl-CoA  from  acetyl-CoA,  one  mitochondrial  and  the  other 
nonmitochondrial  (Green  and  Wakil,  1960).  There  are  obvious  relationships 
between  fatty  acid  synthesis  and  oxidative  metabolism  of  various  sub- 
strates. The  controls  that  establish  the  rates  of  fatty  acid  synthesis,  or  the 
balance  between  oxidation  and  synthesis,  have  not  been  elucidated  and  it  is 
difficult  to  determine  in  a  particular  case  what  the  effect  of  a  cycle  block 
would  probably  be.  The  level  of  acetyl-CoA  and  the  availability  of  the  var- 
ious pathways  for  its  metabolism  must  be  an  important  factor,  but  the 
concentrations  of  coenzyme  A,  ATP,  NADH  and  NADPH  could  also  play 
a  significant  role. 

Malonate  has  been  found  to  produce  a  variety  of  effects.  Most  of  the 
studies  have  involved  the  incubation  of  tissue  preparations  with  acetate- 
1-C^*  and  the  subsequent  determination  of  labeled  fatty  acids  formed  from 
the  acetate.  In  some  cases  a  marked  stimulation  of  fatty  acid  synthesis 
in  the  presence  of  malonate  has  been  observed.  Malonate  at  50  milf  inhib- 
its the  O2  uptake  of  nonparticulate  extracts  of  rat  mammary  gland  and 
yet  increases  the  formation  of  fatty  acids,  sometimes  as  much  as  10-fold 
(Popjak  and  Tietz,  1955),  The  addition  of  oxalacetate  or  a-ketoglutarate 
with  the  malonate  increases  the  synthesis  aven  more  (see  accompanying 


EFFECTS   OF  MALONATE   ON  LIPID  METABOLISM  147 

tabulation).  The  stimulating  action  of  a-ketoglutarate  was  attributed  to 
the  generation  of  NADH  by  which  hydrogen  atoms  are  provided  for  fatty 
acid  synthesis.  Dils  and  Popjak  (1962)  claimed  that  malonyl-CoA  is  not 


Incorporation  of  acetate  into 
fatty  acids  (m/<  moles/ 100  mg) 


Control  18.2 

Malonate  118 

Oxalacetate  72 . 7 

Malonate  +  oxalacetate  241 

a-Ketoglutarate  51.7 

Malonate  +  a-ketoglutarate  517 

formed  from  malonate  in  these  extracts  and  that  the  stimulation  of  fatty 
acid  synthesis  must  be  an  indirect  effect,  possibly  the  suppression  of  the 
deacylation  of  malonyl-CoA  formed  from  acetyl-CoA,  or  the  inliibition  of 
the  decarboxylation  of  malonyl-CoA.  Kallen  and  Lowenstein  (1962)  pointed 
out  that  if  this  were  the  mechanism  by  which  malonate  acts,  it  should  also 
stimulate  the  synthesis  of  fatty  acids  from  malonyl-CoA,  which  it  does  not; 
indeed,  malonate  at  10  niM  inhibits  the  conversion  of  malonyl-CoA  into 
fatty  acids  33%.  There  is  actually  a  stimulation  of  the  formation  of  fatty 
acids  from  acetyl-CoA.  Furthermore,  Spencer  and  Lowenstein  (1962)  found 
that  malonate  is  incorporated  into  fatty  acids  in  an  extramitochondrial 
extract  from  rat  mammary  gland;  acetate  stimulates  malonate  incorpora- 
tion just  as  malonate  stimulates  acetate  incorporation.  All  of  the  stimula- 
tion by  malonate,  however,  cannot  be  explained  by  its  conversion  to  mal- 
onyl-CoA  since  several  times  more  acetate  than  malonate  is  incorporated. 
The  varying  effects  of  malonate  on  different  preparations  from  a  single 
tissue  are  well  illustrated  in  the  studies  of  Abraham  et  al.  (1961)  with  rat 
mammary  gland,  where  malonate  stimulates  fatty  acid  synthesis  markedly 
in  cell-free  systems  (maximal  stimulation  around  130%  at  17  mM  malon- 
ate), inhibits  the  synthesis  63%  in  slices,  and  has  very  little  affect  when 
glucose  is  present.  Glucose  was  assumed  to  provide  NAD(P)H  by  forming 
cycle  substrates  and  also  to  augment  the  ATP  level,  which  in  the  absence 
of  glucose  might  have  been  reduced  by  malonate.  In  the  homogenates  ATP 
was  added  so  that  this  aspect  of  the  action  of  malonate  could  not  be  exhibit- 
ed. The  response  to  malonate  is  markedly  dependent  on  the  experimental 
conditions,  as  shown  by  Hosoya  and  Kawada  (1961)  with  human  placental 
slices,  additions  of  estradiol,  ATP,  NAD,  or  bicarbonate  altering  the  fatty 
acid  synthesis  and  its  modification  by  malonate.  It  may  be  noted  that  fatty 
acid  synthesis  in  particulate  preparations  from  the  locust  fat  body  occurs 
rapidly  only  in  the  presence  of  malonate  (Tietz,  1961). 


148  1.    MALONATE 

So  far  we  have  considered  total  fatty  acid  synthesis.  Separation  of  the 
different  fatty  acids  from  animal  tissues  in  malonate  experiments  has  not 
been  done,  but  in  mycobacteria  malonate  shifts  the  incorporation  of  acetate 
into  the  higher  fatty  acids  (Kusunose  et  al.,  1960).  The  synthesis  of  total 
fatty  acid  is  moderately  increased  and  this  was  attributed  to  the  formation 
of  malonyl-CoA  (see  accompanying  tabulation).  Differential  effects  of  mal- 

Acetate-l-C*    incorporation 

Fatty  acid  %  Change 

Control  Malonate    3.3    mM 

Palmitate  7627 

Stearate  2887 

Arachidate  1144 

Behenate  1089 

Lignocerate  1862 

Total  acids  14609  17234  +   18 


onate  on  the  synthesis  of  short-chain  and  long-chain  fatty  acids  are  also 
seen  in  rat  liver  slices  metabolizing  octanoate-1-C^*  (Lyon  and  Geyer,  1954). 
Although  the  over  all  effect  of  malonate  on  lipid  synthesis  in  a  particulate 
preparation  from  avocado  mesocarp  is  not  marked,  the  incorporation  of 
acetate  is  shifted  from  stearate  to  oleate  (see  accompanying  tabulation) 


3488 

-  54 

1497 

-  48 

1332 

+  16 

2110 

+  94 

8807 

-f373 

Malonate 

Acetate  incorporation 
into  lipid 

Distribution  of  label 

(mM) 

Palmitate 

Stearate 

Oleate 

0 

8.20 

26 

33 

41 

5 

8.50 

23 

12 

65 

10 

8.15 

24 

14 

62 

30 

7.10 

21 

9 

70 

(Mudd  and  Stumpf,  1961).  Although  malonate  may  inhibit  the  cycle,  this 
may  be  counteracted  by  the  formation  of  malonyl-CoA,  which  dilutes  the 
labeled  malonyl-CoA  formed  from  labeled  acetate.  It  is  interesting  that 
malonate  is  formed  from  acetate  in  avocado  and  this  may  be  one  regulatory 
factor  in  fatty  acid  synthesis.  Malonate  has  been  found  in  three  instances 
to  exert  only  inhibitory  effects  on  fatty  acid  synthesis:  in  cell-free  prepara- 
tions from  pigeon  liver,  20  mM  malonate  inhibits  the  incorporation  of 
acetate  into  fatty  acids  32%  (Brady  and  Gurin,  1952);  in  various  tumor 
tissues  (mammary  and  testicular  carcinomata,  and  a  sarcomatoid  ovarian 


EFFECTS   OF  MALONATE   ON  LIPID  METABOLISM  149 

tumor),  30  mM  malonate  inhibits  8-73%  (van  Vals  and  Emmelot,  1957); 
and  in  rat  liver  extracts  containing  mitochondria,  10  mM  malonate  in- 
hibits 51%  (Iliffe  and  Myant,  1964). 

These  divergent  observations  are  difficult  to  explain  satisfactorily  and 
one  must  conclude  that  the  final  effects  of  malonate  must  depend  on  many 
factors.  In  cellular  preparations  malonate  may  alter  the  levels  of  ATP, 
NAD(P)H,  and  coenzyme  A,  as  well  as  divert  the  metabolism  of  acetyl- 
CoA  by  its  inhibition  of  the  cycle.  In  nonmitochondrial  soluble  enzyme 
systems  these  actions  would  be  minimized  or  absent,  and  the  most  important 
factors  might  be  the  facilitation  of  fatty  acid  synthesis  through  the  forma- 
tion of  malonyl-CoA  or  direct  effects  on  the  enzymes  involved,  although 
there  is  no  evidence  for  such  direct  affects  at  present.  When  malonate  is 
itself  incorporated  into  fatty  acids,  as  in  several  examples  above  and  in 
spinach  chloroplasts,  where  malonate  incorporation  occurs  at  about  half 
the  rate  for  acetate  (Mudd  and  McManus,  1964),  additional  complications 
must  be  considered.  Since  the  incorporation  of  acetate- 1-C^*  into  lipid  in 
chloroplasts  is  reduced  71%  by  0.67  mM  malonate  (Mudd  and  McManus, 
1962),  it  would  appear  that  malonate  also  inhibits  some  step  or  steps  in  the 
pathway.  The  compartmentalization  of  the  pools  of  acetyl-CoA,  malonyl- 
CoA,  acetoacetate,  and  the  various  enzymes  and  cofactors  within  the  cell 
must  be  borne  in  mind  in  trying  to  explain  certain  differential  affects  of 
malonate. 

Effects  on  the  Metabolism  of  Fats,  Phospholipids,  and  Sterols 

Several  observations  on  total  lipid  response  to  malonate  are  interesting 
even  though  it  is  impossible  to  assign  a  mechanism.  In  rat  liver  homogenates 
incubated  with  palmitate-1-C^*,  the  lipids  other  than  phosphohpids  increase 
22%  in  the  presence  of  10  mM  malonate  compared  to  controls  (Jedeikin 
and  Weinhouse,  1954).  Whether  this  is  direct  utilization  of  palmitate  or 
lipid  synthesis  with  the  C^^Og  formed  from  palmitate  is  difficult  to  say. 
Malonate  at  50  mM  also  increases  the  total  lipid  content  of  potato  tuber 
slices  some  230%  (Table  1-19)  (Romberger  and  Norton,  1961)  and  this 
could  be  due  mainly  to  an  increased  synthesis  of  fatty  acids.  On  the  other 
hand,  lipid  synthesis  from  glucose-C^*  in  human  placenta  is  depressed  75% 
by  20  mM  malonate  (Hosoya  et  al,  1960).  These  results  again  show  that 
the  action  of  malonate  on  lipid  metabolism  is  quite  variable.  It  will  be  more 
profitable  to  turn  to  the  synthesis  of  particular  lipid  fractions. 

Injections  of  malonate  lead  to  elevation  of  the  free  and  esterified  cho- 
lesterol in  the  liver,  kidney,  and  blood  of  the  rat  (see  tabulation)  (Mook- 
erjea  and  Sadhu,  1955).  Injections  of  800  mg/kg  of  sodium  malonate  were 
made  daily  for  3-4  weeks,  some  toxic  effects  being  noted,  and  the  animals 
then  sacrificed.  Simultaneously,  the  blood  glucose  rose  from  92  mg%  to 
196  mg%  and  the  blood  acetoacetate  from  0.8  mg%  to  3.6  mg%.  Kidney 


150 


1.    MALONATE 


Free  cholesterol 
(mg/100  g  wet  wt.) 

Esterified  cholesterol 
(mg/100  g  wet  wt.) 

Contro: 

Malonate 

%  Change 

Control 

Malonate     %  Change 

Liver 

Kidney 

Blood 

205 

360 

43 

426 
724 

83 

+  107 
+  101 
+  93 

73 

72 
68 

104               +  42 

228               +216 

80               +   17 

and  liver  slices  showed  impaired  respiration  with  succinate,  acetate,  and 
acetoacetate  as  substrates.  This  augmentation  of  tissue  cholesterol  is  clear 
and  is  reasonable  on  the  basis  of  diversion  of  acetyl-CoA  metabolism  by  a 
block  of  succinate  oxidase.  However,  in  vitro  work  has  shown  only  inhibi- 
tion of  cholesterol  synthesis.  The  formation  of  labeled  cholesterol  from 
octanoate-1-C^*  in  rat  liver  slices  is  consistently  depressed  by  5.84  mM 
malonate,  and  fumarate  was  very  ineffective  in  counteracting  this  inhibition 
(Lyon  and  Geyer,  1954).  The  total  lipids  rise  and  this  is  partly  attributable 
to  the  increased  synthesis  of  short-chain  fatty  acids.  The  formation  of 
labeled  cholesterol  from  acetate- l-C^*  in  the  same  tissue  is  inhibited  73% 
by  50  mM  malonate  (Kline  and  DeLuca,  1956)  and  78%  by  30  mM  mal- 
onate (van  Vals  and  Emmelot,  1957).  Cholesterol  synthesis  in  rat  tumors  is 
even  more  strongly  depressed.  The  discrepancy  between  the  in  vivo  and  in 
vitro  results  might  be  due  to  several  factors.  In  the  intact  animal  many 
secondary  effects  may  occur,  e.g.  as  a  result  of  the  marked  rise  in  blood 
glucose.  Also  the  malonate  concentration  in  the  tissues  of  the  rats  is  undoubt- 
edly less  than  in  the  work  with  slices.  It  is  unfortunate  that  most  of  the 
studies  have  been  made  with  unreasonably  high  malonate  concentrations 
so  that  a  specific  inhibition  of  succinate  oxidation  is  doubtful.  The  catabolism 
of  cholesterol,  as  determined  by  the  formation  of  C^'^Og  from  the  labeled 
terminal  methyl  groups  of  cholesterol,  in  suspensions  of  rat  liver  mitochon- 
dria is  inhibited  78%  by  10  mM  malonate  (Whitehouse  et  al.,  1959),  so 
that  this  factor  must  also  be  considered  in  explaining  changes  in  tissue  levels 
over  longer  periods  of  time.  The  synthesis  of  other  sterols  has  been  studied 
very  little.  Pieces  of  rat  adrenal  form  corticosteroids  in  the  presence  of 
glucose  and  this  is  markedly  stimulated  by  the  addition  of  ACTH.  Malonate 
at  10  mM  stimulates  the  formation  of  sterols  in  the  absence  of  ACTH 
from  17  to  22  //g/100  mg/2  hr  (  +  29%)  but  depresses  the  synthesis  in  the 
ACTH-activated  preparations  from  81  to  72  //g/100  mg/2  hr  (-11%) 
(Schonbaum  et  al.,  1956).  Fluoroacetate  also  inhibits  very  little  and  it  was 
concluded  that  the  cycle  does  not  play  a  major  role  in  sterol  synthesis, 
glucose  metabolism  and  particularly  the  pentose  phosphate  pathway  being 
of  more  importance.  The  bearing  of  such  studies  on  the  metabolic  basis  of 
cholesterol  and  hormonal  sterol  levels  in  animals,  especially  the  relationship 


EFFECTS  ON  AMINO  ACID  AND  PROTEIN  METABOLISM  151 

to  the  activity  of  the  cycle  and  the  other  pathways  for  the  utilization  of 
acetyl-CoA,  warrants  further  investigations  of  the  actions  of  malonate  and 
other  cycle  inhibitors  both  in  vitro  and  in  vivo.  One  approach  to  the  met- 
abolic defect  in  hypercholesteremia  could  be  made  in  this  way. 

The  incorporation  of  inorganic  P^^  into  phospholipids  is  almost  invari- 
ably inhibited  strongly  by  malonate.  This  has  been  shown  in  peanut  mito- 
chondria (Mazelis  and  Stumpf,  1955),  mycobacteria  (Tanaka,  1960),  guinea 
pig  brain  dispersions  (R.  M.  C.  Dawson,  1953),  rat  liver  mitochondria 
(Marinetti  et  al.,  1957),  and  other  tissues.  In  cat  brain  slices,  the  effects  of 
malonate  are  very  slight  and  it  is  possible  that  malonate  does  not  penetrate 
well  (Strickland,  1954).  Yet  3  mM  malonate  inhibits  such  incorporation 
87%  in  K+-stimulated  rat  brain  slices,  although  this  may  be  due  to  a  more 
active  cycle  participation  in  the  active  tissue,  inasmuch  as  respiration  is 
93%  inhibited  (Yoshida  and  Quastel,  1962).  The  phosphorylation  of  phos- 
pholipid precursors  probably  involves  the  formation  of  high-energy  phos- 
phate compounds  and  malonate  could  depress  this  as  the  result  of  a  block  of 
the  cycle.  A  direct  effect  on  the  phosphorylation  is  unlikely.  On  the  other 
hand,  the  incorporation  of  activity  into  phospholipids  from  palmitate-1-C^^ 
in  rat  liver  homogenates  (Jedeikin  and  "Weinhouse,  1954)  or  from  acetate- 
1-C^*  in  rat  liver  slices  (Kline  and  DeLuca,  1956)  is  affected  scarcely  at  aU  by 
malonate.  The  phospholipids  comprise  a  very  heterogenous  group  and  the 
response  to  malonate  probably  depends  on  which  type  of  phospholipid  is 
under  investigation. 

EFFECTS  OF  MALONATE  ON  AMINO  ACID 
AND  PROTEIN    METABOLISM 

The  pathways  of  amino  acid  metabolism  often  lead  to  or  from  the  cycle 
so  that  malonate  would  be  expected  to  influence  amino  acid  utilization  and 
formation  by  its  inhibition  of  succinate  oxidase.  The  intracellular  accumu- 
lation of  amino  acids  and  their  incorporation  into  proteins  are  processes 
requiring  energy  and  consequently  malonate  could  depress  these  important 
reactions  involved  in  cellular  growth  by  a  depletion  of  high-energy  phosphate 
derived  from  the  cycle.  Finally,  malonate  might  act  directly  on  the  enzjones 
catalyzing  amino  acid  transformations.  Information  on  these  matters  is 
fragmentary  but  enough  work  has  been  done  to  demonstrate  some  interest- 
ing effects  of  malonate  on  this  phase  of  metabolism. 

Effects   on   Amino   Acid    Metabolism 

None  of  the  enzymes  involved  directly  in  amino  acid  metabolism  seems 
to  be  very  sensitive  to  malonate  (Table  1-12)  but  a  number  of  important 
reactions  have  never  been  studied.  Enz>Tnes  catalyzing  the  reactions  of 


152  1.    MALONATE 

the  dicarboxylic  amino  acids  particularly  might  be  expected  to  bind  mal- 
onate  to  some  extent,  but  there  is  only  indirect  evidence  for  this.  The 
dehydrogenation  of  glutamate  with  methylene  blue  as  an  acceptor  in 
toluene-treated  E.  coli  is  inhibited  about  20%  by  71  vaM  malonate  (Quastel 
and  Wooldridge,  1928),  but  it  may  be  that  all  of  the  hydrogen  atoms  do 
not  arise  from  glutamate  here  and  that  some  other  reaction  is  inhibited. 
In  Walker  carcinosarcoma,  kidney,  and  liver,  glutamate  is  metabolized 
readily  to  succinate  in  the  presence  of  6.3  milf  malonate  (Nyhan  and 
Busch,  1957),  but  no  controls  are  avilable  so  that  some  inhibition  is  pos- 
sible. Aspartate  and  glutamate  are  metabolized  by  Hemophilus  parain- 
fluenzae;  malonate  does  not  interfere  with  the  oxidation  of  the  former  but 
does  inhibit  glutamate  oxidation  (Klein,  1940).  Contrary  to  these  results, 
malonate  completely  inhibits  aspartate  oxidation  in  rat  liver  homogenate 
(Nakada  and  Weinhouse,  1950).  It  was  believed  rightly  that  this  could  not 
be  entirely  attributed  to  an  inhibition  of  succinate  oxidase. 

Glutamate  may  be  converted  to  a-ketoglutarate  by  either  glutamate 
dehydrogenase  or  transamination,  or  it  may  be  decarboxylated  to  y-amino- 
butyrate;  the  decarboxylase  is  limited  mainly  to  certain  bacteria  and  the 
nervous  system  of  animals,  so  the  major  product  is  usually  a-ketoglutarate, 
which  can  be  oxidized  through  the  cycle  or  participate  in  transaminations 
whereby  it  is  reconverted  to  glutamate  (this  occurs  also  with  y-amino- 
butyrate  so  that  the  net  reaction  forms  succinic  semialdehyde,  ammonia, 
and  CO2  from  glutamate).  The  pattern  of  glutamate  metabolism  will  depend 
on  the  relative  activities  of  these  various  enzymes,  the  availability  of  other 
amino  acids  for  transamination,  and  the  supply  of  NAD  for  the  glutamate 
and  a-ketoglutarate  dehydrogenases;  likewise,  the  response  to  malonate 
inhibition  will  depend  on  these  factors.  If  malonate  selectively  inhibits 
succinate  oxidation,  the  O2  uptake  due  to  glutamate  should  be  reduced 
moderately  (perhaps  around  25-50%)  unless  much  of  the  a-ketoglutarate 
formed  is  transaminated  and  does  not  enter  the  cycle.  Malonate,  however, 
occasionally  inhibits  the  formation  of  ammonia  from  glutamate,  indicating 
some  effect  on  the  oxidative  deamination.  Malonate  also  inhibits  the  oxi- 
dation of  glutamate  by  guinea  pig  mammary  gland  mitochondria  completely 
(Jones  and  Gutfreund,  1961),  which  would  not  be  the  case  if  only  succinate 
oxidation  were  blocked.  The  glutamate  respiration  of  rat  brain  mitochon- 
dria is  depressed  88%  by  17.3  mM  malonate  (see  note  in  Table  1-14) 
(Levtrup  and  Svennerholm,  1963),  which  would  indicate  that  glutamate 
is  being  converted  mainly  to  a-ketoglutarate  by  transamination  (glutamate 
decarboxylase  is  not  present  in  brain  mitochondria).  Similar  high  inhibi- 
tions by  20  mM  malonate  are  observed  in  the  mitochondria  from  pigeon 
muscle,  rat  heart,  rat  liver,  and  ascites  cells  (64-99%)  (Borst,  1962).  Con- 
clusions as  to  the  pathway  of  glutamate  catabolism  based  on  the  results 
with  malonate  depend  on  the  assumption  that  the  inhibition  is  specifically 


EFFECTS  ON  AMINO  ACID  AND  PROTEIN  METABOLISM  153 

on  succinate  oxidation,  and  at  these  high  concentrations  this  may  not  be 
true.  On  the  other  hand,  Das  and  Roy  (1961,  1962)  claim  that  transamina- 
tion contributes  little  to  the  metabolism  of  glutamate  in  mitochondria  from 
Vigna  sinensis,  and  since  the  decarboxylase  is  absent,  oxidation  by  gluta- 
mate dehydrogenase  would  seem  to  be  the  major  route.  Glutamate  is  con- 
verted primarily  to  aspartate  in  rat  brain  homogenate  via  the  pathway 
glutamate  -^  a-ketoglutarate  -^  succinate  -^  oxalacetate  -^  aspartate 
(Haslam  and  Krebs,  1963).  The  addition  of  fumarate  removes  this  inhi- 
bition, as  expected. 

Certain  amino  acids  appear  to  be  involved  in  the  functioning  of  nerve 
tissue  and  the  effects  of  inhibitors  on  the  metabolism  of  these  substances 
are  of  particular  interest  in  this  connection.  Glutamate  is  accumulated  in 
brain  and  plays  a  role  in  the  active  transport  of  ions,  while  y-aminobutyrate 
and  A^-acetylaspartate  have  recently  attracted  attention  because  of  their 
ability  to  modify  central  nervous  system  activity.  Glutamate  and  K"*"  are 
taken  up  by  retina  and  brain  slices  in  approximately  equivalent  amounts. 
Malonate  at  20  raM  depresses  the  formation  of  glutamate  +  glutamine 
only  12%  while  it  reduces  K+  uptake  40%  (Terner  et  al.,  1950),  indicating 
that  the  major  effect  of  malonate  is  not  mediated  through  interference  with 
glutamate.  When  guinea  pig  brain  slices  are  incubated  with  glucose-u-C^*, 
a  good  deal  of  the  C^'*  appears  in  amino  acids,  the  most  important  of  which 
is  glutamate  (Tsukada  et  al.,  1958).  Malonate  at  10  mTlf  inhibits  glutamate 
formation  around  25%,  ^-aminobutyrate  formation  around  75%,  and  the 
formation  of  aspartate  appreciably.  The  total  C^*  incorporation  into  amino 
acids  from  glucose-u-C^*  in  rat  brain  slices  is  inhibited  64%  by  10  mM 
malonate  at  normal  K+  concentration  but  83%  in  the  presence  of  105 
mM  K+,  which  produces  an  activation  of  brain  metabolism  (Kini  and 
Quastel,  1959).  Such  results  can  be  readily  explained  on  the  basis  of  a  mal- 
onate-reduced  pool  of  amino  acid  precursors  due  to  the  reduction  in  cycle 
activity.  Glutamate  is  a  central  substance  in  amino  acid  formation  through 
transaminations  and  anything  which  decreases  the  formation  of  a-keto- 
glutarate would  be  expected  to  impair  these  pathways.  Cremer  (1964)  has 
recently  found  that  40  mM  malonate  not  only  reduces  drastically  the 
incorporation  of  glucose-u-C^*  into  glutamate,  aspartate,  y-aminobutyrate, 
and  protein  in  brain  slices,  but  also  causes  a  loss  of  amino  acids  from  the 
cells.  This  concentration,  of  course,  is  probably  not  specifically  inhibiting 
succinate  oxidation.  A  disputation  type  of  reaction  occurs  in  certain 
tissues: 

2  a-Ketoglutarate  +  NH3  +  ADP  +  P,    ^   glutamate  +  succinate  +  CO,  +  ATP 

Tager  (1963)  used  malonate  to  block  succinate  dehydrogenase  and  surpris- 
ingly found  that  it  augments  the  formation  of  glutamate  in  suspensions 
of  rat  liver  mitochondria  (see  accompanying  tabulation).  It  was  suggested 


154 


1.    MALONATE 


Control 


Malonate  20  mM 


A  O2  (/^  atoms) 

-1.2 

-   1.2 

A  a-Ketoglutarate  (^  moles) 

-3.9 

-10.4 

A  Glutamate  (/<  moles) 

+2.0 

+  4.9 

A  Esterified  phosphate  (/<  moles) 

+  1.9 

+  4.8 

that  malonate  is  converted  to  oxalosuccinate  via  malonyl-CoA.  The  oxalo- 
succinate  might  function  in  the  NAD-  and  NADP-dependent  isocitrate 
dehydrogenase  systems  to  form  a  transhydrogenase  so  that  NADPH  is 
the  eventual  donor  for  the  formation  of  glutamate,  oxalosuccinate  acting 
catalytically.  Such  studies  show  how  complex  the  effects  of  malonate  on 
amino  acid  metabolism  can  be.  The  oxidation  of  certain  amino  acids  pro- 
ceeds via  an  initial  transamination  followed  by  degradation  of  the  deaminat- 
ed  acids.  The  oxidation  of  y-aminobutyrate  is  completely  inhibited  by 
1  n\M  malonate  in  rat  brain  mitochondria  (Sacktor  et  al.,  1959)  but  in 
Bacillus  pvmilus  is  not  affected  even  by  40  mM  malonate  (Tsunoda  and 
Shiio,  1959).  Whether  the  former  inhibition  is  the  result  of  an  indirect 
suppression  of  transamination  by  cycle  block  or  a  direct  effect  on  the  oxi- 
dative pathway  of  this  amino  acid  is  not  known. 

iV-Acetylaspartate  occurs  at  a  relatively  high  concentration  in  mamma- 
lian and  avian  brain,  increasing  rapidly  after  birth.  Its  formation  involves 
direct  acetylation  of  aspartate  and  the  brain  has  little  ability  to  metabolize 
this  substance.  When  acetate-1-C^*  is  injected  intracerebrally,  some  of  the 
C^*  is  later  found  in  A^-acetylaspartate  (Jacobson,  1959).  The  injection  of 
malonate  with  the  acetate  reduces  the  incorporation  of  the  C^*  about  50%. 
The  injection  of  acetate  depresses  the  level  of  total  A^-acetylaspartate  and 
malonate  counteracts  this.  These  effects  are  quite  complex  and  difficult  to 
interpret.  The  concentration  of  malonate  injected  was  high  (1.34  M)  and 
could  have  caused  a  severe  fall  in  ATP  so  that  acetate  activation  prior  to 
acetylation  would  be  depressed.  The  rise  in  ^V-acetylaspartate  seen  with 
malonate  might  have  been  due  to  a  cycle  block  counteracting  the  effect 
of  the  injected  acetate,  whereby  cycle  intermediates  involved  in  trans- 
aminations would  be  decreased,  the  level  of  aspartate  being  maintained  with 
more  aspartate  available  for  acetylation.  There  are  so  many  pathways 
associated  with  aspartate  metabolism  and  acetylation  reactions  that  the 
final  effects  of  a  cycle  block  are  difficult  to  predict.  A  good  example  of  the 
complex  effects  of  malonate  on  amino  acid  metabolism  is  seen  in  Table 
1-19,  where  certain  types  of  amino  acid  in  potato  slices  increase  and  other 
types  decrease  during  incubation  with  malonate  (Romberger  and  Norton, 
19C1). 


EFFECTS  ON  AMINO  ACID  AND  PROTEIN  METABOLISM  155 

Effects  on  Protein  Synthesis 

The  intracellular  synthesis  of  protein  requires  the  simultaneous  opera- 
tion of  many  metabolic  pathways  and  thus  is  susceptible  to  inhibition  on  a 
variety  of  reactions.  Some  of  the  processes  involved  in  protein  synthesis 
are:  (1)  the  active  uptake  or  accumulation  of  exogenous  amino  acids,  (2) 
the  production  of  high-energy  substances  such  as  ATP  from  the  oxidative 
reactions  of  the  cycle  (except  in  anaerobes),  (3)  the  formation  of  amino  acid 
precursors,  again  mainly  by  the  operation  of  the  cycle,  and  (4)  all  the  com- 
plex reactions  for  the  activation  and  assemblage  of  the  amino  acids  into 
proteins.  There  are  thus  a  multitude  of  possible  sites  for  malonate  action 
but,  at  reasonable  concentrations,  the  most  important  mechanism  must 
be  a  cycle  block  leading  to  both  depletion  of  energy  supplies  and  decrease 
in  amino  acid  precursors.  There  is  no  evidence  that  malonate  can  interfere 
significantly  either  with  the  proteases  or  peptidases  involved  in  the  break- 
down of  proteins  to  amino  acids  or  with  the  terminal  assembling  reactions 
for  the  formation  of  protein. 

The  effects  of  malonate  on  the  uptake  and  accumulation  of  amino  acids 
by  cells  have  been  studied  in  three  types  of  tissue.  Excised  diaphragm  main- 
tains the  same  tissue/medium  ratio  for  glycine  as  in  the  whole  animal,  and 
the  marked  effects  of  2,4-dinitrophenol  indicate  that  glycine  is  concentrated 
actively  (Christensen  and  Streicher,  1949).  Malonate,  however,  at  concen- 
trations of  3-55  TaM  does  not  uniformly  alter  the  tissue/medium  ratio.  It 
is  possible  that  malonate  does  not  penetrate  adequately,  because  muscle 
is  often  rather  impermeable  to  anions.  The  situation  is  different  in  Ehrlich 
mouse  ascites  carcinoma  cells.  Glycine  is  accumulated  so  that  tissue/medium 
ratios  are  often  10-15.  In  two  experiments,  malonate  at  37  mM  decreased 
this  ratio  from  13.0  to  5.1  and  at  40  mM  from  13.9  to  4.0  (Christensen  and 
Riggs,  1952).  This  occurred  despite  the  fact  that  malonate  increased  the 
synthesis  of  glycine.  In  cell  suspensions  of  Gardner  lymphosarcoma  the 
uptake  of  labeled  glycine  is  inhibited  73%  and  the  uptake  of  alanine  56% 
by  10  mM  malonate  (Kit  and  Greenberg,  1951).  These  studies  demonstrate 
that  malonate  can  interfere  with  protein  s>Tithesis,  at  least  in  some  cells, 
by  inhibiting  the  initial  process  of  amino  acid  uptake. 

The  synthesis  of  protein  is  usually  strongly  inhibited  by  malonate,  but 
no  analyses  of  the  block  have  been  made  and  the  mechanisms  are  un- 
known (see  accompanying  tabulation).  The  formation  of  adaptive  enzymes 
has  often  been  taken  as  indicative  of  the  synthesis  of  general  cell  proteins, 
but  this  is  not  necessarily  so,  as  pointed  out  by  Mandelstam  (1961).  In  E. 
coli  any  substance  acting  as  a  substrate  and  source  of  energy  represses  en- 
zyme synthesis,  whereas  inhibitors,  such  as  malonate  and  2,4-dinitrophenol, 
counteract  such  effects  and  stimulate  the  synthesis.  Furthermore,  under 
conditions  in  which  /5-galactosidase  synthesis  is  inhibited,  the  incorporation 
of  leucine-C^*  into  cell  protein  is  not  affected.  The  lack  of  inhibition  in 


156 


1.    MALONATE 


Process 

Malonate 
(mi/) 

%  Change 

Reference 

Formation    of  induced   /5-galacto- 

sidase  in  E.  coli 

33.5 

+  31 

Mandelstam  (1961) 

100 

-     8 

134 

-  42 

167 

-  79 

Incorporation    of  leucine-C* 

into  tobacco  leaf  proteins 

10 

0 

Stephenson  et  al.  (1956) 

Protein  formation  in  chick  embryo 

tissue  culture 

10 

-100 

Gerarde  et  al.  (1952) 

Incorporation  of  glycine-2-C^* 

into  rat  liver  homogenates 

45 

-  86 

Peterson  and  Greenberg 

Incorporation  of  glycine-C'' 

(1952) 

into    antibody    in    rabbit    lymph 

nodes 

1.2 

-  67 

Ogata  et  al.  (1956) 

Incorporation  of  acetate- l-C^^ 

into  rat  liver  slices 

30 

-  50 

van  Vals  and  Emmelot 

Incorporation  of  acetate- l-C* 

(1957) 

into  various  tumor  slices 

30 

—  74  to  —90 

van  Vals  and  Emmelot 

Incorporation  of  glutamate-u-C^* 

(1957) 

into  Walker  carcinosarcoma 

6.25 

-23  to  -56 

Nyhan  and  Busch  (1957) 

Incorporation  of  glycine- 1-C" 

into  chick  embryo  proteins 

20 

-  57 

Quastel  and  Bickis  (1959) 

Incorporation  of  glycine- 1-C^* 

into  ascites  protein 

20 

-  92 

Quastel  and  Bickis  (1959) 

tobacco  leaves  was  attributed  to  the  presence  of  preformed  precursors  or 
energy  donors,  so  that  interference  with  metabolism  during  the  2-hr  in- 
cubation does  not  modify  the  assembling  of  the  proteins  (Stephenson  et  al., 
1956).  In  two  instances,  glucose  is  able  to  partially  reverse  the  effects  of 
malonate.  Glucose  addition  to  the  Walker  carcinosarcoma  slices  reduces  the 
inhibition  by  malonate,  sometimes  restoring  the  normal  rate  of  protein 
synthesis  (Nyhan  and  Busch,  1957),  and  in  ascites  cell  suspensions  glucose 
decreases  the  malonate  inhibition  from  92%  to  14%  (Quastel  and  Bickis, 
1959),  although  the  inhibition,  is  even  increased  slightly  in  chick  embryo. 
The  marked  glycolytic  activities  of  tumor  tissue  may  be  responsible  for  this 
phenomenon,  sufficient  energy  for  protein  synthesis  being  obtained  from 
noncycle  pathways. 

The  inhibition  of  amino  acid  uptake  and  the  synthesis  of  proteins  and 
enzymes  by  malonate  must  be  considered  in  long-term  experiments  or  in 
whole  animal  experiments,  since  this  could  secondarily  affect  many  other 


EFFECTS  ON  AMINO  ACID  AND  PROTEIN  METABOLISM  157 

metabolic  systems.  Most  enzymes  are  probably  in  a  state  of  simultaneous 
formation  and  degradation,  so  that  an  inhibitor  of  synthesis  would  induce 
a  steady  fall  of  the  enzyme  level  in  the  cells.  This  could  apply,  of  course, 
to  all  inhibitors  of  protein  synthesis. 

Effects   on    Urea   Formation 

The  terminal  product  of  much  protein  and  amino  acid  metabolism  is  urea 
and  it  has  been  found  that  under  certain  circumstances  malonate  inhibits 
the  formation  of  urea  quite  potently.  The  inhibition  has  been  mentioned  in 
connection  with  its  antagonism  by  fumarate  (page  116).  The  most  important 
reactions  of  the  urea  cycle  comj^rise  the  following,  assuming  that  glutamate 
is  the  immediate  amino-group  donor: 

(1)  Glutamate  +  oxalacetate   ->•  aspartate  +  a-ketoglutarate  (transaminase) 

(2)  Aspartate  +  citrulline  -j-  ATP   ->■  argininosuccinate  +  ADP  +  P, 

(argininosuccinate  synthetase) 

(3)  Argininosuccinate   ->  arginine  +  fumarate  {argininosuccinase) 

(4)  Fumarate  ->  oxalacetate                               (fumarase  and  lyialate   dehijdrogenase) 

(5)  Arginine  ->  ornithine  +  urea  (arginase) 

(6)  Ornithine  +  NH3  +  CO2   ~>  citrulline  {citrulline  synthetase) 

Glutamate  +  NH3  +  CO2  +  ATP   -►  a-ketoglutarate  +  urea  +  ADP  +  P, 

This  urea  cycle  thus  makes  contact  with  the  tricarboxylate  cycle  at  several 
points.  The  a-ketoglutarate  formed  in  the  over  all  reaction  can  be  oxidized 
through  succinate  to  oxalacetate  or  can  be  transaminated  to  regenerate  glu- 
tamate. The  operation  of  the  urea  cycle  thus  requires  sources  for  oxalace- 
tate and  ATP,  both  of  which  may  be  blocked  by  malonate. 

Cohen  and  Hayano  (1946)  found  that  5.7  raM  malonate  inhibits  the  con- 
version of  citrulline  to  arginine  90%  in  liver  homogenates  when  glutamate 
is  the  amino  donor.  The  mechanism  of  the  inhibition  was  not  apparent 
at  that  time.  These  results  were  confirmed  by  Fahrlander  et  al.,  (1947) 
and,  in  addition,  they  showed  that  fumarate  or  malate,  can  counteact  the 
inhibition,  indicating  a  block  of  succinate  oxidation.  They  interpreted 
the  mechanism  as  a  depletion  of  ATP  and  a  consequent  inhibition  of  reac- 
tion (2).  Subsequently,  they  showed  that  low  malonate  concentrations  (1-2 
ToM)  inhibit  urea  formation  as  much  as  75%  and  felt  this  was  evidence  for 
a  specific  action  on  succinate  dehydrogenase  (Fahrlander  et  al.,  1948). 
The  ATP  level  in  the  homogenates  drops  from  130  to  49.6  in  the  presence 
of  2.5  yrM  malonate  and  fumarate  restores  the  ATP  level  to  normal.  It 
was  believed  that  glutamate  not  only  furnishes  the  amino  group  but 
also  cycle  substrates  from  which  the  energy  is  derived;  a  block  by  malonate 
at  the  succinate  level  would  reduce  the  amount  of  ATP  formed.  Krebs  and 


158 


1.    MALONATE 


Eggleston  (1948)  then  demonstrated  a  differential  effect  of  malonate  on 
the  formation  of  urea  depending  on  whether  glutamate  or  aspartate  is 
used  as  the  amino  donor,  the  inhibition  being  less  in  the  latter  case.  An 
elucidation  of  the  true  mechanism  of  the  inhibition  was  presented  by 
Ratner  and  Pappas  (1949),  who  showed  a  very  definite  differential  effect 
of  malonate  when  glutamate  and  aspartate  are  used  (see  tabulation).  The 


%  Inhibition 

by  malonate  20  laM 

Substrate 

Aspartate 

Glutamate 

Arginine  synthesis     Oj 

uptake 

Arginine  synthesis    Oj  uptake 

None 

Pyruvate 

Oxalacetate 

Fumarate 

a-Ketoglutarate 

6 
11 

8 

1 
Stim  6 

27 
32 

20 

7 
16 

73                        38 

57                        37 

Stim  22                      5 

Stim     2                      9 

66                        40 

transamination  forming  aspartate  from  glutamate  is  not  inhibited  by  mal- 
onate so  the  mechanism  must  be  sought  elsewhere.  It  was  proposed  that 
malonate  prevents  the  formation  of  oxalacetate  and  thus  indirectly  blocks 
the  formation  of  aspartate;  fumarate  would,  of  course,  overcome  this  block. 
They  opposed  the  idea  that  ATP  depletion  is  important  and  felt  that  the 
ATP  derived  from  a-ketoglutarate  oxidation  would  be  sufficient.  However, 
they  did  not  by  any  means  disprove  the  ATP  depletion  hypothesis  and  it  is 
quite  possible  that  it  also  plays  a  role  in  assigning  an  over  all  mechanism 
for  the  inhibition.  Miiller  and  Leuthardt  (1950)  extended  these  observations 
by  showing  chromatographically  that  malonate  inhibits  the  formation  of  as- 
partate by  reducing  the  formation  of  oxalacetate  from  a-ketoglutarate,  and 
also  demonstrated  conclusively  that  the  transamination  reaction  itself  is 
not  sensitive  to  malonate.  It  should  be  noted  that  in  the  reactions  written 
above,  oxalacetate  appears  to  be  regenerated  in  the  arginosuccinase  reac- 
tion followed  by  the  hydration  and  oxidation  of  fumarate.  but  this  is  appar- 
ently not  sufficient  to  maintain  the  cycle,  probably  because  much  of  the 
oxalacetate  disappears  in  other  reactions.  This  is  why  an  external  source 
of  oxalacetate  is  necessary. 

EFFECTS   OF   MALONATE    ON    PORPHYRIN    SYNTHESIS 


The  pathway  for  the  synthesis  of  porphyrins  in  both  animals  and  plants 
originates  in  the  cycle  in  the  condensation  of  succinyl-CoA  with  glycine 
(Fig.  1-16).  The  succinyl-CoA  can  be  formed  either  from  a-ketoglutarate 


EFFECTS   OF  MALONATE   ON  PORPHYRIN   SYNTHESIS  159 

or  from  succinate;  the  latter  reaction  requires  ATP  and  is  catalyzed  by 
succinyl-CoA  synthetase  (P-enzyme).  A  total  of  8  molecules  of  succinate  and 
8  molecules  of  glycine  is  required  for  the  synthesis  of  a  molecule  of  proto- 
porphyrin. The  close  connection  between  this  pathway  and  the  succinate 
steps  of  the  cycle,  and  the  great  iniportance  of  porphyrin  synthesis  in  all 
tissues,  make  the  study  of  the  action  of  malonate  on  this  system  interesting. 
We  may  speculate  on  the  various  ways  in  which  malonate  could  modify 
porphyrin  synthesis.  (1)  If  the  succinyl-CoA  is  formed  in  the  cycle  through 
a-ketoglutarate,  malonate  could  restrict  its  formation  by  blocking  the  cycle 


Acetyl  -Co  A 

Fumarote 

(Malonate) 


I 


a-Ketoglutarote  Succmote 


-Succinyl  -Co  A 

^^  Glycine 

a  -Ammo-  /9  -  ketoadipote 
S  -  Aminolevulinote 

I 

Porphobilinogen 


Protoporphyrin 

Fig.   1-16.   The  pathways  involved  in 
porphyrin  biosynthesis. 

and  reducing  the  rate  of  acetyl-CoA  entry,  especially  if  no  noncycle  source 
of  oxalacetate  is  available.  (2)  If  succinyl-CoA  can  be  formed  through  the 
cycle  readily  in  spite  of  a  malonate  block,  malonate  might  divert  more  suc- 
cinate into  the  synthesis  of  porphyrin  by  inhibiting  succinate  oxidation. 
(3)  If  the  succinyl-CoA  arises  from  succinate,  this  requires  ATP  and  mal- 
onate could  deplete  the  system  of  ATP.  (4)  Malonate  might  deplete  the 
system  of  coenzyme  A  by  the  formation  of  malonyl-CoA.  (5)  It  is  possible 
in  some  way  that  malonate  might  inhibit  the  formation  of  glycine,  although 
this  is  rather  unlikely  because  there  are  usually  several  pathways  avail- 
able for  glycine  synthesis.  The  effects  of  malonate  will  thus  depend  on  the 
type  of  preparation  used  and  the  conditions  of  the  experiment. 

Duck  erythrocytes  (intact  or  hemolyzed)  incubated  with  succinate  and 
glycine  form  porphyrin.  Succinyl-CoA  could  be  formed  from  succinate  either 
directly  or  through  the  cycle  and  the  relative  importance  of  these  pathways 
may  be  demonstrated  by  the  use  of  succinate-C^'*  with  subsequent  deter- 


160  1.    MALONATE 

mination  of  the  porphyrin  labeling  (Shemin  and  Kumin,  1952).  Succinate- 
C^^OO"  when  oxidized  through  the  cycle  gives  rise  to  a-carboxyl-labeled 
a-ketoglutarate  and  hence  to  unlabeled  succinyl-CoA;  therefore,  no  porphy- 
rin labeling  should  result  from  this  pathway.  However,  succinate-C"00~ 
could  also  directly  form  succinyl-CoA,  which  in  this  case  would  be  labeled 
and  C^*  would  be  found  in  porphyrin.  On  the  other  hand,  succinate-C^^Hg 
would  form  labeled  succinyl-CoA  by  both  pathways.  If  it  is  assumed  that 
malonate  inhibits  the  oxidation  of  succinate  and  the  cycle  pathway  only, 
malonate  should  not  inhibit  any  porphyrin  labeling  after  incubation  with 
succinate-C^*00~,  but  should  inhibit  appreciably  the  porphyrin  labeling 
from  succinate-C^^Hg.  This  was  found  by  Shemin  and  Kumin,  as  the  aver- 
aged results  in  the  accompanying  tabulation   show  (figures   are   counts/ 

Hemin  Control         Malonate  20  mif        %  Change 

From  succinate-C'^OO-  215  210  -  2 

From  succinate-Ci^Hj  1025  421  -59 

minute  for  intact  erythrocytes).  It  would  appear  that  both  pathways  are 
operative  in  these  cells.  The  failure  of  malonate  to  increase  the  porphyrin 
labeling  from  succinate-C^*00~  is  rather  surprising  because  one  might 
expect  malonate  to  divert  some  of  the  succinate  from  the  cycle  into  the 
formation  of  succinyl-CoA.  It  is  possible  that  this  effect  is  somewhat  coun- 
teracted by  an  inhibition  on  succinyl-CoA  synthetase. 

Further  information  on  porphyrin  synthesis  and  the  effects  of  malonate 
were  obtained  by  Wriston  et  al.  (1955)  by  the  use  of  labeled  acetate.  Different 
malonate  effects  were  obtained  when  methyl-labeled  and  carboxyl-labeled 
acetate  were  incubated  with  glycine,  the  inhibition  of  porphyrin  labeling 
being  much  greater  with  the  former  (see  tabulation).  This  is  the  expected 

Hemin  Control         Malonate  20  mM      %  Change 

From  acetate-Ci^Ha  376  174  -54 

From  acetate-C'^OO-  72.5  67  -  8 


result,  because  the  formation  of  labeled  succinyl-CoA  from  acetate-C^*00~ 
does  not  involve  the  complete  cycle  and  the  C^*  pathway  does  not  go  through 
the  succinate  oxidation  step,  whereas  porphyrin  labeling  from  acetate- 
C^^Hg  depends  on  the  operation  of  the  entire  cycle  (except  for  the  contri- 
bution from  the  y-C  atom  of  a-ketoglutarate).  Furthermore,  the  labeling 


EFFECTS   OF  MALONATE   ON   PORPHYRIN   SYNTHESIS  161 

in  the  porphyrin  from  acetate-C^^Hg  should  be  altered  by  malonate. 
Labeling  in  the  carbon  atoms  of  the  A  and  B  pyrrole  rings  of  protoporphyrin 
occurs  after  incubation  of  duck  erythrocytes  with  acetate-C^^Hg  and  glycine. 

CH,  9 

II 
6  H3C  CH    8 

'  I  I 

'       fi^N       \ 

I 

H 


Control 

Malonate  10 

mM 

%  Inhibition 

Total  porphyrin 

186,000 

54,000 

71 

Pyrroles  A  and  B 

88,000 

27,500 

69 

Carbon  4 

13,000 

3,700 

72 

Carbon  5 

12,000 

900 

93 

Carbon  6 

20,000 

9,400 

53 

Acetate-Ci^Hg  will  lead  directly  to  -OOC-C^^HaCHa-CO-CoA  and  if  the 
cycle  is  blocked  completely  by  malonate,  carbons  6  and  9  only  will  be 
labeled,  except  for  some  labeling  of  carbons  4  and  8  due  to  the  reversible 
reaction  succinyl-CoA  :^  succinate  (as  long  as  ATP  is  available).  Carbons  2, 
3,  and  5  should  not  be  labeled.  This  is  essentially  seen  in  the  tabulation. 
The  cycle,  of  course,  is  not  blocked  completely  so  that  some  labeling  in 
carbon  5  occurs.  The  over  all  inhibition  is  due  to  a  depression  of  the  entry 
of  acetate  into  the  cycle.  These  experiments  not  only  show  the  variable 
effects  of  malonate  on  a  pathway  associated  with  the  cycle  but  well  illus- 
trate the  use  of  an  inhibitor  to  elucidate  a  metabolic  pathway. 

The  analysis  of  the  action  of  malonate  on  porphyrin  synthesis  was  ex- 
tended by  Granick  (1958)  in  his  work  with  chicken  erythrocytes.  The  for- 
mation of  protoporphyrin  is  innibited  90%  by  10  mM  malonate  when  only 
glycine  is  present,  85%  when  succinate  is  added,  and  80%  when  a-keto- 
glutarate  is  added.  The  effects  of  different  concentrations  of  malonate  are 
shown  in  Fig.  1-17.  Malonate  could  decrease  the  incorporation  of  succinate 
into  porphyrin  by  blocking  the  cycle  and  reducing  the  ATP  level,  and  thus 
inhibit  both  pathways  of  succinate-CoA  formation  from  succinate.  However, 
the  quite  strong  inhibition  of  protoporphyrin  formation  from  glycine 
-f  a-ketoglutarate  is  surprising.  Inhibition  of  the  step  a-ketoglutarate  — * 
succinyl-CoA  is  not  likely  as  the  only  explanation,  because  1  raM  malonate 
inhibits  protoporphyrin  synthesis  32%  and  there  is  no  reason  for  thinking 
that  this  low  concentration  would  inhibit  a-ketoglutarate  oxidase.  The  for- 


162 


1.    MALONATE 


mation  of  protoporphyrin. from  S-aminolevulinate  is  not  inhibited  by  mal- 
onate  so  that  an  action  on  this  part  of  the  pathway  is  excluded.  Granick 
suggested  that  malonate  reacts  with  coenzyme  A  and  thus  depletes  the 
system  so  that  succinyl-CoA  cannot  be  so  readily  formed.  These  effects 
were  confirmed  in  lysed  chicken  erythrocytes  by  Brown  (1958).  Porphyrins 
are  not  formed  in  these  preparations  but  (5-aminolevulinate  is  formed  from 
glycine  and  succinyl-CoA  derived  from  a  variety  of  cycle  substrates.  Mal- 
onate at  10  mM  inhibits  this  reaction  30%  when  the  incubation  is  with 
glycine  and  citrate,  substantiating  the  action  on  this  region  of  the  pathway. 


1  50 


2  I 

A  Arsenite 

B  DNP 

C  Fluoroacetote 

D  Malonate 


pi  ► 

Fig.  1-17.  Effects  of  four  inhibitors  on  the  sjti thesis 

of  protoporphyrin  in   chicken  erythrocytes  with   the 

substrates  as  indicated.  (From  Granick,  1958.) 


It  is  interesting  that  the  addition  of  succinate  to  glycine  and  citrate  in  the 
incubation  medium  leads  to  an  inhibition  of  (5-aminolevulinate  synthesis. 
This  was  shown  to  be  due  to  the  formation  of  oxalacetate,  which  inhibits 
a-ketoglutarate  oxidase.  Malonate  is  able  to  overcome  this  inhibition  by 
succinate  through  the  prevention  of  oxalacetate  formation.  This  indicates 
another  minor  mechanism  for  the  effect  of  malonate  on  porphyrin  synthesis, 
namely,  the  reduction  in  oxalacetate  concentration  and  a  consequent  release 
from  any  inhibition  on  the  oxidation  of  a-ketoglutarate.  Finally,  we  may 
note  that  coproporphyrin  synthesis  from  glycine  and  a-ketoglutarate  in 


EFFECTS    ON    METABOLIC    PATHWAYS  163 

Rhodopseitdomonas  spheroides  is  inhibited  50%  by  20  mif  malonate  and 
75%  by  40  mM  malonate  (Lascelles,  1956),  indicating  again  some  inhibition 
of  the  formation  of  succinyl-Co A. 

The  incorporation  of  iron  into  heme,  as  demonstrated  with  Fe^^,  is  not 
inhibited  by  10  mM  malonate  in  canine  reticulocytes  (Yoshiba  et  al.,  1958) 
but  is  inhibited  26%  in  chicken  erythrocytes  (Kagawa  et  al.,  1959).  It  is 
possible  in  the  latter  case  that  the  inhibition  is  due  to  the  chelation  of  part 
of  the  Fe^^,  making  it  unavailable  for  incorporation. 

EFFECTS  OF   MALONATE 
ON    MISCELLANEOUS   METABOLIC   PATHWAYS 

There  have  been  many  reports  on  the  actions  or  lack  of  action  of  malonate 
on  enzyme  reactions  or  metabolic  pathways  of  varying  degrees  of  impor- 
tance. Some  of  these  are  worth  mentioning,  either  because  they  indicate 
areas  where  further  study  might  be  profitable  or  because  they  provide  some 
evidence  for  noncycle  actions  of  malonate. 

One  might  expect  very  little  effect  of  malonate  on  photosynthesis  but, 
although  very  little  work  has  been  done,  in  every  case  some  effect  has  been 
observed.  Even  the  Hill  reaction  is  susceptible  to  inhibition  (Ehrmantraut 
and  Rabinowitch,  1952).  This  reaction  is  the  photochemical  oxidation 
of  water  with  the  production  of  oxygen  and  the  reduction  of  a  substance, 
usually  quinone,  other  than  COg.  In  Chlorella  this  reaction  is  inhibited 
30%  by  6  mM  malonate  and  50%  by  60  mM  malonate.  The  inhibition  is, 
surprisingly,  prevented  by  fumarate,  indicating  that  the  site  of  action  is 
succinate  dehydrogenase  and  that  this  enzyme  takes  part  in  the  transport 
of  hydrogen  in  the  Hill  reaction,  which  would  not  be  the  case  if  quinone 
were  serving  as  the  immediate  hydrogen  acceptor.  It  may  also  be  that 
malonate  does  not  inhibit  the  Hill  reaction  directly,  but  depletes  the  cells 
of  cycle  intermediates  or  other  cycle  products  necessary  for  the  Hill  reaction 
to  proceed.  Malonate  not  only  inhibits  the  total  incorporation  of  C^^Og  in 
Scenedesmus  by  about  20%,  but  almost  completely  blocks  the  formation  of 
labeled  malate  (Bassham  et  al.,  1950).  This  was  taken  as  evidence  that  mal- 
ate  is  not  on  the  direct  line  of  phosphoglycerate  synthesis,  but  it  also  dem- 
onstrates that  by  some  mechanism  malonate  can  inhibit  COg  incorporation. 
An  inhibition  of  glucose  formation  in  Chlorella  by  malonate  has  also  been 
reported  (Kandler,  1955),  although  there  is  less  inhibition  in  the  light  than 
in  the  dark.  The  synthesis  of  glucose  was  believed  to  be  closely  related  to 
the  formation  of  high  energy  phosphate  intermediates,  and  it  is  thus  inter- 
esting that  malonate  inhibits  the  photosynthetic  phosphorylation  of  ADP 
in  Rhodospirillum,  although  the  inhibition  is  only  17%  at  the  very  high 
concentration  of  100  mM  (Smith  and  Baltscheffsky,  1959).  With  these  lim- 
ited observations  on  the  effects  of  malonate,  it  must  be  admitted  that  it 


164  1.    MALONATE 

is  difficult  to  fit  the  data  into  the  modern  concepts  of  the  carbon  pathway  in 
photosynthesis,  which  does  not  directly  involve  the  cycle,  and  particularly 
to  understand  the  mechanism  whereby  the  Hill  reaction  is  inhibited. 

Malonate  is  also  able  to  interfere  in  the  metabolism  of  glycerol.  Glycerol 
is  fermented  to  succinate,  accompanied  by  the  uptake  of  CO2,  in  Propioni- 
hacterium  pentosaceum,  and  30  mM  malonate  inhibits  both  the  glycerol 
fermentation  and  the  CO2  uptake  around  10%  (Wood  and  Werkman,  1940). 
It  is  impossible  to  attribute  this  to  an  action  on  succinate  dehydrogenase. 
The  oxidation  of  glycerol  may  involve  an  initial  phosphorylation  with 
subsequent  formation  of  pyruvate  and  entry  into  the  cycle: 

Glycerol  +  ATP   ->  glycerophosphate  ->  pyruvate   ->  cycle 

The  phosphorylation  is  inhibited  in  rat  liver  homogenates  (Ruffo  and  D'A- 
bramo,  1952),  but  the  total  oxidation  is  not  markedly  affected  in  the 
mycobacteria  (G.  J.  E.  Hunter,  1953).  Malonate  at  10  mM  inhibits  5%  in 
M.  stercoris  and  stimulates  4-6%  in  M.  smegmatis  and  M.  butyricum.  It  is 
surprising  that  greater  inhibition  is  not  observed  if  the  oxidation  does 
involve  the  cycle.  In  castor  bean  cotyledons,  malonate  at  high  concentra- 
tions has  very  marked  effects  on  the  utilization  of  glycerol  (Beevers,  1956). 
At  70  mM  the  C^^Oa  formation  from  labeled  glycerol  is  inhibited  97%  and 
the  sucrose  formation  is  inhibited  82%,  while  at  130  mM  the  oxygen 
uptake  is  inhibited  60%.  Such  high  concentrations  may  interfere  with  the 
formation  of  ATP  and  hence  depress  phosphorylation  and  also  block  the 
cycle  CO2  release. 

The  stimulation  of  muscle  respiration  by  insulin  is  inhibited  potently  by 
malonate  (Stare  and  Baumann,  1940).  Insulin  almost  doubles  the  respiration 
of  minced  breast  muscle  from  depancreatized  pigeons  and  1  mM  malonate 
inhibits  this  increase  92%.  Fumarate  is  able  to  overcome  both  this  inhibi- 
tion and  the  inhibition  of  nonstimulated  respiration  completely.  This 
interesting  action  in  vitro  led  to  a  study  of  the  antagonism  in  the  whole 
animal.  Solutions  of  sodium  malonate  were  injected  subcutaneously  in 
rabbits  either  before  or  with  insulin  and  the  drop  in  blood  glucose  was  much 
less  than  with  insulin  alone.  Insulin  (4  units)  decreases  the  blood  glucose  65% 
in  4  hr,  whereas  with  malonate  present  the  reduction  is  only  13%  and  none 
of  the  rabbits  goes  into  convulsions.  Malonate  alone  increases  the  blood 
glucose  26%.  The  respiratory  inhibition  in  the  muscle  mince  could  be 
explained  on  the  basis  of  a  typical  cycle  block  (although  the  degree  of  inhi- 
bition is  surprising  for  a  concentration  of  1  mM),  but  the  inhibition  of  glu- 
cose utilization  in  the  animal  is  more  complicated.  The  initial  phosphoryla- 
tion of  glucose  and  its  uptake  could  have  been  inhibited  indirectly  by  a  re- 
duction of  the  available  ATP,  or  it  could  have  resulted,  at  least  in  part, 
from  a  hyperglycemic  action  of  malonate  unrelated  directly  to  the  insulin 
stimulation. 


EFFECTS    ON    METABOLIC    PATHWAYS  165 

Succinate  and  propionate  are  formed  anaerobically  in  Ascaris  muscle 
from  glucose  and  lactate,  presumably  by  the  following  pathway: 

Glucose  ^ 

^>fc^  +CO,  +4H      „  -COj    ^ 

Pyruvate  »=-  Oxalacetate  *-  Succinate *~  Propionate 


Lactate  -""^ 

Malonate  at  20  mM  does  not  appreciably  inhibit  the  decarboxylation  of  suc- 
cinate to  propionate  (about  a  13%  reduction  in  total  radioactivity)  but  the 
small  inhibition  indicates  a  possible  competition  with  succinate  for  the 
enzyme.  However,  the  incorporation  of  lactate-2-C'-^  into  succinate  is  inhibited 
almost  90%.  If  succinate  is  formed  by  reduction  of  fumarate  derived  from 
oxalacetate,  malonate  would  be  expected  to  inhibit  well,  not  only  because 
of  the  effect  on  the  succinate  dehydrogenase  but  also  by  an  inhibition  of 
oxalacetate  formation.  Malonate  inhibits  the  formation  of  labeled  propionate 
from  lactate-2-C^*  65%.  The  smaller  inhibition  compared  to  that  for  suc- 
cinate formation  implies  another  less  important  pathway  for  the  formation 
of  propionate,  perhaps  by  direct  reduction,  as  shown  in  several  bacteria. 

The  metabolism  of  glyoxylate  by  Kver  mitochondria  is  rather  complex; 
it  is  decarboxylated  to  formate  by  a  devious  route,  it  may  be  oxidized  to 
oxalate,  or  it  may  be  aminated  to  glycine  (Crawhall  and  Watts,  1962). 
Malonate  inhibits  the  decarboxylation  competitively  but  does  not  interfere 
with  the  formation  of  oxalate  or  glycine;  indeed,  the  latter  may  be  stimul- 
ated slightly  due  to  diversion  in  a  branched  chain.  The  decarboxylation 
reaction,  which  requires  glutamate,  is  quite  sensitive  to  malonate,  around 
50%  inhibition  occurring  at  0.15  mM,  both  substrates  being  at  3  mM. 
This  would  certainly  appear  to  be  one  system  in  which  a  marked  effect  can 
be  exerted  by  malonate  at  low  concentrations  and  which  is  unrelated  to 
succinate  oxidation. 

The  synthesis  of  acetylocholine  is  an  endergonic  process  and  is  related 
to  the  cycle  both  for  the  supply  of  energy  and  with  respect  to  the  utilization 
of  acetyl-CoA.  The  effects  of  inhibitors  on  acetylcholine  synthesis  and  hy- 
drolysis are  particularly  important  when  considering  the  mechanisms  by 
which  malonate  can  alter  nerve  and  muscle  function.  Unfortunately,  only 
one  study  of  the  action  of  malonate  has  been  made  (Torda  and  Wolff, 
1944  a).  The  formation  of  free  acetylcholine  in  minced  frog  brain,  in  the 
presence  of  physostigmine  to  prevent  hydrolysis,  is  inhibited  32%  by  0.08 
mM,  46%  by  0.8  mM,  and  49%  by  8  mM;  the  inhibition  of  total  acetyl- 
choline is  about  the  same.  Succinate,  fumarate,  and  citrate  increase  acetyl- 
choline formation.  It  would  be  interesting  to  investigate  the  effects  of  mal- 
onate on  the  purer  enzyme  systems  now  available  for  acetylcholine  synthe- 
sis to  determine  if  the  inhibition  is  a  direct  effect  or  secondary  through 
ATP  depletion.  The  effects  of  malonate  in  the  intact  cell  may  be  quite 
complex,  because  malonate  might  suppress  the  incorporation  of  acetyl-CoA 


166  1.    MALONATE 

into  the  cycle  and  thereby  lead  to  a  greater  availability  of  acetyl-CoA  for 
choline  acetylation.  However,  it  is  not  known  if  the  acetyl-CoA  pool  is 
common  to  both  the  cycle  and  the  synthesis  of  acetylcholine.  In  many 
cases  the  effects  of  malonate  are  due  only  to  a  depression  of  the  cycle  oper- 
ation and  a  decreased  formation  of  ATP.  For  example,  in  the  synthesis  of 
chondroitin  sulfate  in  tibial  condyles  of  chick  embryos,  the  fixation  of  sulfate 
is  inhibited  by  malonate  in  a  parallel  fashion  to  the  inhibition  of  respiration 
(Boyd  and  Neuman,  1954).  The  fixation  of  sulfate  requires  ATP,  as  shown 
by  the  marked  inhibition  with  2,4-dinitrophenol,  so  that  here  the  mechanism 
of  malonate  action  is  simply  an  inhibition  of  energy  formation.  Other  proc- 
esses, such  as  calcium  deposition  in  tibial  cartilage  (Hiatt  et  al.,  1953), 
do  not  require  energy  and  are  not  inhibited  by  malonate. 

EFFECTS   OF   MALONATE 
ON  THE  ENDOGENOUS  RESPIRATION 

The  alterations  of  the  most  important  metabolic  pathways  by  malonate 
have  been  discussed,  and  we  shall  now  conclude  this  aspect  of  the  subject 
with  a  survey  of  the  effects  on  the  total  oxygen  uptake  of  cells  respiring  in 
the  absence  of  any  external  substrate.  Although  the  interpretation  of  the 
results  of  such  studies  is  very  difficult,  the  changes  in  the  endogenous  respi- 
ration have  been  examined  more  frequently  than  any  other  response  to 
malonate.  There  is,  thus,  a  vast  and  variable  mass  of  data,  some  of  which 
is  summarized  in  Table  1-26.  The  aim  of  most  of  these  investigations  has 
been  to  demonstrate  the  absence  or  presence  of  the  cycle  in  the  types  of 
cells  tested,  and  we  must  attempt  to  assess  the  validity  of  conclusions  based 
on  the  response  to  malonate.  The  most  unsatisfactory  work  has  been  done, 
and  the  most  unjustified  conclusions  have  been  drawn,  in  studies  of  this 
type,  inasmuch  as  the  inherent  complexities  of  the  situations  have  seldom 
been  appreciated.  Although  the  cycle  has  a  wide  distribution  in  the  cells  of 
microorganisms,  plants,  and  animals,  its  operation  during  the  metabolism 
of  endogenous  substrates  is  quite  variable  and  dependent  on  the  state  and 
past  history  of  the  cells. 

Factors  That  May  Determine  the  Degree  of  Malonate  Inhibition 

Certain  basic  factors  should  be  considered  in  every  investigation  of  the 
susceptibility  of  the  endogenous  respiration  to  malonate.  Although  some 
of  these  have  been  mentioned  previously  and  some  will  be  taken  up  in 
greater  detail  later,  it  may  be  convenient  to  enumerate  here  the  most  im- 
portant. 

(a)  Intrinsic  susceptibility  of  succinate  dehydrogenase  to  malonate.  This  en- 
zyme from  different  species  varies  a  good  deal  in  its  ability  to  bind  mal- 


EFFECTS    ON   THE    ENDOGENOUS    RESPIRATION  167 

onate,  as  is  evident  from  the  range  of  K^  values  observed  (page  33).  It  is, 
perhaps,  too  often  assumed  that  in  every  organism  the  succinate  dehydro- 
genase wiU  be  readily  blocked  by  malonate  and  that  the  inhibition  of  the 
endogenous  respiration  depends  only  on  the  importance  of  the  enzyme  in 
the  total  oxygen  uptake.  It  is  mandatory  to  demonstrate  the  sensitivity 
of  succinate  dehydrogenase  to  malonate  in  the  preparation  being  studied. 

(b)  Degree  to  which  intracellular  succinate  dehydrogenase  is  inhibited.  In 
addition  to  the  intrinsic  susceptibility,  there  are  other  factors  which  can 
alter  the  inhibition  occurring  within  the  cell.  The  concentration  of  succinate, 
both  initially  and  following  the  accumulation  resulting  from  the  inhibition, 
may  be  high  enough  to  oppose  the  malonate  effect  appreciably.  The  relative 
stability  of  plant  respiration  to  malonate  has  been  attributed  to  the  high 
concentrations  of  succinate  and  other  organic  anions  in  plant  cells.  It  is 
probably  very  seldom  that  an  inhibition  even  approaching  completeness 
can  be  achieved  in  cells  at  concentrations  likely  to  be  specific. 

(c)  Specificity  of  malonate  inhibition.  Inhibition  of  the  endogenous  res- 
piration can,  of  course,  arise  from  actions  other  than  on  succinate  dehydro- 
genase. If  the  object  of  the  study  is  to  evaluate  the  contribution  of  the 
cycle  to  the  oxygen  uptake,  inhibitions  on  noncycle  pathways  must  be 
eliminated.  At  the  high  malonate  concentrations  often  used  (see  Table  1-26), 
there  is  certainly  no  assurance  that  the  inhibition  is  specific. 

(d)  Intracellular  concentration  of  malonate.  Malonate  does  not  penetrate 
readily  into  most  cells,  especially  at  physiological  external  pH  values,  so 
that  the  internal  concentrations  of  malonate  may  be  far  below  those  in 
the  surrounding  medium  (see  page  190).  The  degree  of  respiratory  inhibition 
observed  is  probably  often  more  of  a  measure  of  malonate  penetrability 
than  of  the  nature  or  susceptibility  of  the  metabolic  systems.  There  are 
several  instances  in  Table  1-26  in  which  the  inhibition  rises  with  destruction 
of  the  normal  tissue  structure  or  the  removal  of  permeability  barriers.  In 
general,  the  inhibition  is  greater  in  homogenates  than  in  minces,  and  greater 
in  minces  than  in  slices  or  intact  cells.  The  results  of  Bonner  (1948)  on 
Avena  coleoptiles  are  interesting  in  this  regard.  Soaking  in  water  for  24  hr 
increases  the  susceptibility  to  malonate  and  removal  of  the  endosperm 
further  increases  the  inhibition.  The  many  observations  that  a  lowering 
of  the  pH  augments  the  inhibition  also  provide  evidence  of  the  importance 
of  permeability. 

(e)  Metabolism  of  malonate.  Many  tissues  and  organisms  can  metabolize 
malonate  to  acetyl-CoA,  the  oxidation  of  which  contributes  to  the  oxygen 
uptake  (see  page  228).  Some  of  the  respiratory  stimulations  noted  with 
malonate  must  be  due  to  this,  and  it  is  likely  that  the  experimentally  de- 
termined inhibition  in  other  cases  is  reduced  from  that  which  would  be  ob- 
served if  malonate  were  not  metabolized.  The  best  way  to  test  for  and 


168 


1.    MALONATE 


^ 


1^ 

1— 1 

05     'ki 

(N 

t^ 

1— 1       CO 

-* 

-^ 

■ — -  -k^ 

05 

o 

!h 

'a 

o 

" — ' 

S 

ft 

"e 

'^ 

>« 

-c 

S 

!3     a; 

u 

c 

c 

cS     O 

S 

ce 

o 

-iU     t^ 

"m 

u     |3 

S3 

o 

> 

m 


it       >« 


-d 

rt 

c 

rt 

TJ 

C 

o 

m 

t^ 

o 

i 

PQ 

s  a  s  s 

"-J3      '-iS       +3       -t3 

M   02   02   03 


« 


ft 


CO     ■*     QO    "M     O    "M     O    'M 


GO    00 


'MOO-HOOOTt<OiCOO'OiC 


O    -Tit 


11^ 


«       (B 

O)     03 

ft     ft 

t»     tc 

s    s 

^    r*^ 

02   02 

02    W 

>. 

% 


ft 

02 


oq  ^ 


g^ 


•i 

as 

=0 

^ 

^ 

_£ 

=0 

s*. 

O 

'S. 

O 

C 

"^ 

e 

i 

1 

=5 

0^  -^ 


EFFECTS    ON   THE    ENDOGENOUS    RESPIRATION 


169 


c3    ^' 


2    OJ  ^ 


Oi 


o 


Z  o 


W 


t^    O    <M    00    Tj<    <M 


-a 


^         3    c 


c 

c 

C 

C 

c 

£ 

j5 

c 

c 

g 

c 

3 

C 

0 

_o 

_o 

_o 

0 

0 

_o 

_o 

_o 

_o 

.2 

0 

_o 

a 

tr. 

■5 

'S 

'S 

'S 

'S 

'm 

'm 

"m 

'ro 

"m 

s 

s 

c 

c 

c 

05 

-2 

C 

C 

fi 

C 

C 

f^ 

C 

0) 

(L> 

V 

D 

v 

<a 

a:i 

v 

«> 

0) 

ID 

0 

CO 

a 

a 

a 

a 

a 

15 

a 

a 

a 

a 

a 

a 

a 

« 

CO 

m 

m 

03 

m 

P 

CO 

CO 

m 

(O 

(C 

03 

ai 

_o 

S 

S 

3 

^ 

3 

^ 
P 

3 

3 

3 

3 

3 

3 

3 

CO 

CB 

J» 

M 

M 

M 

m 

M 

m 

m 

CC 

CO 

s 

03     ro 

3     3 

-     3 


I  i 

O      S 
CO      «i) 


.e   -3 


M     ®    'S     -g 


&q 


o     o     cS  '  «     „     „ 


3     S 

«=  J  t;  S 


-3    -3 


«     e6 


S   3  2 


<=«  ^    a  .'^ 

m  ft  w  w 


^ ^ 

^^^ 

W 

^_^ 

00 

00 

05 

-* 

Tl< 

T3 

■* 

05 

Oi 

C 

Oi 

)-H 

y—i 

cS 

1— 1 

^-^ 

^— ^ 

^— ^ 

Lh 

tH 

3 

tl 

« 

<U 

3 
-0 

« 

3 

3 

3 

3 

c 

3 

0 

0 

0 

m 

pq 

< 

m 

COI>l>^(Nf-ifCt^C0005 


00 

0-^0>CO»0000000»OM 


-H      O      10 

00   >c   Tt< 


10  >o  o  »c 

■^      T)H      O     ■<* 


o 


,— V 

0 

-0 

-o 

<o 

c 

M 

Ci 

ee 

0 

0 

3 

c 

_« 

^ 

-2 

s. 

-1-^ 

-»^ 

■k^ 

-u 

a 

a 

a 

a 

0 

0 

0 

0 

-2 

® 

V 

0) 

"o 

"o 

0 

0 

0 

0 

0 

Q 

S     c3 


O 


170 


1.    MALONATE 


P5 


a 

Tt< 

05 

■? 

■* 

05 

Oi 

i-H 

r— 1 

^"^ 

OJ 

^ 

■-3 

J 

§ 

Oi    -^ 


w  o 


c3    X!  ^ 

.£   r-  t- 

M   IC  >0 

'O  ri  rl 

H-)    ^  ^ 

-§    ^  ^ 


-5    ,«    ,::i; 
S      cS      cS 

1:15   PH   fe 


a  2 


P^  Q 


o  00  00  o  10 

10    00    I-H    ffO    10 


iCTj<iO00O5f0l>iO 


000^000 


iC   o 


O    >C    lO  >0    Tj< 

t^      TjH      -^  ■<*      CO 


10  ira  o 

TjH     O     l> 


flH 


o    o 
o    o 

Pi    05 


o    o 
o    o 


g    oj    o 


til     a     P,     ^    -72     ^ 
03    pR    O  W    CB 


pq 


^ 


EFFECTS    ON   THE    ENDOGENOUS    RESPIRATION  171 


^ 

3 

ert 

-tJ 

C/J 

<N 

'O 

cS 

^^ 

C 

^ 

c 

01 

^3 

c 

c 

O) 


<M    O    •>*     CO    O    O    C^ 


00«0->*0     »C»CiOiO'0>OlOOOOOOOOOOOOCO-*CDTf<0 
•^    Oi    -^     ^  »ClO»OOiClCiMlOOOO  CO  M 

—1  -H     CO     GO 


»CO»CiOOOiCOOOOOOO  »0  CO  CO  o 

eO'*'*>CCOl:-»0-<*CO->*OCOI>>C  -*  »C  CD  ;o 


OC  Coo  O  S  rf 


o  •£  'I 

i  g  I  ^ 


172 


1.    MALONATE 


m 


2?     <» 


M 


P. 

a 

T3 

-d 

< 

< 

i 

G 
c6 

T3 

T3 

bl 

b 

C 

C 

<u 

0) 

c3 

cS 

M 

M 

Lh 

(H 

>^ 

>> 

<D 

4) 

-2 

(D 

-Q 

,0 

"t^ 

" 

s 

g 

ce 

C8 

o 

O 

K 

ffi 

tf 

tf 

s  -^ 


(M     C<1     O     O     O 
Ttt    Tt    »0    lO    lO 


-£3    -a 


.2  S  .S 


PL, 


§      >>     3      S     g 

P?  S  H  H  1-^ 


J< 


a 

CO 


to 

3 

S 

M 

i< 

eS 

S 

fe 

o 

e 

S 

« 

r^ 

^ 

o 

^i:) 

o 

c3 

c3 

^ 

+3 

o 

X5 
O 

pq 

(X, 

H 

EFFECTS    ON    THE    ENDOGENOUS    RESPIRATION 


173 


W5 

OS  r\ 


m         ^    —H 


o 

s 


>■    o 


05   eg 


^    t^    1-    ^ 

1^    O     O     Ci 
0^0     05     — ' 


05     1— I 


'O       43 

(^     Pi 


■>* 

b 

o> 

« 

>_^ 

c 

3 

S 

X 

m    is     S     =* 


?►.    O     O     «      MO     X^^ 

a5fiQW<:!i!a50 


o^ 


tf 


05 


T3 
05 


o 


CCi    Tfi     00    <M    O    Oi    »C 
^H    ^     -^    ^H    CO    CC    "^ 


O     CO    00    O    -M     O     O 


.s  .s  .s 

'-3   '-3    ^3 
CO  M    OQ 


O    O     -H     U5 

lO   lO    i>    ■* 


JC    '^     '^     '^ 
t^    CD     CC     t^ 


o  o 


m     xn     " 


O     O 


o    o    c,   "5 

05  P5  fe  (1^ 


s 

s. 

c 

c 

C 

C 

C 

C 

C 

j2 

J^ 

_3 

cS 

_o 

o 

_o 

o 

o 

o 

.2 

_o 

.2 

C 

'ro 

■^ 

'S 

'm 

'm 

■^ 

"S 

■^ 

"« 

O 

« 

C 

c 

c 

fl 

c 

c 

c 

c 

© 

S 

<» 

(D 

« 

(B 

s 

0 

« 

<» 

© 

& 

& 

c 

ft 

& 

ft 

ft 

ft 

ft 

^ 

"o 

m 

n 

OJ 

OJ 

CO 

tc 

05 

02 

tc 

o 

5 

3 

3 

3 

3 

3 

3 

3 

s 

> 

m 

a; 

CO 

02 

CB 

02 

02 

02 

02 

> 

Oh 


W 


a? 


a  S 

ft    o 

<i1  03 


00 

00 

'S. 

•^ 
-< 

=0 

Q 

B 

S£ 

S 

p 

OQ    &i 


_s    2? 


*>.        *s.        \,«i        -o        (O        -_,- 

E-1    ^    ft;    Oh    o    S 


^     «      © 


rfi     cS      00 


K3 


-  t  ° 

cS      ?5     e8 


01       B      Oi 


174 


1.    MALONATE 


lO  2 


tf 


o 

05 

O 

P5 

o 

P3 

t3 
C 

C 

T3 

c 

Ph 


SU 

r] 

lO 

" — 

O 

05 

_ 

■5 

a 

^ 

-*:i 

2 

03 

'$ 

2 

5 
3 

c 

-O 

^ 

05 

Oi 

s 

"o 

O 

•«* 

-* 

Oi 

05 

t3 

t3 

I-H 

fi 

C 

■ — ■ 

^-' 

a 

cS 

-o 

73 

>i 

c 

c 

C 

« 

o 

-2 

-S 

L4 

h 

"S 

"3 

-^ 

o 

o 

'TS 

w 

X 


«0      T(<      Tt* 

t>   I>   t^ 


>0    LO    ^  <N 

i>  i>  00        00 


Ph 


i  i  i  I 


o    ^    ^ 


N    .5      CD      2      lU      3 

6  §  W  ^  W  S 


o 

C 

'm 

4J 

c 

M 

ID 

O 

S 

oo 

o 

^ 

WJ 

M 

M 

c 

bC 

M 

cS 

W 

w 

§ 

f»H 


•c* 

S 

!^ 

to 

<U 

e 

«0 

V 

e 

»o 

^ 

"to 

J 

-S 

5s 

.e 

"e 

^_^ 

e 

^^.^ 

<a 

^— ^ 

'ft. 

'•^ 

§> 

® 

s 

(D 

?^ 

CD 

«0 

Q 

00 

t3 
S 

s 
s 

c 

T3 
S 

S 

1 

o 

1 

'ft, 

1 

1 

^ 

^"^ 

fe; 

(^ 

"^"^ 

'^ 

^ 

&5 

CQ 


O 


— .    CI 


O    e  •= 


EFFECTS    ON   THE    ENDOGENOUS    RESPIRATION  175 


O     ^     C5 

•^   n   n 

Oi    Oi     Oi 


2  _  H^ir^i^^J^s^'  -25 


5        >^ 


N       3 


:  i 


^  >>  3      C      C 


Me'iOitNTti-^ccaor-CLoao-^'i^ioOfcOt^^ioo 


OOOOO-^OOOOOO-^-^OOO 
-Hrt-*OOfC'MTtiO    —    O-^'N'MOO-H 


5 


^  -^ 


c 


S 

cS 

bi 

X! 

2 

g 

c3 

43 

<u 

bo 

O 

o 

g 

1 

o 

M 

>> 

ei 

« 

_o 

t3 


^ 

M 

'S 

tc 

o 

'£ 

O 

o 

o 

C/3 

S 

o 

^ 

■^ 

>> 

c« 

_» 

^ 

^ 

Lri 

^ 

'3 

'S 

O 

to 

3 

73 

2 

o 

CO 

3 

WW  ^p:S  tiio  WWfqpq^SSS 


I  I   g  § 

;^     I  a    5     i 

C     =c     C  .S    -2    -^ 


S     i; 


S    -^ 


6    t:  ]B  .3'  5 


176 


1.    MALONATE 


P5 


w 


n 

to 

05 

^— V 

i-H 

l~- 

^—^ 

lO 

<35 

s? 

^ 

^^ 

s 

"e 

-d 

.^ 

c 

<u 

cS 

.^ 

^ 

c3 

cS 

3 

s 

w 


to  Ji 

OJ  o 

-H  > 

^  o 

-*^  0-1 


^  a 

lO    W    CQ 


0)      CO 


£-2   2 

l-J  cc  c» 


-^  73  M 

t-  ti  Sh 

cS  <S  3 

J  h-l  pq 


«  o  o  o  o 


a 

02 


1     o 
o    -* 


K 

a 


<o 

f—i 

o 

_o 

3 

_o 

"m 

"oo 

3 

bo 

C 

3 
3 

cS 

'S 

t-i 

PQ 

m 

3  ^ 

8   8   :>.  §   §   I        g  .£  .£ 

CO       GO       w       fefl    ^7     TT!  ""^       '-"       '~> 


o    > 


««     y    .S     3 


pq  pa  g  "-  J  J 


a;     4) 


3    ^     0?    -3    S 


3     3     S    ^    ^     £"    p 


c   P   c 


cc  cc  cS 


^^ 


w 


a 

CO 


EFFECTS    ON    THE    ENDOGENOUS    RESPIRATION  177 


TjH  2 


^  cd   c^  ^ 


cS 


cS 


^^      >-*      -^         r^        1-^        /•■\        I--1        (•-!         r:)         r^         fi         r-i        ,-^  ~.'       ~.*       Q 


3     3 


— . 

o 

3 

05 

>. 

C 

_ 

c 

o 

o 

>.  s 

C 

T3 

T3 

CO 

o 

C 

0) 

73 

Oi 

3 
O 

C     1* 

a 

cS 

T1 

cS      O 

« 

Ph 

|-^ 

W 

Q  W 

WKKpqpqpqpqpqpqpqpqpqm    -^-SWp^oS    Go 

COCD^OOO>OOCC'000      iCOiMOOiC      00-<*-^OCO'*fO'MOO 
(M-HCOiMQOOSOOt^CO'-fOCQDiO      »OfOMTt<»OfO      lO  00Or-<M-HC0'*>0 


t^    t^    t^    o    o 

-^     -^     -H     O     O 


CO  eo  C5     I  ■* 

t-    C^    l>  o 


c  .z  ■-.    ?  •- 


^ 

.2 

c8 

c 

^TO 

XH 

'S 

"S 

'E 

'C 

o 

IS 

L4 

u 

73 

o 

2 

§ 

'S 

-2 

3 

s 

Xi 

M 
3 
_3 

S 
o 

•*3 

3 
O 

o 
O 

o 
00 

3 

S 

o 

JD 

^ 

^ 

(h 

^ 

o 

o 

O 

s 

3 

0) 

00 

> 

o 
3 

3 
3 

x> 
S 

a 

J2 

s 

'S 

o 

2 

-2 

•--      r      5;    •■-      S;      ^      H      <D      y      2^  X!    -Q    J2      -3      ni      C3    .ii    — H  .3    .3    .3    .3 


178 


1.    MALONATE 


P5 


1^ 

cs    -3- 


PLh 


CO  Ci 

C5     -H 


<o 

.5 

05 

■*^ 

'(C 

-* 

o 

73 

05 

73 

C 

S 

c 

eS 

^^ 

(S 

« 

73 

-Q 

•^    o 
P5  T  '2 


M  T3  ii^     ^  S    ^    - 


CLh 


c5     g:i 


Oi 

rj 

O) 

TjH 

Ol 

O 

00   "-' 

fO 

o 

o 

"m 

■ — -  -u 

05 

73 

CO 

o 

C 

>>    PL, 

>. 

eS 

73 

o 

1^  "^ 

o 
i:i5 

£? 

73     c3 

73 

0) 

73 

C     «^ 

c 

C 

i:« 

2     0) 

en 

c 

c3 

03     ^ 

t« 

i^s 

o3 

Q    P^ 

Q  fc^ 

W 

Tt<   c-1   CO 

■^    »0    00 


—    C<l'>^<MO00OC0C0»Ct^O000000C30O 

c^oo-^^ooc^t^ooccioecoo-^ooooT^os 


«      IB 


i§ 

-S 

cS 

cS 

ce 

c 

C 

c 

m 

0^ 

o 

M 

M 

03 

OO 

o 

O 

8 

o 

O 

o 

<B 

_C 

S 

c 

o 

S 

"So 

'x 

's 

o 

^ 

^*^ 

^^ 

>) 

>. 

>. 

>■. 

4^ 

.JJ 

+:j 

<D 

a> 

QJ 

« 

;4 

;-! 

c 

c 

cS 
« 

03 

rs 

2 

73 

-0 

o 


HH     )-H     hH  k>      ky*  ls>     Vv< 

HH     HH     H-t  HH     hH  t-M     HH 


EFFECTS    ON   THE    ENDOGENOUS    RESPIRATION 


179 


CO   CO 

CO    CO 
05     05 


2   o 


-H      o 


pq  ^ 


O    -H    O    C    CO    CD    O 


ce    2 


CO      ?      S 


QQ 

s 

O 

ti 

;i3 

GQ 

s 

cS 

2 

2 

S 

£2 

C3 

o 

o 

o 

CJ 
c8 

o 

o 

tl 

L. 

cS 

rt 

cc 

cc 

tc 

cS 

^ 

C 

TS 

» 

<D 

'^ 

OJ 

OQ 

0) 

C 
4) 

CO    uo    73 
2    ^     eS 


IC    lO    lO    -H    ic 

CO  CO  CO  lo  CO 

C^    Oi    O    Oi    Ot! 


03 

CD 

cS 

-O 

T3 

S 

S 

M 

o 

^ 

eg 

eg 

eg 

eg 

eg 

.2f 

-4-^ 

+2 

">> 

>> 

C 

C 

C 

'E 

"3 

^ 

O 

O 

eg 

C 

C 

fQ 

^ 

h-1 

pq 

m 

H 

<1 

<: 

~  ,2 


tf 


eg 


eg 
H 


<»       03 


eg  eg 

>■.  >. 

O  O 

CO  03 


^O^^'C—     "     C     C     O 


K  pq  <5  ffi  ffi 


H  P^        d  ^  ^  ^ 


lO   O    lO   o   o    o   o 
C<I    ^  O    O    <M    Oq 


•<*    Tjt    CO    Tt<    (M    -^    05 

i>   t^   t^   t>   i>   r-   CO 


C      eo  U 

O     <0    .-,    -^    -^ 


Q    a>  .-.  .2 


2    ^ 


_^  O  K       03     ^ 

"^  '^  >■■>>>■. 

^  ^  QJ       2       05 

^  fl^  rS       -S       -S 

>  >  T3    ^    T3 

;j  ;3  S  5  i5 


O      03      "3    .X    ;:;      O 


o  -d   .ii   -X 


^    S   ^    =« 


^   -35     Q     =?     ?? 


s  :g 


05  fe  M  <3  fiH 


P     eg 


PM 


"B 

K 

.4J 

& 

-2 

<B 

X 

-j3 

-^ 

-bS 

0 

S-i 

C 

a 

£    S 


y.  H 


li    T3      OJ 


«    s 


T3 


r"      ^,      ■" 
S      <5^      3 


^-     O    J2 


bC 

c 

0 

eg 

T3 

•+i 

03 

5 

_>■. 

0 

3 

c 

t-i 

-t^J 

_c 

03 

4^ 

c 
0 

3 
,0 

1 

4i 
03 

ft 

03 

«       C       05 


c 

eg 

-Q 

'g 

^ 

r^ 

05 

eg 

05 

> 

-t^ 

05 

S 

'Si 

5 

<» 
c;> 

<» 
45 

_3 

05 

3 
0 

"3) 

> 
eg 

J3 

5£ 

oT 

^ 

03 

>> 

05 
0 

eg 

^ 

c 

eg 
4-:> 

_C 

T3 

03 

45 

c 

8 

T5 

05 

05 

73 

r^ 

w 

c 

r-.  JC>  05  03 
"  'S  05  S 
'^     •"     -O      01 


180  1.    MALONATE 

correct  for  this  phenomenon  is  to  determine  the  C^^02  formed  from  labeled 
malonate. 

(f)  Nature  of  the  cycle  operation  and  the  presence  of  alternate  pathways. 
The  inhibition  of  the  oxygen  uptake  associated  with  the  cycle  will  depend 
on  a  number  of  factors  in  addition  to  the  inhibition  of  succinate  dehydro- 
genase. The  availability  of  a  large  pool  of  organic  acids  to  form  oxalacetate, 
or  the  presence  of  pathways  from  which  oxalacetate  may  arise  (e.g.,  by 
carboxylation  of  pyruvate,  or  from  aspartate  by  transamination),  will  re- 
duce the  inhibition  of  oxygen  uptake  from  that  which  would  be  observed 
if  all  the  oxalacetate  had  to  be  derived  from  the  cycle.  Other  pathways 
for  the  metabolism  of  succinate  may  circumvent  the  block  to  some  extent. 
These  matters  have  been  discussed  in  some  detail  (see  pages  72-88). 

(g)  Adaptive  changes  in  the  presence  of  malonate.  Inhibition  of  the  cycle 
may  accelerate  other  pathways.  The  increased  uptake  and  metabolism  of 
glucose  brought  about  by  malonate  have  been  noted  in  several  types  of 
cells,  and  such  a  phenomenon  will  tend  to  counteract  the  malonate  inhi- 
bition on  the  oxygen  uptake.  Adaptive  changes  in  enzyme  concentrations 
probably  are  seldom  important  in  short-term  experiments  but  cannot  be 
completely  ignored  in  work  with  certain  microorganisms.  Inhibitions  by 
malonate  have  occasionally  been  noted  to  decrease  with  time,  and  adaptive 
changes  are  the  most  obvious  explanation. 

These  and  other  more  subtle  factors  determine  the  effect  of  malonate  on 
the  total  oxygen  uptake  of  a  preparation,  and  it  should  be  apparent  that 
deductions  based  exclusively  on  the  inhibitions  of  endogenous  respiration 
are  frequently  untenable.  A  definite  inhibition  with  a  reasonable  malonate 
concentration  is  more  significant  than  a  negative  result,  because  there  are 
many  factors  which  can  reduce  or  abolish  the  action  of  malonate  even 
though  the  cycle  is  present  and  active. 

The  Time  Course  of  Malonate  Inhibition 

The  inhibition  of  respiration  by  malonate  may  occur  fairly  rapidly  and 
remain  constant,  or  it  may  increase  slowly  to  a  level  at  which  it  is  main- 
tained, or  it  may  gradually  disappear,  or  it  may  vary  in  quite  complex 
fashion  with  time.  A  slowly  developing  inhibition  would  not  be  unexpected 
with  a  substance  which  does  not  penetrate  readily.  The  inhibition  of  suc- 
cinate dehydrogenase  is  essentially  instantaneous  so  that  an  approximately 
linear  increase  in  the  inhibition  to  a  constant  level  would  imply  that  the 
rate  of  inhibition  is  determined  by  the  penetration.  On  the  other  hand, 
secondary  effects,  such  as  would  result  from  the  depletion  of  ATP,  may 
also  contribute  to  a  progressive  inhibition.  Greville  (1936)  noted  that  mal- 
onate does  not  immediately  inhibit  the  respiration  of  rat  diaphragm  but 


EFFECTS    ON    THE    ENDOGENOUS    RESPIRATION  181 

that  the  inhibition  developes  over  1-2  hr.  A  very  similar  time  course  was 
observed  in  barley  roots  by  Laties  (1949  a).  After  the  addition  of  10  mM 
malonate,  the  respiration  drops  linearly  for  60-90  min  and  then  becomes 
relatively  constant  at  about  40%.  However,  over  the  next  5  hr,  the  inhi- 
bition lessens  somewhat.  This  was  postulated  to  be  due  to  increasing  suc- 
cinate concentration,  but  actually  such  changes  in  succinate  take  place 
much  more  rapidly  in  most  cases.  It  must  be  admitted  that  in  spinach 
leaves  the  maximal  succinate  accumulation  occurs  at  4  hr  (Laties,  1949  b), 
so  that  this  explanation  can  by  no  means  be  eliminated.  The  inhibition 
of  the  spinach  leaf  respiration  by  malonate  is  greater  at  6  hr  than  at  3  hr 
but  not  enough  data  are  available  to  correlate  the  changes  in  inhibition 
with  succinate  levels.  Malonate  requires  about  1  hr  to  produce  its  maximal 
inhibition  of  sea-urchin  egg  homogenate  respiration  (Yeas,  1950).  This  is 
the  only  report  of  such  a  slowly  developing  inhibition  in  subcellular  pre- 
parations and  no  explanation  is  evident.  Another  situation  seems  to  exist 
in  bovine  kidney  culture  cells,  where  the  inhibition  by  100  mM  malonate 
slowly  increases  from  23%  at  1  hr  to  35%  at  5  hr  (Polatnick  and  Bachrach, 
1960).  It  is  more  likely  here  that  secondary  changes  are  responsible. 

Definite  decrease  in  the  respiratory  inhibition  of  pigeon  brain  dispersions 
(probably  similar  to  homogenates)  with  time  was  reported  by  Banga  et  al. 
(1939).  The  inhibition  is  41.7%  at  10  min,  29.5%  at  30  min,  and  20.5% 
at  50  min.  Since  the  malonate  concentration  was  24  mM,  it  seems  unlikely 
that  metabolism  of  malonate  could  have  reduced  its  concentration  signifi- 
cantly. Adaptive  changes  in  homogenates  are  improbable  and  sufficient 
accumulation  of  succinate  from  the  relatively  limited  substrate  supply  is 
scarcely  possible.  Besides,  the  inhibition  of  pyruvate  oxidation  increased 
over  this  interval.  Sometimes  changes  in  tissue  metabolism  occur  during 
the  course  of  an  experiment  independent! j^  of  malonate  action.  When  mal- 
onate is  added  to  human  brain  slices  immediately,  the  endogenous  respira- 
tion is  inhibited  around  25%  at  5  mM,  but  if  malonate  is  added  after 
90  min  incubation  of  the  slices,  there  is  no  inhibition  (Elliott  and  Suther- 
land, 1952).  The  role  of  succinate  oxidase  in  the  respiration  must  change 
as  a  result  of  the  slicing  or  the  abnormal  medium. 

The  inhibition  of  the  respiration  of  rat  ventricle  slices  by  malonate 
5-20  mM  follows  a  more  complex  course  (Webb  et  al.,  1949).  Following 
an  initial  inhibition,  the  respiration  rises  for  approximately  1  hr  and  then 
begins  to  fall  again.  There  is  thus  a  maximum  or  hump  in  the  respiration 
curve.  After  1  hr,  the  respiratory  level  with  malonate  is  higher  than  in 
the  controls.  The  reversal  of  the  inhibition  is  inhibited  by  fluoride,  which 
would  indicate  that  the  hump  is  due  to  augmented  glucose  oxidation  or  to 
the  metabolism  of  malonate.  Calcium  is  also  necessary  for  the  typical  re- 
sponse to  malonate.  Very  complex  effects  of  malonate  were  also  found  by 
Turner  and  Hanly  (1947)  and  Hanly  et  al.  (1952)  in  carrot  slices.  The  var- 


182  1.    MALONATE 

iation  of  the  inhibition  depends  on  the  pH.  At  pH  4,  the  inhibition  de- 
velops over  1  hr  and  remains  constant,  but  at  higher  pH's  the  inhibition 
may  disappear  or  only  stimulation  may  be  seen.  The  value  of  these  interest- 
ing experiments  is  greatly  reduced  by  the  unaccountable  use  of  potassium 
malonate  rather  than  the  sodium  salt.  Potassium  at  50  mM  (which  is  the 
concentration  of  malonate  generally  used  by  them)  stimulates  the  respira- 
tion and  alters  its  character,  so  that  aU  the  results  must  be  the  summation 
of  two  usually  opposing  actions.  This  illustrates  how  the  incorrect  choice 
of  an  inhibitor  salt  can  vitiate  the  results  of  an  otherwise  excellent  in- 
vestigation. 

Most  of  the  work  on  the  malonate  inhibition  of  endogenous  respiration 
has  been  done  without  regard  for  possible  alterations  in  the  inhibition 
with  time.  The  inhibitions  have  simply  been  determined  over  an  arbitrary 
interval.  Inasmuch  as  changes  in  the  inhibition  by  malonate  occur  quite 
frequently,  it  is  likely  that  over  all  inhibitions,  such  as  are  presented  in 
Table  1-26,  are  often  mean  values  and  do  not  reflect  either  the  initial  inhi- 
bition or  the  maximal  inhibition.  As  was  pointed  out  in  Chapter  1-12,  the 
value  of  many  studies  on  inhibitors  would  be  increased  by  determinations 
of  the  variation  of  the  inhibitions  with  time. 

Effects  of  Different  Conditions  on  the  Inhibition  of  Endogenous  Respiration 

One  of  the  most  important  variables  affecting  the  response  of  the  endog- 
enous respiration  to  malonate  is  the  age  of  the  tissue,  particularly  as  it 
relates  to  the  stage  of  development  or  the  interval  between  the  preparation 
of  the  tissue  and  the  experimental  testing.  The  respiration  of  plant  tissues 
usually  becomes  more  sensitive  to  malonate  with  time.  This  indicates  a 
progressive  change  in  the  metabolic  pattern  in  the  direction  of  a  greater 
participation  of  the  cycle.  The  changes  in  the  inhibition  during  malonate 
inhibition,  discussed  in  the  previous  section,  can  be  due  to  the  effects  of 
the  malonate  or  to  inherent  metabolic  alterations.  It  is  thus  important  in 
such  studies  to  determine  both  the  changing  inhibition  in  the  presence  of 
malonate  and  the  changing  susceptibility  as  malonate  is  added  at  various 
intervals.  The  malonate  inhibition  rises  with  time  in  the  Avena  coleoptile 
(Bonner,  1948),  rose  petals  (Siegelman  et  al.,  1958),  chicory  root  slices  (La- 
ties,  1959  a),  Arum  spadix  slices  (Simon,  1959),  and  potato  tuber  slices 
(Romberger  and  Norton,  1961).  These  changes  are  usually  associated  with 
an  increase  in  the  total  uninhibited  respiration.  For  example,  in  potato 
slices  there  is  a  4-fold  rise  in  the  respiration  during  incubation  for  30  hr; 
the  malonate  resistant  fraction  doubles  and  the  malonate-sensitive  fraction 
increases  10-  to  15-fold.  Carrot  slice  inhibition  by  malonate,  on  the  other 
hand,  decreases  steadily  up  to  376  hr  after  cutting  the  sections  (Hanly 
et  al.,  1952),  although  the  uninhibited  respiration  first  rises  and  then  falls. 
The  results  obtained  with  animal  tissues  are  less  striking  and  more  variable. 


EFFECTS    ON   THE    ENDOGENOUS    RESPIRATION  183 

The  inhibition  of  rabbit  ova  respiration  at  various  times  postcoitum  does 
not  change  significantly  (Fridhandler  et  al.,  1957),  while  the  inhibition  of 
trematode  respiration  decreases  with  time  from  excystment  (Vernberg  and 
Hunter,  1960).  When  20  raM  malonate  is  added  to  rat  ventricle  slices  1  hr 
after  slicing,  the  inhibition  of  the  respiration  is  only  25%,  whereas  initially 
the  inhibition  is  near  50%  (Webb  et  al.,  1949),  and  this  relationship  holds 
for  all  malonate  concentrations  up  to  100  vaM.  We  have  seen  that  in  the 
presence  of  malonate  the  inhibition  has  been  replaced  by  stimulation  at 
1  hr.  Therefore  the  metabolism  changes  differently  during  the  action  of 
malonate  and  the  maxima  in  the  time  curves  cannot  be  explained  by 
inherent  alterations  of  the  metabolic  pattern.  All  of  these  results  point  to 
the  importance  of  considering  the  time  factor  in  studies  of  malonate  inhi- 
bition. 

Another  apparently  important  factor  is  the  ion  and  buffer  composition  of 
the  medium,  although  no  thorough  studies  have  been  done  and  the  mecha- 
nisms are  not  understood.  Many  years  ago  Annau  (1935)  observed  that  the 
inhibition  of  the  respiration  of  both  rabbit  liver  and  kidney  slices  is  less 
in  Ringer  than  in  phosphate  medium.  The  results  were  quite  variable  but 
on  the  average  the  inhibition  is  30%  in  phosphate  and  15-20%  in  Ringer 
medium.  Unfortunately,  Annau  did  not  state  what  form  of  malonate  was 
used  nor  did  he  mention  pH  control,  so  the  results  are  perhaps  unreliable. 
The  presence  of  bicarbonate  abolishes  the  inhibition  by  malonate  in  ox 
retina  homogenates  (Burgess  et  al.,  1960),  and  it  is  probable  that  inhibition 
in  intact  cells  would  also  vary  with  the  bicarbonate  concentration.  Bicarbon- 
ate can,  of  course,  facilitate  the  formation  of  oxalacetate  through  carboxyla- 
tion  reactions.  When  Ca++,  Mg++,  or  K"*"  is  removed  from  the  medium,  the 
inhibition  of  rat  brain  slice  respiration  by  10  mM  malonate  is  not  altered, 
but  in  the  presence  of  fumarate  the  inhibition  becomes  progressively  greater 
as  these  ions  are  successively  removed  (Greville,  1936).  The  addition  of 
Ca+"'"  to  nematode  minces  increases  both  the  respiratory  inhibition  and  the 
inhibition  of  succinate  oxidation  (Massey  and  Rogers,  1950).  The  sensitivity 
of  chicory  root  slice  respiration  to  malonate  is  markedly  affected  by  K+  and 
Li+  (Laties,  1959  b).  Slices  incubated  with  50  milf  K+  are  inhibited  more 
and  with  50  mM  Li+  inhibited  less  than  the  fresh  slices.  It  is  also  interesting 
that  increase  in  CO2  tension  results  in  progressive  disappearance  of  the  mal- 
onate inhibition,  whereas  increase  in  Og  tension  augments  the  malonate- 
sensitive  fraction  of  the  respiration.  It  is  thus  clear  that  the  medium  can 
play  an  important  role  in  the  response  to  malonate.  Much  work  has  been 
done  in  quite  nonphysiological  media  and  the  results  are  thus  difficult  to 
apply  to  the  actions  of  malonate  in  situ.  Much  more  effort  should  be  di- 
rected at  creating  approximately  physiological  conditions. 

The  functional  activity  of  the  tissue  determines  the  level  and  type  of 
respiration,  and  therefore  is  often  a  major  factor  in  the  sensitivity  to  mal- 


184  1.    MALONATE 

onate.  This  has  been  shown  particularly  clearly  in  brain  slices,  stimulated 
both  electrically  (Heald,  1953)  and  by  K+  (Kimura  and  Niwa,  1953;  Yoshida 
and  Quastel,  1962).  The  stimulated  respiration  is  readily  inhibited  by  mal- 
onate  (Fig.  1-14)  whereas  the  resting  respiration  is  insensitive.  This  be- 
havior is  probably  exhibited  by  many  tissues.  It  is  often  very  difficult  to 
determine  exactly  the  functional  state  of  isolated  tissues,  such  as  slices, 
but  where  possible  this  should  be  attempted.  We  shall  find  later  that 
active  tissues  are  more  easily  functionally  depressed  by  malonate  and 
this  may  have  a  metabolic  basis.  Indoleacetate  stimulates  the  growth  of 
Avena  coleoptiles  and  increases  the  respiration  simultaneously.  This  ad- 
ditional respiration  brought  about  by  indoleacetate  is  readily  inhibited  by 
malonate  (Bonner,  1949),  and  it  is  likely  that  the  respiration  of  rapidly 
growing  tissue  is  generally  inhibited  more  strongly  by  malonate  than  that 
of  resting  or  slowly  proliferating  tissue. 

Consideration  must  also  be  given  to  the  history  of  the  tissue.  Bonner  (1948) 
has  shown  that  the  nutritional  state  of  the  Avena  coleoptile  determines  the 
inhibition  by  malonate,  and  it  is  probable  that  the  same  applies  to  animal 
tissues.  The  inhibition  of  wheat  seedling  respiration  by  malonate  depends 
on  a  number  of  factors,  including  the  type  and  duration  of  irradiation, 
the  nutrition,  and  the  region  from  which  the  plants  come  (Farkas  et  al., 
1957  a,  b).  This  is  a  field  that  has  been  very  little  explored.  The  changes 
in  the  respiratory  inhibition  of  animal  tissues  with  nutrition  might  not  only 
provide  information  on  the  metabolic  patterns  under  various  conditions, 
but  be  important  in  the  use  of  the  inhibitor  to  selectively  depress  the  me- 
tabolism and  growth  of  neoplastic  tissues. 

Effects  on  the  Respiratory  Quotient 

The  effects  of  an  inhibitor  on  the  respiratory  quotient  (R.Q.  =  COg  form- 
ed/Og  uptake)  are  often  indicative  of  shifts  in  metabolic  pathways.  Let  us 
first  consider  the  theoretical  values  of  the  R.Q.  for  the  metabolism  of 
various  substrates  (see  tabulation)  in  the  presence  and  absence  of  malonate, 
assuming  that  malonate  is  able  to  block  succinate  oxidation  completely. 
Cases  in  which  oxalacetate  is  formed  in  the  cycle  and  from  noncycle  sources 
must  be  separated.  Summarizing  these  results,  one  would  expect  malonate 
to  increase  or  decrease  the  R.Q.,  depending  on  the  substrate  and  the  nature 
of  the  cycle  operation.  Since  a  complete  block  of  succinate  oxidation  would 
prevent  the  formation  of  oxalacetate  through  the  cycle,  malonate  may 
shift  the  pathway  from  cycle  oxalacetate  to  externally  formed  oxalacetate, 
if  the  latter  reaction  is  possible.  If  this  is  so,  the  R.Q.  should  rise  in  every 
case. 

This  prediction  is  quite  consistently  borne  out  experimentally.  The  R.Q. 
of  rat  liver  slices  rises  from  0.72  to  0.77  in  the  presence  of  20  mM  malonate, 
at  which  concentration  the  respiration  is  inhibited  14%  (Elliott  and  Greig, 


EFFECTS    ON    THE    ENDOGENOUS    RESPIRATION  185 


Substrate  metabolism  R.Q. 


Glucose  ->  CO2  +  H2O  1.0 

Glucose  ->  succinate  -f  CO2  +  H^O  0.8 

Glucose  +  2  oxalacetate  ->•  CO2  +  HgO  1.27 

Glucose  +  2  oxalacetate  ->  2  succinate  +  CO2  +  H2O  1.5 

Pyruvate  ->  COj  +  HjO  1.2 

2  PjTuvate  ->  succinate  +  CO2  +  H2O  1 .  33 

Pyruvate  +  oxalacetate  ->  COj  +  H2O  1.4 

Pyruvate  +  oxalacetate  ->■  succinate  +  COj  +  O2  2.0 

ButjTate  -^  CO2  +  H2O  0.8 

ButjTate  -r  oxalacetate  ->  COj  +  HjO  1 .  23 

Butyrate  +  oxalacetate  ->  succinate  +  CO2  +  HjO  1.0 

Butyrate  +  2  oxalacetate  ->  CO2  +  HjO  1.2 

But>Tate  +  2  oxalacetate  ->  2  succinate  +  CO2  +  H2O  1.33 


1937).  However,  malonate  decreases  the  R.Q.  of  kidney  slices,  both  en- 
dogenous and  with  pyruvate  as  the  substrate.  Malonate  elevates  the  R.Q. 
of  rat  adipose  tissue  from  1.0  to  1.13  endogenously  and  from  1.14  to  1.41 
in  the  presence  of  glucose  (Haugaard  and  Marsh,  1952).  In  frog  muscle, 
the  R.Q.  first  rises  from  0.9  to  0.97  at  10  mM  malonate,  but  then  progres- 
sively decreases  as  the  malonate  concentration  is  raised  so  that  at  200  mM 
malonate  the  R.Q.  is  0.39  (Thunberg,  1909).  In  plant  tissues,  the  effects 
are  less  variable.  Malonate  has  been  shown  to  increase  the  R.Q.  of  barley 
roots  from  0.97  to  1.14  (Machlis,  1944),  of  maize  roots  (Beevers,  1952), 
of  carrot  roots  up  to  values  as  high  as  3  (Hanly  et  al.,  1952),  of  chicory 
roots  from  1.03  to  1.14  (Laties,  1959  a),  and  of  rhubarb  leaves  at  pH  5.3 
(Morrison,  1950). 

Of  course,  there  are  many  factors  which  must  be  taken  into  account, 
since  malonate  can  secondarily  alter  several  metabolic  pathways.  A  stimula- 
tion of  glucose  uptake  could  change  the  R.Q.  in  either  direction,  depending 
on  the  nature  of  the  substrates  used  in  the  uninhibited  tissue;  in  the  pre- 
sence of  a  significant  cycle  block,  this  would  usually  depress  the  R.Q.  and 
counteract  the  more  direct  effects  described  above.  On  the  other  hand, 
metabolism  of  malonate  would  tend  to  elevate  the  R.Q.  since  the  complete 
oxidation  would  give  R.Q.'s  of  1.50-1.55  and  the  oxidation  to  succinate 
3.0-4.0.  A  final  factor  of  importance  is  the  relative  dependence  of  glucose 
and  fatty  acid  metabolism  on  the  operation  of  the  cycle  and  the  levels  of 
ATP,  since  malonate  could  alter  the  oxidative  contribution  from  these  sub- 
strates secondarily. 


186  1.    MALONATE 

Significance  of  Respiratory   Inhibition 

Does  the  degree  of  malonate  inhibition  indicate  the  contribution  of  the 
cycle  to  the  total  oxygen  uptake?  This  must  certainly  be  answered  in  the 
negative.  Lack  of  inhibition  can  be  due  to  a  failure  to  penetrate,  the  me- 
tabolism of  malonate,  a  source  of  oxalacetate  external  to  the  cycle,  met- 
abolic adaptations  of  the  cells,  and  many  other  factors.  Positive  evidence 
of  inhibition  is  more  valuable  than  absence  of  inhibition,  but  even  when 
definite  inhibition  is  observed  the  possibility  of  actions  other  than  in  the 
cycle  must  be  considered,  especially  when  the  malonate  concentration  must 
be  high  to  achieve  an  effect.  It  is  doubtful  if  anyone  examining  Table  1-26 
would  attempt  to  correlate  the  inhibitions  with  the  importance  of  the 
cycle  in  the  organisms  and  tissues.  For  example,  in  general  there  is  greater 
inhibition  of  mammalian  endogenous  respiration  than  of  the  respiration  of 
microorganisms  or  plants.  This  might  indicate  a  greater  role  of  the  cycle 
in  mammalian  tissues,  but  it  could  also  be  attributed  to  a  poorer  penetration 
in  the  plants  and  microorganisms,  or  to  a  greater  metabolic  flexibility  and 
adaptability  in  these  more  resistant  forms.  It  must  also  be  clear  that  the 
degree  of  respiratory  inhibition  bears  no  necessary  relationship  to  the  de- 
gree of  inhibition  of  succinate  dehydrogenase.  A  significant  inhibition  by  a 
reasonable  concentration  of  malonate  is  evidence  for  the  operation  of  the 
cycle,  but  the  quantitative  aspects  of  the  contribution  cannot  be  derived 
from  these  data  alone.  The  effects  of  malonate  on  the  endogenous  respira- 
tion are  sometimes  of  greater  physiological  significance  than  effects  on  the 
oxygen  uptake  in  the  presence  of  high  concentrations  of  often  abnormal 
substrates,  since  the  endogenous  metabolism  may  be  representative  of  a 
more  normal  balance  of  substrates.  In  this  connection,  studies  of  inhibitors 
would  often  be  improved  if  the  attempt  were  made  to  provide  the  cells 
with  a  mixture  of  physiologically  pertinent  substrates  at  the  concentrations 
normally  occurring  in  the  cellular  environment. 

PERMEABILITY  OF  CELLS  TO  MALONATE 

One  of  the  major  problems  in  the  use  of  malonate  has  always  been  the 
degree  of  penetration  of  the  inhibitor  into  the  cells  or  tissues,  and  it  has 
been  frequently  stated  that  this  is  the  primary  factor  responsible  for  the 
low  inhibitions  observed  in  many  cases.  It  is  true  that  the  plasma  mem- 
brane is  relatively  impermeable  to  most  ions,  particularly  anions  and  those 
carrying  two  or  more  charges,  but  if  this  is  so  how  can  one  explain  the 
marked  respiratory  stimulations  usually  seen  with  succinate  or  other  di- 
carboxylate  ions?  Furthermore,  malonate  is  often  metabolized  readily  by 
tissues  and  this  presupposes  entrance  into  the  cells.  Since  there  are  other 
possible  reasons  for  a  resistance  to  malonate,  the  permeability  hypothesis 
must  be  examined  critically. 


PERMEABILITY  OF  CELLS  TO  MALONATE  187 

Experimental  Evidence  Relating  to  the  Penetration  of  Malonate 

Malonic  acid  is  more  than  10  times  as  lethal  on  injection  into  frogs  as 
is  sodium  malonate  (Heymans,  1889).  No  explanation  was  offered  for  this 
observation  but  it  could  have  been  due  to  the  greater  permeability  of  the 
cells  to  the  acid  or,  on  the  other  hand,  to  a  nonspecific  acidification  of  the 
animals.  Malonate  administered  to  rabbits  circulates  initially  in  a  volume 
equivalent  to  the  extracellular  compartment  and  the  intracellular  transfer 
occurs  slowly  (Wick  et  al.,  1956).  The  failure  of  malonate  to  alter  the  me- 
tabolism of  labeled  acetate  was  attributed  to  both  the  slow  penetration  into 
the  tissues  and  the  simultaneous  metabolism  of  the  malonate,  both  factors 
keeping  the  intracellular  concentration  at  low  levels.  The  inability  of  mal- 
onate to  alter  gastric  acid  secretion  in  frogs,  even  at  lethal  doses,  was  sim- 
ilarly attributed  to  these  factors  (Davenport  and  Chavre,  1956).  Inasmuch 
as  succinate  oxidase  is  present  in  the  secretory  cells  and  the  cycle  is  im- 
portant in  secretion  (as  shown  by  the  inhibition  with  fluoroacetate),  the 
lack  of  action  must  be  due  to  an  insufficient  concentration  within  the  cells. 

Turning  to  isolated  tissues  and  cell  suspensions,  the  augmentation  of 
malonate  effects  by  procedures  designed  to  reduce  or  abolish  the  perme- 
ability barriers  has  been  demonstrated  many  times  (Table  1-26).  Rat  dia- 
phragm respiration  is  inhibited  slowly  by  malonate,  but  if  the  diaphragm 
is  cut  into  small  pieces,  and  hence  presumably  damaged,  the  inhibition  is 
immediate  (Greville,  1936).  The  oxygen  uptake  of  pigeon  brain  brei  is 
inhibited  rather  poorly  by  24  mM  malonate,  but  when  the  brain  is  dispersed 
more  completely  in  the  form  of  a  homogenate  the  inhibition  is  more  marked 
(Banga  et  al.,  1939).  The  succinate  dehydrogenase  of  yeast  incubated  for 
5  hr  in  liquid  nitrogen  is  much  more  susceptible  to  malonate  than  in  normal 
cells  (Lynen,  1943),  and  the  same  holds  for  E.  coli  treated  with  toluene 
(Ajl  and  Werkman,  1948).  Sensitivity  to  malonate  can  be  induced  by  liquid 
nitrogen  treatment  in  the  fungus  Zygorrhynchus  (Moses,  1955)  and  by  dry- 
ing Pseudomonas  (Gray,  1952).  Malonate  does  not  inhibit  glutamate  oxida- 
tion in  intact  cells  of  Pasteurella,  but  inhibits  well  in  sonic  lysates  (Kann 
and  Mills,  1955).  All  of  these  phenomena  have  been  interpreted  in  terms  of 
permeability.  This  is  certainly  the  most  obvious  explanation  and  it  is  prob- 
ably generally  correct,  but  it  must  be  admitted  that  such  drastic  treat- 
ments could  affect  many  other  things;  for  example,  alter  the  organized 
enzyme  structure  so  that  the  attacked  enzyme  is  more  exposed,  or  reduce 
the  ability  of  the  cells  to  metabolize  malonate. 

Only  one  investigation  of  the  relative  permeabilities  of  the  dicarboxy- 
late  anions  has  been  made.  Giebel  and  Passow  (1960)  determined  the  half- 
times  for  penetration  of  these  ions  into  bovine  erythrocytes  and  the  results 
are  given  in  Table  1-27.  Giebel  and  Passow  attempted  to  correlate  the  per- 
meabilities with  the  ionic  sizes  and  the  acidic  ionization  constants.  The 
ionic  volumes  and  lengths  presented  in  the  table,  which  are  somewhat 


188 


1.    MALONATE 


Table  1-27 

Erythrocyte  Permeabilities  and   Molecular  Properties   of  Dicarboxylate 

Anions  " 


Anion 


Relative 
permeability 


Ionic 
volume 


Ionic 
length 

(A) 


(H,B) 


(HB-) 


Oxalate 

100 

54 

5.1 

0.00015 

276 

Malonate 

17 

66 

7.5 

0.112 

6590 

Maleate 

14 

75 

7.9 

0.051 

252 

Fumarate 

1.4 

75 

7.6 

0.014 

445 

Succinate 

0.45 

78 

7.5 

2.56 

6410 

Malate 

0.31 

91 

7.5 

0.135 

1860 

Glutarate 

0.085 

90 

8.5 

1.99 

3780 

Tartrate 

0.045 

104 

7.5 

0.0103 

338 

"  The  permeabilities  were  determined  in  bovine  erythrocytes  by  Giebel  and  Passow 
(1960).  The  values  given  here  for  the  relative  permeabiHties  are  the  reciprocals  of  the 
entrance  rate  half-times  (ti/^)  multipHed  by  100.  Tlie  calculations  of  the  ionic  volumes 
and  lengths  are  approximate  and  usually  depend  on  the  configuration  of  the  ions. 
The  concentrations  of  HjB  and  HB    are  given  for  a  total  concentration  of  1  M. 


different  than  those  given  by  Giebel  and  Passow,  must  be  considered  as 
only  relative  values,  neglecting  hydration  and  special  configurations  of  the 
ions.  Their  experiments  were  run  at  a  pH  of  7.35  so  the  concentrations  of 
the  undissociated  and  singly  dissociated  forms  of  the  acids  are  given  for 
this  pH.  The  ionization  constants  are  sometimes  quite  different  from  those 
assumed  by  Giebel  and  Passow  and  are  those  in  Table  1-2.  There  is  certainly 
little  or  no  general  correlation  between  permeabilities  and  the  concentra- 
tions of  either  H2B  or  HB".  Succinate,  for  example,  penetrates  one-thirty 
eighth  as  rapidly  as  malonate  and  yet  (H2B)  for  succinate  is  23  times  higher 
than  for  malonate.  This  does  not  necessarily  invalidate  the  assumption 
that  for  a  single  substance  the  unionized  forms  penetrate  more  rapidly  than 
the  ionized,  but  it  shows  that  there  are  other  factors  which  are  quite  im- 
portant. There  is  also  no  correlation  with  the  ionic  length  and  it  is  unlikely 
that  one  would  be  expected.  However,  there  is  some  correlation  with  ionic 
volume,  leading  Giebel  and  Passow  to  suggest  that  the  dicarboxylate  anions 
penetrate  through  pores  in  the  membrane,  whereas  the  monocarboxylates 
pass  through  the  lipid  phase  of  the  membrane.  They  calculate  the  pore 
radius  to  be  between  3.8  and  4.5  A.  If  these  ions  pass  through  the  pore 
channels  in  the  extended  form,  which  is  likely,  there  are  two  major  factors 
which  may  contribute  to  the  permeability:  the  cross-sectional  area  perpen- 


PERMEABILITY  OF  CELLS  TO  MALONATE  189 

dicular  to  the  direction  of  passage  and  the  degree  of  interaction  of  the 
molecules  with  the  walls  of  the  pores.  This  interaction  may  be  of  various 
types  and  includes  steric  repulsion  and  attractive  forces  (such  as  van  der 
Waals'  forces  or  hydrogen  bonds).  The  configuration  of  the  ion  must  be 
important  in  this  connection.  Maleate  penetrates  10  times  faster  than  fu- 
marate  and  this  must  be  mainly  due  to  the  structure  of  maleate  wherein 
the  carboxylate  groups  are  much  closer  than  in  fumarate.  That  both  of 
these  ions  penetrate  more  rapidly  than  succinate  may  be  due  to  the  greater 
rigidity  of  the  former,  energy  perhaps  being  required  for  succinate  to  change 
from  its  statistically  most  probable  configuration  to  that  necessary  for 
penetration.  It  is  evident,  however,  that  none  of  these  explanations  satis- 
factorily fits  the  experimental  data  and  that  we  need  to  know  much  more 
about  the  membrane  before  accurate  interpretations  can  be  made.  It  should 
be  mentioned  that  these  results  on  erythrocytes  do  not  apply  to  other 
types  of  cells  or  tissues,  inasmuch  as  erythrocyte  permeability  is  in  some 
senses  unique. 

Malonate  inhibits  the  succinate  dehydrogenase  of  calf  thymus  nuclei  and 
yet  at  10  mM  has  no  effect  on  the  respiration  or  ATP  level  of  intact  nuclei 
(McEwen  et  al.,  1963  b).  This  indicated  a  failure  to  penetrate  and  it  was 
shown  with  labeled  malonate  that  this  is  indeed  the  case,  which  is  some- 
what surprising  in  view  of  the  usual  concepts  of  the  nuclear  membrane. 

Malonic  acid  does  not  enter  organic  solvents  from  water  readily,  due 
probably  to  the  dipolar  nature  of  the  unionized  carboxyl  groups.  The  par- 
tition ratios  for  malonic  acid  are  given  as  (concentration  in  solvent/concen- 
tration in  water):  oleyl  alcohol  0.049  (Collander,  1951),  ether  0.083,  iso- 
butanol  0.62,  methylisobutylketone  0.15,  and  methylisobutylcarbinol  0.37 
(Pearson  and  Levine,  1952).  The  partition  ratios  for  succinic  acid  are  always 
somewhat  higher,  as  expected.  These  data  would  indicate  that  even  the 
unionized  malonic  acid  would  not  penetrate  through  the  lipid  phase  of  the 
membrane  too  readily.  The  fact  that  the  un-ionized  forms  penetrate  better 
than  the  ionized  does  not  imply  that  passage  through  a  lipid  phase  occurs. 
The  negative  charge  on  the  ions  could  impede  movement  through  pores, 
especially  when  it  is  considered  that  in  most  cells  these  ions  must  move 
up  an  electrical  potential  gradient  to  cross  the  membrane. 

Variation  of  Malonate  Inhibition   with   pH 

One  of  the  strongest  arguments  for  the  preferential  uptake  of  the  less 
ionized  forms  of  malonic  acid  is  the  rise  in  the  inhibition  observed  with  a 
lowering  of  the  pH.  This  has  been  examined  particularly  in  plant  tissues 
and  the  results  are  quite  uniform.  Such  effects  have  been  observed  in  to- 
mato stem  slices  (Link  et  al.,  1952),  maize  roots  (Beevers,  1952),  rhubarb 
leaves  (Morrison,  1950),  barley  roots  (Laties,  1949  a),  spinach  leaves  (Bon- 
ner and  Wildman,  1946),  carrot  root  slices  (Hanly  et  al,  1952),  and  Avena 


190  1.    MALONATE 

coleoptile  (Cooil,  1952).  These  results  are  plotted  in  Fig.  1-14-19.  On  the 
other  hand,  rather  insignificant  effects  of  pH  have  been  noted  in  fungi, 
such  as  Microsporum,  Trichophyton,  Epidermophyton  (Chattaway  et  al., 
1956),  and  Pullularia  (Clark  and  Wallace,  1958).  In  Pullularia  malonate 
is  readily  oxidized;  the  oxygen  uptake  from  malonate  increases  with  a 
lowering  of  the  pH  along  with  the  inhibition  of  the  cycle  and  the  effects 
tend  to  cancel  one  another.  The  effects  of  malonate  in  tobacco  leaves  are 
not  changed  greatly  by  lowering  the  pH  from  7  to  4  (Vickery  and  Palmer, 
1957),  although  down  to  pH  5  the  uptake  of  malonate  becomes  progressively 
greater.  The  incubation  here  was  very  long  (48  hr)  and  hence  there  was 
more  opportunity  for  malonate  to  penetrate  than  in  shorter  experiments. 
It  is  unfortunate  that  no  quantitative  work  on  pH  has  been  done  with 
animal  tissues. 

It  is  also  regrettable  that  in  those  instances  in  which  malonate  inhibits 
more  strongly  at  low  pH  values  the  reversibility  of  the  inhibition  has  not 
been  adequately  tested.  Lowering  the  pH  of  the  medium  in  the  presence 
of  a  weak  acid  increases  the  amount  of  unionized  acid  entering  the  cells 
and  can  decrease  the  intracellular  pH  to  a  degree  causing  cell  damage.  It 
was  noted  in  carrot  root  tissue  that  injury  to  the  cells  occurred,  including 
loss  of  turgor  and  release  of  some  of  the  cell  contents,  with  malonate  at  a 
pH  around  4  (Hanly  et  al.,  1952).  It  is  very  difficult  in  experiments  of  this 
type  to  distinguish  between  a  specific  malonate  effect  and  a  nonspecific 
acid  damage.  Examination  of  the  reversibility  of  the  inhibitions  might 
provide  some  evidence  on  this  point. 

In  Volume  I  it  was  shown  that  the  intracellular  concentration  of  a  di- 
carboxylate  anion  is  related  to  the  total  external  concentration  in  the  follow- 
ing manner  (Eq.  I-i4-178): 

(i=),  =    I  J;     (!,)„  (1-4) 

oJi   \n.  )'i 

where  the  subscripts  o  and  i  refer  to  outside  and  inside  the  cells  and  ^f-' 
is  the  appropriate  pH  function  for  the  external  inhibitor  (see  Eq.  1-14-12). 
Two  assumptions  are  involved  in  this  formulation:  (1)  only  the  Hgl  form 
of  the  inhibitor  penetrates,  and  (2)  the  cells  are  internally  completely  buf- 
fered. The  variation  of  the  intracellular  active  1=  form  with  external  pH 
is  very  marked,  as  shown  in  the  accompanying  tabulation,  assuming  an 
intracellular  pH,  of  6.8.  Of  course,  cells  are  not  completely  buffered  and, 
as  the  internal  pH,  drops,  the  entrance  of  the  inhibitor  is  slowed,  so  that 
with  decrease  in  the  buffering  capacity  the  ratios  given  will  be  lessened. 
Here  one  is  presented  with  the  dilemma  that  at  low  pH  values  one  must 
either  assume  a  high  internal  inhibitor  concentration  or  a  significant  fall 
in  pH,.  Since  the  inhibitions  observed  are  not  as  high  as  would  be  predicted 
on  the  basis  of  the  above  equation  and  tabulation,  one  is  forced  to  conclude 


PERMEABILITY  OF  CELLS  TO  MALONATE  191 

that  the  intracellular  pH^  must  fall.  This  may  not  only  damage  the  met- 
abolic systems  but  wiU  tend  to  reduce  the  inhibition  on  succinate  dehydro- 
genase by  decreasing  the  concentration  of  dicarboxylate  anion. 


pHo 

a-)i/(it)o 

8.2 

0.0016 

7.4 

0.063 

7.0 

0.39 

6.0 

34.8 

5.0 

1610 

If  one  assumes  that  the  HI~  form  can  also  penetrate,  the  internal  con- 
centration of  the  active  1=  form  will  not  be  so  strongly  dependent  on  the 
external  pH^,  and  the  inhibition  will  increase  significantly  as  the  pH^  is 
lowered  from  6  to  5,  but  otherwise  the  same  behavior  will  be  expected.  It 
is  usually  dif&cult  to  distinguish  between  penetration  by  the  Hgl  form  only 
and  penetration  by  the  HI"  form  also.  Some  arguments  have  arisen  on  this 
point,  Bonner  and  Wildman  (1946)  believing  that  the  HI~  form  penetrates 
and  Beevers  (1952)  holding  that  only  the  Hgl  penetrates.  A  Simon-Beevers 
plot  of  the  data  from  maize  roots  (see  Chapter  1-14)  shows  that  (Hgl)  does 
not  remain  constant  for  50%  inhibition  of  respiration  over  the  pH  range 
3-6.5,  which  would  indicate  a  possible  contribution  from  the  HI~  form. 
Actually,  it  might  be  better  to  express  the  total  entry  rate  of  malonate  as: 

Entry  rate  =  Phj(H2I)o  +  Phi-(HI-)o  +  Pi=(I=)<,  (1-5) 

where  the  P's  represent  the  permeabilities  to  the  various  forms  of  the  inhi- 
bitor. Although  Pjj  I  >  Pri-  >  Pi-^  above  pH  4  (HI~)  is  much  greater 
than  (Hoi)  so  that  the  contribution  of  the  second  term  to  the  total  rate  may 
be  significant.  One  would  like  to  know  the  relative  values  of  the  P's  for  a 
particular  tissue  and  these  could  be  determined  if  the  entry  rate  of  mal- 
onate were  determined  (for  example,  with  labeled  malonate)  at  different 
pH  values.  One  must  also  bear  in  mind  that  a  change  of  pH  could  alter 
the  permeability  properties  of  the  membrane. 

The  only  data  on  the  effects  of  pH  on  malonate  action  in  animal  tissues 
were  obtained  on  rat  ventricle  strips  by  Masuoka  et  at.  (1952).  Malonate 
stimulates  the  amplitude  of  hypodynamic  strips  much  more  at  pH  6.2 
than  at  7.4.  This  positive  inotropic  action  may  be  unrelated  to  the  inhi- 
bition of  succinate  dehydrogenase,  since  succinate  at  pH  6.2  gives  essenti- 
ally the  same  response,  and  could  depend  on  the  oxidation  of  malonate 
(see  page  216).  Whatever  the  mechanism,  these  results  indicate  that  mal- 
onate penetrates  more  readily  at  the  lower  pH. 


192  1.    MALONATE 

In  comparing  the  actions  and  penetrations  of  malonate  with  succinate, 
or  with  other  dicarboxylate  anions,  it  is  necessary  to  consider  that  the 
relative  concentrations  of  the  ionic  species  can  be  quite  different.  Thus 
the  HgB  form  of  succinate  is  at  a  much  higher  concentration  than  the  same 
form  of  malonate  at  the  same  total  concentrations  (Table  1-3).  It  is  also 
possible  that  some  cells  possess  active  transport  or  carrier  systems  for  the 
substrate  dicarboxylate  anions,  allowing  succinate  to  penetrate  more  readi- 
ly than  malonate.  Permeability  in  some  cases  seems  to  be  as  specific  as 
are  enzyme  reactions,  and  can  be  quite  dependent  on  the  configurations 
and  charge  distributions  of  the  transported  substances.  This  could  be  due 
to  either  the  structure  of  the  membrane  pores  or  the  nature  of  a  carrier. 
For  these  reasons  one  must  anticipate  striking  differences  between  different 
tissues  with  regard  to  the  relative  permeabilities  to  the  various  ions,  and 
it  is  particularly  important  not  to  apply  the  results  on  plants  unreservedly 
to  animal  tissues  or  microorganisms. 


GROWTH,    DEVELOPMENT,   AND    DIFFERENTIATION 

The  responses  of  growth,  cleavage,  and  histogenesis  to  inhibitors  are 
interesting  because  they  often  demonstrate  the  nature  of  the  metabolic 
basis  for  these  important  biological  processes.  The  results  may  also  have 
bearing  on  the  possible  use  of  the  inhibitors  for  the  selective  depression 
of  the  growth  of  organisms  or  abnormal  cells  which  are  detrimental  to  the 
host.  These  processes  all  require  energy  from  the  metabolism  so  that  any 
reduction  of  either  the  exergonic  reactions  or  their  coupled  phosphoryla- 
tions would  be  expected  to  interfere  in  some  manner.  In  addition,  more 
specific  effects  may  occasionally  be  observed.  The  selective  inhibition  of 
the  growth  of  certain  cells  can  result  from  different  rates  of  growth  of  the 
cells  involved,  or  from  differences  in  the  metabolic  requirements  for  growth. 
It  is  generally  true  that  rapidly  proliferating  cells  are  more  readily  affected 
by  inhibitors  than  are  the  same  or  other  cells  growing  or  multiplying  at 
slower  rates.  It  has  also  been  demonstrated  that  various  types  of  cells  may 
utilize  different  enzymes  or  metabolic  pathways  to  support  proliferation. 
With  respect  to  malonate,  one  might  anticipate  that  cells  whose  growth 
is  in  one  way  or  another  significantly  dependent  on  the  cycle  would  be 
inhibited  more  than  cells  not  requiring  the  operation  of  the  cycle.  However, 
other  factors,  such  as  the  degree  of  penetration  of  the  malonate  or  the 
susceptibility  of  the  succinate  dehydrogenases  to  malonate,  may  be  im- 
portant. 

Virus  Multiplication 

Malonate  is  able  in  some  instances  to  suppress  the  intracellular  formation 
of  virus  without  permanently  damaging  the  host  cells.  The  results  obtained 


GROWTH,     DEVELOPMENT,    AND    DIFFERENTIATION 


193 


with  influenza  type  A  virus  isolated  from  man  and  cultured  in  chick  chorio- 
allantoic membrane  illustrate  this  well.  Ackermann  (1951)  found  that  mal- 
onate  inhibits  the  formation  of  virus  and  simultaneously  reduces  the  oxygen 
uptake  of  the  host  cells  (see  accompanying  tabulation).  Malonate  is  not 


Malonate 

%  Inhibition 

Final 

Infectivity 

(mi/) 

of  respiration 

virus  titer 

titer 

None 

213 

107.8 

20 

3 

192 

107.8 

40 

18 

69 

106.5 

60 

47 

0 

103.3 

virucidal  since  the  original  virus  can  be  recovered  from  the  infected  cul- 
tures. The  major  effect  of  malonate  is  not  to  prevent  infection  of  the  cells, 
since  essentially  the  same  results  were  obtained  when  malonate  was  added 
4  hr  after  the  inoculation  of  the  virus.  The  inhibition  is  thus  exerted  on 
the  synthesis  of  new  virus  material.  Furthermore,  the  host  cells  are  not 
damaged;  if  infected  chorioallantoic  tissue  is  exposed  to  60  mM  malonate 
for  24  hr,  virus  proliferation  is  completely  inhibited  but,  if  the  tissue  is 
washed  free  of  malonate,  becomes  susceptible  to  infection  and  supports 
virus  multiplication.  These  results  were  confirmed  and  extended  by  Eaton 
(1952),  who  reported  that  42  niM  malonate  completely  inhibits  virus  mul- 
tiplication, inhibits  respiration  around  50%,  and  does  not  alter  aerobic 
glycolysis.  The  formation  of  virus  in  minced  chorioallantoic  membrane  is 
inhibited  85%  by  27  mM  malonate  when  glucose  is  the  substrate  and  95% 
when  glutamate  is  the  substrate  (Eaton  and  Scala,  1957).  The  injection 
of  1  mg  sodium  malonate  into  chick  embryos  1  hr  after  inoculation  with 
virus  reduces  the  amount  of  virus  after  48  hr  by  55%  (Hannoun,  1952). 
The  energy  for  the  synthesis  of  new  virus  material  must  come  mainly  from 
the  host  cell  tricarboxylate  cycle.  It  is  impossible  to  reduce  the  energy 
supply  for  virus  synthesis  without  simultaneously  restricting  the  energy 
for  host  cell  activities.  However,  it  appears  that  the  virus  propagation  is 
selectively  depressed  because  it  is  the  most  endergonic  process  occurring, 
the  chorioallantoic  cells  otherwise  not  being  very  functionally  active.  As 
long  as  the  low  maintenance  energy  requirement  is  satisfied,  the  cells  are 
not  damaged  (see  Fig.  1-9-6).  If  one  were  dealing  with  virus  in  an  active 
tissue,  such  as  nerve  or  muscle,  it  would  be  much  more  difficult,  or  im- 
possible, to  selectively  block  virus  multiplication. 

Other  viruses  growing  in  animal  cells  have  been  studied  with  various 
results.  Malonate  at  6.7  mM  has  a  definite  depressant  effect  on  the  proli- 
feration of  vaccinia  virus  in  chick  embryo  cultures  (R.  L.  Thompson,  1947) 


194  1.    MALONATE 

and  similar  results  were  reported  for  psittacosis  virus  (Morgan,  1954).  In 
neither  case  does  malonate  depress  the  growth  of  the  tissue  cultures,  noi 
does  it  inhibit  the  cellular  contractions  of  chick  heart  cultures.  Very  slight 
effects  of  malonate  at  concentrations  from  10  to  100  mM  were  observed 
with  foot-and-mouth  disease  virus  in  bovine  kidney  culture  cells  (Polatnick 
and  Bachrach,  1960).  Since  there  is  little  inhibition  of  the  respiration,  it  is 
possible  that  malonate  does  not  penetrate  adequately.  At  100  mM,  respira- 
tion is  inhibited  23%,  virus  yield  perhaps  20%,  and  there  is  a  10  min  delay 
in  the  appearance  of  virus.  Finally,  malonate  was  found  to  actually  stimu- 
late feline  pneumonitis  virus  proliferation  in  the  isolated  chick  embryo  yolk 
sac,  10  mM  increasing  the  virus  titer  about  33%  (Moulder  et  al.,  1953). 
Thus  a  wide  variety  of  effects  have  been  observed  with  different  viruses 
and  host  cells,  no  general  conclusions  being  possible  at  this  time.  I  am  not 
aware  of  any  studies  on  the  effects  of  malonate  on  the  course  of  virus  in- 
fections in  whole  animals. 

Plant  viruses  have  been  inadequately  investigated  and  results  on  the 
tobacco  mosaic  virus  only  are  available.  Although  malonate  at  0.5  mM  de- 
creases the  number  of  lesions/cm^  in  detached  tobacco  leaves  from  28.9 
to  21.9,  Chiba  et  al.  (1953)  felt  that  this  result  is  statistically  insignificant. 
Ryzhkov  and  Marchenko  (1954,  1955)  reported  that  malonate  inhibits 
multiplication  of  this  virus  and  that  this  is  reversed  by  fumarate,  but 
Schlegel  (1957)  found  only  variable  effects  of  3  mM  malonate  on  the  yield 
of  virus  in  leaf  discs.  The  spraying  of  10  mM  malonic  acid  solutions  (pH 
2.7-3.6)  onto  the  leaves  of  bean  plants  decreases  the  number  of  virus  lesions 
68%  without  leaf  damage  (Matthews  and  Proctor,  1956),  but  this  may  be 
unrelated  to  the  action  of  malonate  on  the  cycle  inasmuch  as  succinic  acid 
is  even  more  inhibitory.  This  is  probably  a  nonspecific  acid  effect  because 
the  cycle  intermediates  usually  increase  the  virus  yield  when  they  are  added 
to  leaf  cultures. 

It  is  somewhat  surprising  that  malonate  has  no  demonstrable  effect  on 
the  multiplication  of  E.  coli  phage.  Spizizen  (1943)  found  no  effect  at  10  mM 
under  any  conditions  of  virus  growth,  and  Czekalowski  (1952)  reported  no 
actions  on  either  T2  phage  or  host  cells  at  concentrations  from  0.1  to  100  mM. 
It  may  well  be  that  phage  proliferation  is  not  directly  dependent  on  the 
energy  derived  from  the  cycle,  but  inhibition  by  2,4-dinitrophenol,  cyanide, 
and  fluoride  is  observed.  In  fact,  Czekalowski  stated  that  all  the  inhibitors 
that  depress  phage  selectively  seem  to  act  in  some  manner  on  the  cyto- 
chrome system  and  are  able  to  inhibit  succinate  oxidase;  yet  the  most 
specific  inhibitor  for  this  enzyme  is  inactive.  Lack  of  penetration  is  an 
unlikely  explanation  and  this  failure  of  malonate  to  inhibit  deserves  further 
study. 


GROWTH,    DEVELOPMENT,    AND    DIFFERENTIATION  195 

Multiplication  of  Bacteria,  Fungi,  and  Other  Microorganisms 

Bacterial  growth  is  apparently  fairly  resistant  to  malonate,  despite  the 
many  observations  of  enzyme  and  metabolic  inhibitions  in  these  organisms. 
The  absence  of  any  effect  on  E.  coli  at  malonate  concentrations  up  to  100  mM 
reported  by  Czekalowski  was  mentioned  above,  but  Loveless  et  al.  (1954) 
found  50%  inhibition  of  growth  at  19.2  ml/  malonate,  with  no  effect  on  cell 
size.  Malonate  at  300  roM  produces  somewhat  elongated  E.  coli  cells  and 
at  800  mM  they  are  markedly  lengthened  and  often  U-shaped:  division  is 
abolished  but  the  cells  continue  to  elongate  slowly  (Schweisfurth  and 
Schwartz,  1959).  The  effects  of  malonate  must  depend  on  many  factors, 
including  the  nutrient  medium,  the  duration  of  the  growth  phase  studied, 
and  the  pH.  Although  Rosenberg  (1948)  found  the  growth  of  Clostridium 
saccharobutyricum  to  be  inhibited  by  malonate,  a  concentration  of  100  mM 
was  used,  so  that  the  mechanism  of  the  effect  is  not  clear.  His  observation 
that  the  inhibition  is  overcome  by  meso-inositol  and  borate  must  be  inter- 
preted as  indicating  a  unique  approach  to  the  study  of  malonate  inhibition. 
Malonate  at  10  mM  has  a  slightly  depressant  action  on  rate  of  germination 
of  Bacillus  subtilis  spores  but  does  not  affect  growth  in  culture  (Hachisuka 
et  al.,  1955).  The  bacteriostatic  concentrations  of  malonate  were  given  as 
3.2  mM  for  Pseudomonas  fluorescens  and  7.7  mM  for  B.  aerogenes  (Cooper 
and  Goddard,  1957)  but  the  acid  was  used,  the  pH  being  2.5  and  1.5,  re- 
spectively, so  that  a  nonspecific  acid  effect  is  the  most  likely  explanation, 
especially  as  succinic  acid  is  as  inhibitory.  It  is  clear  that  the  investigations 
of  the  effects  of  malonate  on  bacterial  growth  leave  everything  to  be  desired. 

The  sporulation  and  growth  of  yeast  are  inhibited  by  malonate  quite 
potently  and  in  parallel  fashion,  50%  depression  of  each  occurring  at  5  mM 
(Miller  and  Halpern,  1956).  On  the  other  hand,  Hensenula  ellipsoidospora, 
a  vellum-forming  yeast,  grows  more  rapidly  in  the  presence  of  malonate 
(Luteraan,  1953).  It  would  certainly  not  be  too  surprising  to  find  certain 
microorganisms  stimulated  by  malonate  inasmuch  as  many  can  oxidize 
malonate  readily  (page  228).  Malonate  arrests  sporulation  of  Aspergillus 
niger  without  suppressing  growth  (Behal,  1959)  but  germination  of  Neuro- 
spora  ascospores  is  not  blocked  by  10  mM  malonate  even  at  pH  2.3  and 
after  24  hr  (Sussman  et  al.,  1958),  nor  is  the  germination  of  Puccinia  uredo- 
spores  affected  significantly  by  20  mM  malonate  at  pH  4.8,  although  the 
respiration  is  inhibited  around  37%  (Farkas  and  Ledingham,  1959).  The 
formation  of  conidia  is  often  essential  for  the  spread  of  the  scab  disease 
of  apple  caused  by  Venturia  inaequalis  and  so  Kirkham  and  Flood  (1963) 
investigated  the  effects  of  various  respiratory  inhibitors  on  ascosporulation. 
Malonate  was  found  to  inhibit  at  high  concentrations,  the  inhibition  sur- 
prisingly increasing  with  increase  in  the  pH  (see  accompanying  tabulation). 
This  might  imply  an  action  on  or  within  the  membrane;  this  is  supported 
by  the  relative  lack  of  effect  on  the  respiration  and  the  rather  potent  inhi- 


196  1.    MALONATE 

bition  produced  by  ^ra ws-aconitate.  The  injection  of  50  milf  malonate  into 
the  leaf  petioles,  however,  increases  infectivity  so  that  a  more  significant 


Malonate 
(mil/) 

Initial  pH 

0/ 

/o 

Inhibition   of 
sporulation 

50 

4.2 

4 

100 

39 

40 

5.0 

Stim  7 

100 

71 

50 

6.2 

68 

100 

96 

effect  on  the  host  tissue  is  evident.  The  development  of  Puccinia  rust  on 
wheat  seedling  leaves  is  inhibited  by  10  mM  malonate,  but  the  leaf  tissue 
is  damaged  (Samborski  and  Forsyth,  1960).  In  this  particular  case  the 
phytotoxicity  is  greater  than  the  rust  suppression  so  that  malonate  could 
not  be  used  commercially.  Malonate  esters  have  been  tested  for  inhibition 
of  mold  growth  in  syrups,  but  13  mM  does  not  have  much  effect  over  144  hr 
(Lord  and  Husa,  1954);  however,  these  esters  are  used  as  fungistatic  agents 
in  soy  sauce  in  Japan  (Tsukamoto,  1951).  Another  instance  of  growth  stim- 
ulation by  malonate  was  reported  for  Endamoeba  histolytica  (Nakamura 
and  Baker,  1956);  the  average  cell  count  per  field  at  the  end  of  3-4  days 
was  1  in  the  control  and  16  in  the  presence  of  12.8  mM  malonate,  indicating 
possible  metabolism  of  the  malonate  by  these  organisms. 

Plant  Growth 

Avena  coleoptile  growth  is  sometimes  stimulated  and  sometimes  depressed 
by  malonate,  the  response  depending  on  the  strain  of  oats  used,  the  pH, 
and  whether  the  sodium  or  potassium  salt  of  malonate  is  used.  The  marked 
inhibition  (61%)  reported  by  Commoner  and  Thimann  (1941)  for  10  mM  po- 
tassium malonate  over  24  hr  has  not  been  observed  by  others.  Albaum  and 
Eichel  (1943)  found  only  stimulation  (around  30%)  from  1-5  niM  potassium 
malonate  over  a  period  of  160  hr,  and  it  was  felt  that  malonate  was  serving 
as  a  substrate,  which  was  substantiated  by  the  higher  respiration  in  the 
presence  of  malonate.  Thimann  and  Bonner  (1948)  provided  further  evi- 
dence for  this  by  the  finding  that  malonate  at  1  mM,  having  little  effect  by 
itself,  antagonizes  the  marked  inhibition  produced  by  iodoacetate.  How 
much  of  this  is  due  to  malonate  and  how  much  to  potassium  is  difficult 
to  say.  Cooil  (1952)  confirmed  the  counteraction  of  iodoacetate  inhibition 
by  potassium  malonate,  but  found  that  the  sodium  salt  is  not  nearly  as 
potent,  implicating  the  potassium  ion.  The  failure  of  malonate  to  inhibit 
the  growth  is  probably  the  result  of  poor  penetration,  as  shown  by  the 


GROWTH,    DEVELOPMENT,    AND    DIFFERENTIATION  197 

effects  of  pH  (see  accompanying  tabulation);  the  pH  alone  has  little  effect 
on  growth.  The  malonate  inhibition  is  satisfactorily  reversed  by  fumarate. 


pH  Mean  growth  (mm)  in  malonate  3  mM 

6.5  1.56 

6.0  1.70 

5.5  1.55 

5.0  0.22 

4.5  0.16 

The  mitotic  activity  of  the  excised  roots  of  the  garden  pea  {Pisum  sativum) 
is  stimulated  by  glucose.  This  is  very  strongly  blocked  by  malonate  at  pH 
5.5.  A  concentration  of  0.01  mM  delays  the  action  of  glucose  2  hr  but  does 
not  inhibit  mitosis;  0.1  mM  inhibits  mitotic  activity  around  50%;  0.5  milf 
almost  completely  inhibits  mi^-^ses;  and  1  mi!f  not  only  inhibits  completely 
but  produces  some  toxic  effects  (Wilson  et  al.,  1959).  It  was  postulated 
that  the  initiation  of  mitosis  is  dependent  on  the  cycle,  since  once  mitosis 
began  it  proceeded  to  telophase  normally.  The  progression  through  mitosis 
may  be  coupled  with  a  more  anaerobic  type  of  metabolism.  The  growth 
and  cell  proliferation  in  tissue  cultures  of  the  crown  galls  of  various  plants 
(marigold,  Paris  daisy,  periwinkle,  and  sunflower)  are  inhibited  to  different 
degrees  by  malonate.  Cultures  from  normal  tobacco  stem  are  inhibited 
similarly.  At  10  mM,  the  following  inhibitions  may  be  estimated  from  the 
curves  given  by  Hildebrandt  et  al.  (1954):  sunflower  0%,  marigold  29%, 
tobacco  40%,  Paris  daisy  60%,  and  periwinkle  67%.  At  80  mM  malonate, 
all  are  inhibited  completely.  A  question  arises  again  as  to  whether  these 
effects  are  related  to  cycle  inhibition,  since  succinate,  pyruvate,  acetate, 
and  other  organic  anions  inhibit  also.  The  pH  was  6.0  so  that  acid  effects 
should  not  be  important. 

Egg  Cleavage  and   Embryogenesis 

The  best  and  most  interesting  work  on  the  growth  responses  to  malonate 
has  been  done  with  marine  invertebrate  eggs  and  embryos.  Since  this  work 
was  done  in  sea  water,  we  must  bear  in  mind  that  the  concentration  of 
free  malonate  is  much  less  than  the  total  concentration  due  to  the  high 
amounts  of  Ca++  and  Mg+"'".  When  malonate  is  added  at  a  total  concen- 
tration of  25  mM,  it  is  likely  that  the  free  malonate  is  around  4  mM  (see 
Table  1-5).  In  addition,  the  pH  of  sea  water  is  near  8.2  and  this  is  unfavor- 
able to  malonate  penetration  into  the  cells.  Considering  these  factors,  it  is 
surprising  that  such  definite  and  characteristic  effects  of  malonate  have 
been  observed. 


198  1.    MALONATE 

Egg  cleavage  is  generally  rather  sensitive  to  malonate.  Although  Arbacia 
eggs  divide  normally  in  1  mM  malonate  (Krahl  and  Clowes,  1940),  Dend- 
raster  eggs  are  inhibited  quite  well  at  this  concentration  (Pease,  1941).  The 
development  of  bilaterality  in  Dendraster,  seen  with  many  inhibitors,  does 
not  occur  with  malonate  even  at  10  mM,  demonstrating  a  true  differential 
effect  on  cleavage.  Division  of  Arbacia  and  Chaetopterus  eggs  is  inhibited 
completely  by  malonate  at  70  mM,  although  40  mM  is  essentially  without 
action,  and  this  is  completely  reversed  by  fumarate,  indicating  a  block  of 
the  cycle  (Brust  and  Barnett,  1952;  Barnett,  1953).  This  is  not  a  sodium 
effect  since  NaCl  does  not  inhibit.  Such  high  concentrations  of  malonate 
are  not  unreasonable  in  work  in  sea  water  and  the  inhibitions  here  may  be 
quite  specific.  Egg  cleavage  seems  to  depend  on  the  ATP  generated  in  the 
cycle,  because  concentrations  of  malonate  that  completely  inhibit  the  cleav- 
age of  Chaetopterus  and  Lytechinus  eggs  cause  an  immediate  drop  in  the 
high-energy  phosphate  to  the  unfertilized  levels,  and  inorganic  phosphate 
is  actually  lost  from  the  cells  (Barnett  and  Downey,  1955). 

The  effects  of  malonate  on  the  development  of  Arbacia  eggs  were 
thoroughly  investigated  by  Rulon  (1948),  who  demonstrated  abnormal  dif- 
ferentiation with  low  concentrations  of  malonate.  If  fertilized  eggs  in  the 
1-2-cell  stage  are  placed  in  1.2  mM  malonate,  a  slight  retardation  of  cleav- 
age is  observed,  and  at  13  hr  (when  the  controls  are  swimming  bilateral 
gastrulae)  there  is  no  evidence  of  gastrulation,  development  having  pro- 
gressed only  to  spherical  blastulae  with  no  ventral  flattening.  After  24  hr 
(when  the  controls  are  plutei),  abnormal  gastrulae  with  thickened  apical 
ends  and  long  active  cilia  are  seen.  At  48  hr  some  had  developed  into  ab- 
normal plutei  with  ciliated  apical  knobs  rather  than  normal  arms.  When 
malonate  is  removed  after  13  hr,  quite  normal  plutei  are  formed,  showing 
that  the  effect  is  readily  reversible.  It  is  interesting  that  eggs  exposed  to 
malonate  for  varying  times,  then  washed  free  of  malonate  and  fertilized, 
show  abnormal  development,  demonstrating  that  malonate  can  so  disturb 
egg  metabolism  that  the  effects  are  made  evident  later  after  the  malonate 
is  no  longer  present.  Exposure  of  unfertilized  eggs  to  1.44  mM  malonate 
for  12  hr,  for  example,  leads  to  only  30  40%  cleavage  with  irregular  cleav- 
age furrows  and  cells  of  unequal  sizes,  development  not  progressing 
beyond  irregular  blastulae.  Rulon  postulated  a  gradient  of  succinate  de- 
hydrogenase throughout  the  cells  and  embryos,  paralleling  the  physiological 
activity  gradient,  but  it  is  not  necessary  to  assume  this  to  explain  effects  on 
differentiation.  In  connection  with  our  work  on  parapyruvate  (Montgomery 
and  Bamberger,  1955),  we  examined  the  development  of  Strongylocentrotus 
purpuratus  in  25  mM  malonate.  Up  to  24  hr  no  discernible  differences  from 
the  controls  are  seen,  but  at  44  hr  a  slight  inhibition  of  development  can  be 
detected,  with  less  formation  of  the  primary  mesoderm.  Incoordination  of 
ciliary  activity  is  also  evident,  most  of  the  blastulae  simply  rotating  rather 


GROWTH,    DEVELOPMENT,    AND    DIFFERENTIATION  199 

than  swimming.  At  64  hr,  when  the  controls  are  beginning  to  gastrulate, 
the  treated  embryos  are  still  spherical  blastulae,  and  at  72  hr  have  not 
progressed  beyond  this  point.  In  this  species,  malonate  would  appear  to 
be  a  rather  specific  inhibitor  of  gastrulation  without  altering  cleavage  pri- 
marily. It  may  be  noted  that  other  cycle  inhibitors,  such  as  parapyruvate 
and  fluoroacetate,  also  specifically  block  gastrulation  in  Strongylocentrotus. 
It  is  difficult  to  understand  these  marked  differences  between  the  behaviors 
of  various  sea  urchin  eggs,  and  especially  the  striking  effects  obtained  by 
Rulon  with  such  low  malonate  concentrations.  The  free  malonate  concen- 
trations in  his  work  would  have  been  around  0.17  milf  (he  used  Ca++-free 
sea  water)  and  the  intracellular  concentration  presumably  much  less. 

There  is  evidence  that  insect  spermatogenesis  is  an  aerobic  process  with 
the  terminal  electron  transport  through  the  cytochrome  system,  and  that 
the  cycle  is  the  primary  pathway  for  substrate  oxidations.  Yet  no  inhi- 
bition of  meiosis  or  differentiation  into  spermatids  and  spermatozoa  by 
50  ToM  malonate  is  observed  in  hanging-drop  cultures  from  the  Cecropia 
silkworm  (Schneiderman  et  at.,  1953).  These  experiments  were  carried  out 
at  pH  6.8-7.2  and  it  is  possible  that  malonate  failed  to  penetrate. 

Amphibian  gastrulation  is  inhibited  by  high  concentrations  of  malonate. 
Frog  embryos  at  the  early  dorsal  lip  stage  were  dissected  to  give  explants 
which  were  exposed  to  40  raM  malonate  at  pH  8.0  for  18  hr.  Development 
does  not  proceed  beyond  the  next  stage  (Ornstein  and  Gregg,  1952;  Gregg 
and  Ornstein,  1953).  There  is  no  differential  effect  on  the  respiration  of 
dorsal  and  ventral  explants,  both  being  inhibited  about  60%.  Unfortunately, 
there  is  some  doubt  that  the  block  of  development  is  due  to  any  specific 
effect  of  the  malonate  since  45  vciM  NaCl  apparently  inhibits  to  the  same 
degree.  Thus  the  mechanism  of  the  block  could  have  been  osmotic  or  a  Na"*" 
effect.  The  chick  embryo  seems  to  be  much  more  sensitive  to  malonate. 
Explants  of  chick  embryo  in  the  presence  of  glucose  undergo  morphogenesis 
and  differentiation  to  the  formation  of  the  central  nervous  system  and  an 
actively  beating  heart.  Malonate  at  1-2  mJf  exerts  striking  differential  ef- 
fects (Spratt,  1950).  Although  no  differences  were  noted  during  the  first 
20  hr,  afterwards  the  central  nervous  system  degenerates  completely  while 
the  heart  forms  normally  and  beats.  Malonate  is  the  most  specific  inhibitor 
for  the  development  of  the  nervous  system.  Some  antagonism  of  this  effect 
was  seen  with  succinate  but  none  with  fumarate.  No  inhibition  of  mitoses 
in  cultures  of  chick  bone  is  observed  with  malonate  from  65  to  138  nxM 
(A.  F.  W.  Hughes,  1950). 

Mitoses   in    Mammalian   Tissues 

Epidermal  mitotic  activity  in  mouse  ear  fragments  was  determined  over 
4  hr  periods  at  pH  7.4  and  38^  by  BuUough  and  Johnson  (1951).  From  the 
effects  of  anaerobiosis  and  various  substrates  it  was  concluded  that  the 


200  1.    MALONATE 

energy  for  mitosis  is  derived  from  cycle  oxidations.  Malonate  at  20  mM 
prevents  the  cell  from  entering  mitosis  for  3  hr  but  evidence  of  recovery  is 
seen  after  this  time.  However,  at  4  hr,  although  some  cells  are  progressing 
through  mitosis,  the  number  of  mitoses  is  definitely  less  than  in  the  controls. 
There  is  no  evidence  that  malonate  can  stop  mitosis  once  it  has  begun. 
Epidermal  cells  adjacent  to  the  wound  show  a  higher  number  of  mitoses 
than  normal  epidermis  and  malonate  depresses  both  strongly  (see  accom- 
panying tabulation).  (Bullough  and  Laurence,  1957).  Malonate  can  thus 


Malonate 


Average  number  of  mitoses 


(milf)  Epidermis  adjacent  to  wound  Normal  epidermis 

0  8.20  0.96 

10  1.23  0.22 

20  0.05  0.18 

30  0.05  0.01 


inhibit  healing  without  significant  damage  to  the  tissue,  since  no  necrosis 
is  seen  following  incubation  with  malonate.  Similar  results  were  obtained 
by  Gelfant  (1960)  and  concentrations  as  low  as  0.5  mM  are  definitely  inhi- 
bitory. After  4  hr  50  m  M  malonate  produces  some  necrosis.  The  mitotic 
rate  in  the  germinal  epithelium  of  rat  ovaries  is  also  suppressed  by  malo- 
nate, 2  mM  having  variable  effects  but  inhibiting  21%  on  the  average  and 
10  mi/  inhibiting  59%  (Weaver,  1959).  The  relatively  high  sensitivity  of 
mammalian  tissue  mitosis  to  malonate  would  implicate  the  cycle  as  an 
important  source  of  energy  for  this  process. 

Growth  of  Neoplastic  Tissues 

The  respiration  of  various  types  of  tumor  cells  is  inhibited  by  malonate 
(Table  1-26)  but  comparisons  with  the  appropriate  normal  tissues  have 
seldom  been  made.  Amino  acid  uptake  is  also  inhibited  (Kit  and  Greenberg, 
1951)  and  high-energy  phosphate  compounds  reduced  (Greaser  and  Schole- 
field,  1960),  but  whether  tumor  tissue  is  more  or  less  sensitive  than  normal 
tissue  to  malonate  is  not  known.  Fishgold  (1957)  obtained  evidence  that 
hepatoma  succinate  oxidase  is  inhibited  more  readily  than  the  enzyme  from 
normal  mouse  liver,  but  it  is  not  certain  if  this  is  a  true  difference  in  the 
affinities  of  the  dehydrogenase  active  center  for  malonate  or  the  result  of 
other  factors.  Other  than  this,  there  is  no  demonstrated  metabolic  difference 
between  tumor  and  normal  tissues  with  respect  to  inhibition  by  malonate. 
The  time  course  of  the  accumulation  of  succinate  in  the  Flexner-Jobling 
tumor  is  essentially  the  same  as  in  other  tissues  (Fig.  1-11),  but  the  relation- 


GROWTH,    DEVELOPMENT,    AND    DIFFERENTIATION 


201 


ship  of  the  accumulation  to  malonate  concentration  is  not  linear  for  the 
tumor  (Fig.  1-12).  As  pointed  out  by  Busch  and  Potter  (1952  b),  the  ac- 
cumulation of  succinate  in  the  tumor  may  be  superficially  indistinguishable 
from  that  in  normal  tissues,  but  there  is  reason  to  believe  that  the  succinate 
in  the  tumor  arises  by  somewhat  different  pathways,  mainly  from  gluta- 
mate  and  related  compounds.  There  is  little  reason  to  believe  from  the 
known  metabolic  characteristics  of  tumor  tissues  that  malonate  would  se- 
lectively depress  their  growth;  indeed,  one  might  expect  them  to  be  less 
sensitive  to  malonate,  except  for  the  fact  that  tumor  cells  are  often  more 
rapidly  proliferating  and  more  active  metabolically  than  normal  tissues. 

Neoplastic  growth  in  general  seems  to  be  relatively  resistant  to  malonate. 
Malonate  at  30  mM  is  not  toxic  to  cultures  of  various  tumors  but  50  mM 
is  toxic  to  all  types  of  cells  in  a  few  hours  (Chambers  et  al.,  1943).  It  is 
interesting  that  no  differences  in  susceptibility  of  lymphocytes  and  other 
wandering  cells,  whether  from  normal  or  neoplastic  tissues,  are  seen.  Eagle's 
KB  strain  of  human  carcinoma  cells  is  more  sensitive,  the  50%  inhibitory 
concentration  of  malonate  being  around  4  ml/  (Smith  et  al.,  1959).  Malo- 
nate, like  many  metabolic  inhibitors,  is  capable  of  producing  acentric  blebs 
on  Sarcoma  37  ascites  cells  (Belkin  and  Hardy,  1961).  Although  this  indi- 
cates some  alteration  of  the  membrane  properties  and  the  permeability  to 
water,  the  relationship  to  growth  inhibition  is  unknown. 

Studies  of  the  action  of  malonate  on  tumors  growing  in  whole  animals 
are  more  pertinent  to  the  problem  of  the  possible  value  of  this  inhibitor 
in  therapy.  The  earliest  work  was  done  by  Boyland  (1940)  following  his 
investigations  of  the  effects  of  malonate  on  tumor  respiration.  Definite  sup- 
pression of  growth  was  observed  with  malonate  at  a  dose  well  below  the 
lethal,  the  carcinoma  being  more  sensitive  than  the  sarcoma  (see  accom- 
panying tabulation).  Actual  regression  of  the  carcinomata  was  observed. 


Inhibitor 


Dose 


%  Inhibition  of  growth  of 


LD,„ 


(mg/day)       Qrafted  sarcomata    Spontaneous  carcinomata 


Malonate 

20 

32 

Malonamide 

25 

36 

Ethylmalonate 

40 

27 

Glutarate 

25 

8 

Adipate 

25 

0 

31 

100 

44 

150 

25 

160 

48 

150 

35 

200 

as  indicated  by  the  131%  inhibition.  Whether  ethylmalonate  and  malon- 
amide are  metabolized  to  malonate  and  are  active  for  this  reason  is  not 
known,  but  the  lesser  potencies  compared  to  malonate  do  not  suggest  that 
these  uncharged  substances  penetrate  better.  Malonate,  fluoride,  iodoacetate 


202  1.    MALONATE 

and  azide  were  administered  to  patients  with  advanced  neoplasia  and  tem- 
porary suppressive  effects  were  noted  (Black  and  Kleiner,  1947;  Black  et  at., 
1947).  Sodium  malonate  was  given  orally  at  doses  of  1-1.5  g/day.  It  is 
difficult  to  state  clearly  the  effects  of  malonate,  since  the  inhibitors  were 
usually  given  sequentially  or  together,  but  it  was  stated  that  hematological 
remissions  occurred  in  acute  myeloblastic  leukemia  and  that  shrinkage  of 
solid  tumors,  with  relief  of  pain,  was  evident.  The  tumor  cells  usually  be- 
come refractory  to  these  inhibitors.  After  resistance  to  fluoride  and  iodo- 
acetate  has  developed,  a  beneficial  effect  is  seen  with  malonate.  It  would 
seem  that  these  results  are  encouraging  enough  to  warrant  further  study, 
particularly  with  combinations  of  the  inhibitors  to  prevent  or  reduce  the 
development  of  resistance.  Several  derivatives  of  malonate  were  tested 
against  mammary  Carcinoma  755  in  mice  and  suppressive  action  was  dem- 
onstrated (Freedlander  et  al.,  1956).  Malonic  acid  at  1.2%  in  the  diet  did 
not  affect  either  the  tumor  size  or  the  growth  of  the  mice.  The  most  active 
ethyl  ester  was  diethylethoxymethylenemalonate  (the  group  =CH — 0 — 
— CH2CH3  on  C-2),  which  at  1.2%  in  the  diet  reduces  the  surface  area  of 
the  tumors  83%  while  causing  minimal  loss  of  body  weight.  This  substance 
is  as  effective  as  8-azaguanine  and  is  less  depressant  on  the  total  body 
growth.  Some  diamides  are  also  active,  iV-dimethylmalondiamide  being  the 
most  active,  reducing  tumor  area  68%  with  no  effects  on  body  growth. 
It  was  thought  that  these  substances  may  be  inhibitory  to  succinate  de- 
hydrogenase, probably  after  hydrolysis,  but  there  is  no  evidence  at  present 
for  this  and  it  is  quite  possible  that  the  mechanism  is  entirely  different. 
Despite  the  lack  of  evidence  for  a  high  susceptibility  of  the  metabolism  or 
growth  of  isolated  neoplastic  cells  to  malonate,  the  in  vivo  work  has  brought 
out  interesting  effects  that  deserve  more  thorough  investigation. 


CELLULAR   AND  TISSUE    FUNCTION 

Many  studies  of  the  effects  of  malonate  on  physiological  function  with 
the  object  of  relating  the  cellular  activity  to  succinate  oxidase  or  the  cycle 
have  been  reported,  but  in  only  a  few  instances  have  the  necessary  data 
been  obtained  and  a  relationship  adequately  established.  The  general  rela- 
tions between  enzyme  inhibition  and  changes  in  cellular  function  were 
discussed  in  Chapter  1-9,  and  several  methods  for  demonstrating  correla- 
tions were  presented.  The  complexities  of  such  studies  were  emphasized 
and  the  difficulties  commonly  encountered  are  well  illustrated  in  the  results 
of  malonate  inhibition.  In  addition  to  the  various  possible  metabolic  effects 
of  malonate,  one  must  bear  in  mind  that  malonate,  or  the  cations  added 
with  it,  can  directly  alter  functional  processes,  actions  which  can  be  dis- 
tinguished by  the  proper  controls. 


CELLULAR    AND    TISSUE    FUNCTION  203 

Single   Cell    Motility 

A  thorough  study  of  the  respiration  and  motility  of  the  ciliate  Paramecium 
caudatum  was  made  by  Holland  and  Humphrey  (1953).  Malonate  at  20  mM 
inhibits  the  endogenous  respiration  31%  but  has  no  effect  on  the  ciHary 
activity  over  1  hr.  The  oxidation  of  cycle  substrates,  such  as  citrate,  a- 
ketoglutarate,  fumarate,  and  malate,  is  well  inhibited,  and  it  is  likely  that 
the  cycle  is  present.  Furthermore,  the  succinate  oxidase  in  homogenates 
is  quite  sensitive  to  malonate.  One  must  conclude  either  that  malonate 
does  not  penetrate  sufficiently  or  that  the  motility  is  not  entirely  dependent 
on  the  cycle  for  an  energy  supply.  The  answer  to  this  problem  may  lie  in 
the  observation  that  malonate  does  not  inhibit  the  oxidation  of  pyruvate 
or  acetate.  Holland  and  Humphrey  point  out  that  there  are  many  alternate 
pathways  for  metabolism  in  paramecia.  Thus  it  is  possible  that  normally 
energy  is  derived  from  the  cycle  but  during  a  cycle  block  energy  is  provid- 
ed by  other  alternate  reactions.  In  the  human  parasitic  ciliate  Balantidium 
coli,  malonate  depresses  both  respiration,  which  is  probably  endogenous, 
and  motility  (Agosin  and  von  Brand,  1953).  The  flagellar  motility  of  Ba- 
cillus brevis  is  not  affected  by  10  mM  malonate  (De  Robertis  and  Peluffo, 
1951).  The  motility  of  bull  sperm  is  reduced  appreciably  by  10  mM  mal- 
onate and  the  endogenous  respiration  is  simultaneously  inhibited  55% 
(Lardy  and  Phillips,  1945).  However,  the  results  are  quite  different  when 
various  substrates  are  present  and  motility  is  depressed  very  little  or  not 
at  all.  Since  motility  is  not  depressed  with  glucose  or  pyruvate  as  sub- 
strate, whereas  it  is  with  acetate,  it  is  likely  that  energy  sources  other 
than  the  cycle  can  be  used.  Ciliary  and  flagellar  activities  in  different  or- 
ganisms are  thus  affected  in  various  ways  by  malonate  and  no  uniform 
picture  emerges  from  the  data  at  present  available. 

Chemotaxis  and  phagocytosis  in  guinea  pig  leucocytes  are  partially  inhi- 
bited by  malonate  but  the  concentration  used  (140  mM)  was  so  high  that 
little  significance  can  be  attached  to  these  results  (Lebrun  and  Delaunay, 
1951).  Bacterial  phagocytosis  by  human  neutrophiles  is  depressed  moder- 
ately by  malonate  from  33.5  to  100  mM  but  no  effect  is  seen  with  6.7  mM 
(Berry  and  Derbyshire,  1956).  Malonate  at  1  mM  has  no  effect  on  the  mi- 
gration of  amphibian  chromatophores  in  cultures  of  the  neural  crest  (Flick- 
inger,  1949).  At  10  mM,  malonate  is  toxic  to  these  preparations  and  not 
specifically  depressant  to  the  motility.  These  limited  results  point  to  a 
relative  insusceptibility  of  ameboid-type  movements  to  malonate,  which  is 
not  surprising  considering  the  known  metabolic  characteristics  of  such  cells. 

Renal  Tubular  Transport 

Most  of  the  studies  of  the  effects  of  malonate  on  the  kidney  have  involved 
the  active  transport  of  p-aminohippurate.  The  accumulation  of  this  sub- 


204  1.    MALONATE 

stance  by  kidney  cortex  slices,  which  is  quite  marked  (slice/medium  ratios 
around  10),  is  reduced  by  malonate:  in  slices  from  dogs,  5  raM  inhibits  25%, 
10  mM  inhibits  35%,  and  20  mM  inhibits  55%  (Shideman  and  Rene,  1951  b) 
and  in  slices  from  rabbits  20  mM  inhibits  72%  while  reducing  the  respiration 
with  acetate  by  53%  (Cross  and  Taggart,  1950).  This  action  can  be  shown 
to  occur  in  the  whole  animal  also.  Dominguez  and  Shideman  (1953,  1955) 
removed  one  kidney  from  rats,  administered  solutions  of  sodium  malonate, 
removed  the  other  kidney,  and  determined  the  accumulation  of  j^-amino- 
hippurate.  When  approximately  10  millimoles/kg  of  malonate  are  injected 
subcutaneously,  the  uptake  of  p-aminohippurate  is  depressed  52%,  the 
average  slice/medium  ratio  falling  from  9.28  to  4.47.  The  decrease  in  the 
slice/medium  ratio  is  linear  with  intravenous  doses  of  4-7  millimoles/kg. 
Some  effect  occurs  at  15  min  after  the  injections,  the  maximal  inhibition 
is  around  60  min,  and  the  transport  mechanisms  have  returned  to  normal 
by  150  min,  indicating  the  ready  reversibility  of  the  action.  The  transport 
inhibition  can  also  be  demonstrated  by  the  renal  clearance  technique  in 
dogs  (Shideman  and  Rene,  1951  b).  Malonate  at  a  dose  of  0.96  millimole/kg 
depresses  the  p-aminohippurate  Tm*  73%.  It  was  stated  that  50%  inhi- 
bition of  renal  succinate  dehydrogenase  is  produced  by  1.32  mM  malonate 
and  thus  the  dose  given  would  be  expected  to  inhibit  in  vivo,  but  the  con- 
centration of  malonate  in  the  tubular  fluid  is  probably  much  higher  than 
in  the  blood  and  permeability  factors  must  also  be  important.  Farah  and 
Rennick  (1954,  1956)  studied  the  effects  of  many  inhibitors  on  the  p-amino- 
hippurate  uptake  in  guinea  pig  kidney  slices  and  the  results  are  summarized 
in  Fig.  1-18.  Malonate  is  one  of  the  weakest  inhibitors  but  yet  exhibits  a 
marked  effect  at  10  mM.  Koishi  (1959  a)  confirmed  the  inhibition  by  mal- 
onate on  p-aminohippurate  accumulation  in  rat  kidney  slices,  obtaining 
slight  inhibition  at  1  mM  and  65%  inhibition  at  5  mM.  The  effects  of  mal- 
onate are  expressed  in  the  following  equation: 

log  (S/M)  =  1.452  +  0.457  log  (/)  (1-6) 

where  S/M  is  the  slice/medium  ratio  and  (/)  is  the  molar  concentration  of 
malonate.  The  active  transport  of  p-aminohippurate  is  thus  definitely  relat- 
ed in  some  manner  to  succinate  dehydrogenase  and  the  cycle,  assuming  a 
specific  action  of  malonate,  which  is  likely  at  the  generally  low  concentra- 
tions used. 

This  conclusion  is  somewhat  substantiated  by  the  findings  that  the  renal 
transport  of  other  substances  is  frequently  not  inhibited  potently  by  mal- 
onate. The  accumulation  of  tetraethylammonium  ion  is  scarcely  affected 
by  malonate  up  to  40  mM  (Farah  and  Rennick,  1956;  Farah,  1957),  the 
metabolic  requirements  apparently  being  different  than  for  jj-aminohippu- 

*  Tm  is  the  tubular  transport  maximal  rate  for  a  substance. 


CELLULAR    AND    TISSUE    FUNCTION 


205 


rate.  Phenol  red  accumulation  is  also  not  depressed  (Shideman  and  Rene, 
1951  b),  although  in  isolated  flounder  tubules  it  is  suppressed  by  10-20  roM 
malonate  (Forster  and  Goldstein,  1961).  It  was  claimed  that  there  is  a  cor- 
relation between  succinate  oxidase  activity  and  transport  in  different  spe- 
cies. Clearance  studies  with  glucose  and  phosphate  show  that  there  is  little 
effect  of  malonate  on  their  transport:  e.g.,  when  p-aminohippurate  trans- 
port is  inhibited  73%,  glucose  Tm  is  decreased  only  10%.  On  the  other 
hand,  Malvin  (1956)  reported  that  malonate  quite  definitely  depresses  the 
phosphate  Tm,  although  no  data  were  given.  Since  malonate  interferes 
with  the  uptake  of  inorganic  phosphate  in  kidney  homogenates,  it  was 
concluded  that  malonate  in  some  manner  suppresses  the  esterification  of 
phosphate  during  its  transport. 


Fig.  1-18.  Effects  of  inhibitors  on  the  slice/medium  (S/M)  ratio 
for  p-aminohippurate  in  kidney  shces.  DNP  =  2,4-dinitrophenol, 
FA  =  fluoroacetate,  lA  =  iodoacetate,  DHA  =  dehydroacetate, 
CN  =  cyanide,  and  F  =  fluoride.  (Modified  from  Farah  and 
Rennick,    1956.) 


Renal  electrolyte  transport  is  disturbed  by  malonate  (Mudge,  1951). 
Rabbit  renal  cortex  slices  were  leached  for  2.5  hr  in  0.15  M  NaCl,  this 
lowering  the  tissue  K+  concentration  and  reducing  the  endogenous  respir- 
ation. The  slices  were  then  incubated  for  30  min  in  medium  containing 
10  mM  K+,  10  mM  acetate,  NaCl  to  provide  a  constant  osmotic  pressure, 
and  phosphate  buffer,  during  which  time  K+  enters  the  cells.  Malonate  at 
50  mM  almost  completely  blocks  this  return  of  K+  into  the  cells  and  re- 
verses the  movement  of  water  (see  accompanying  tabulation).  The  Na+ 
changes  are  not  so  significant  because  the  substitution  of  divalent  anions 
for  chloride  increases  the  external,  and  presumably  the  internal,  Na+  con- 
centration. If  fresh  slices  are  treated  with  malonate,  a  loss  of  cell  K+  and 


206  1.    MALONATE 


Initial  Control       Malonate  50  xnM      %  Change 


espiration  (Qo^) 

— 

0.69 

0.37 

'^ater  content  (%) 

79.0 

78.2 

81.1 

+  (meq/kg  wet  wt.) 

24.7 

58.0 

28.1 

a+  (meq/kg  wet  wt.) 

112 

93.8 

139 

+  +  Na+ 

136.7 

151.8 

167.1 

-  46 

-  90 
+  100 


an  equivalent  gain  of  Na+  would  be  expected  due  to  the  inhibition  of  the 
transport  mechanism  responsible  for  K+  accumulation.  The  effect  on  water 
transport  is  really  quite  marked  and  greater  than  with  some  fifty  other 
inhibitors  used.  Mudge  suggested  that  malonate  acts  by  depression  of 
aerobic  metabolism  in  general  and  not  necessarily  by  a  specific  inhibition 
of  succinate  oxidase,  which  at  the  high  concentration  used  is  quite  possible. 
The  results  on  malonate  in  vivo  occasionally  do  not  correspond  to  the 
in  vitro  experiments,  nor  do  they  always  correspond  to  each  other.  The  in- 
jection of  10  millimoles/kg  malonate  into  rats  leads  to  a  considerable  diu- 
resis which  lasts  for  several  days  (Angielski  et  al.,  1960  a).  On  the  other 
hand,  infusion  of  malonate  into  the  renal  artery  of  a  dog  at  a  concentration 
of  8.7  mM  causes  no  significant  change  in  creatine,  p-aminohippurate,  or  Na+ 
clearances,  and  produces  a  15%  suppression  of  urinary  volume  (Strickler 
and  Kessler,  1963).  The  effects  in  intact  animals  are  related  to  acid-base 
imbalance  in  addition  to  direct  renal  action,  as  shown  by  unpublished  ex- 
periments by  Goldberg  (1963)  in  which  rats  were  water  loaded  and  received 
17  millimoles/kg  sodium  malonate  with  17  ml  10%  mannitol/kg,  this  pro- 
ducing certain  toxic  symptoms  (e.g.,  respiratory  difiiculty,  sluggishness, 
and  mild  cyanosis).  The  results  are  summarized  in  the  accompanying  tab- 
ulation and  it  is  clear  that  a  systemic  acidosis  was  produced.  A  rather 

Determination  Control  Malonate 


Urinary  flow  (ml/min)  0.054  0.058 

Urinary  pH 

Titratable  acidity  (meq/liter) 

Creatinine  clearance  (ml/min) 

Na+  excretion  (meq/liter) 

K+  excretion  (meq/liter) 

NH4+  excretion  (meq/liter) 

Plasma  pH 

Plasma  Na+  (meq/liter) 

Plasma  K+  (meq/liter) 


6.72 

5.45 

17.0 

46.7 

1.1 

0.89 

9.5 

110 

10.3 

57.5 

25.7 

16.3 

7.39 

7.23 

50 

198 

5.4 

9.5 

CELLULAR    AND    TISSUE    FUNCTION  207 

clear  increase  in  phosphate  excretion  with  a  fall  in  plasma  phosphate  was 
also  observed. 

The  effects  of  malonate  on  the  renal  excretion  of  malate  are  interesting, 
but  it  is  not  known  if  this  phenomenon  is  related  to  an  action  on  transport 
mechanisms  or  to  more  general  metabolic  effects.  Vishwakarma  (1957, 
1962)  showed  that  malonate  has  no  effect  on  the  excretion  df  malate  in- 
duced by  malate  infusion.  However,  when  succinate  is  infused,  the  in- 
creased malate  excretion  is  due  to  both  an  increased  filtration  and  a  marked 
tubular  secretion.  Malonate  inhibits  the  tubular  secretion  of  malate  and 
converts  the  excretion  to  a  purely  filtration  process  or  causes  a  net  resorp- 
tion. Vishwakarma  and  Lotspeich  (1960)  continued  this  study  in  chickens 
and  found  that  when  malonate  is  infused  with  succinate,  instead  of  block- 
ing the  formation  of  malate,  it  further  increases  the  malate  excretion. 
Malonate  was  infused  at  a  rate  of  about  6.8  //moles/kg/min.  This  could 
mean  that  malonate  (1)  enhances  the  formation  of  malate  from  succinate, 
(2)  facilitates  the  tubular  secretion  of  malate,  or  (3)  gives  rise  to  malate  by 
metabolic  conversion.  Since  malonate  infused  alone  did  not  significantly 
increase  malate  excretion,  the  last  explanation  is  unlikely.  In  the  dog,  mal- 
onate inhibits  the  tubular  secretion  rather  than  stimulating  it  and  does  not 
inhibit  the  resorption  of  malate.  The  mechanism  for  this  paradoxical  effect 
is  unknown.  Reference  may  be  made  to  studies  of  Lotspeich  and  Woron- 
kow  (1964),  who  unilaterally  perfused  chicken  kidneys  and  found  complex 
effects  on  the  excretions  of  various  organic  acids,  and  concluded  that  the 
cycle  must  be  involved  in  some  manner. 

Transintestinal  Transport 

Quastel  has  studied  the  effects  of  various  inhibitors  on  the  transfer  of 
glucose  across  the  guinea  pig  intestinal  wall.  This  is  an  active  transport 
and  depends  strongly  on  the  aerobic  metabolism  (as  shown  by  the  marked 
inhibition  by  cyanide  and  azide)  and  the  associated  phosphorylations  (as 
shown  by  the  2,4-dinitrophenol  inhibition).  When  malonate  at  20  mM  is 
present  in  the  lumen,  there  is  18.5%  inhibition  of  the  glucose  transported, 
but  if  malonate  is  present  both  inside  and  outside,  the  inhibition  is  44.3% 
(Darlington  and  Quastel,  1953).  An  increase  of  K"'"  from  6  to  15.6  n\M  ac- 
celerates glucose  transport  about  50%.  Malonate  inhibits  the  K^-stimulated 
transport  completely  at  concentrations  as  low  as  2  mM  (Riklis  and  Quastel, 
1958).  This  result  may  be  related  to  the  greater  sensitivity  of  K+-stimulated 
brain  slices  to  malonate.  It  was  claimed  that  20  mM  malonate  depresses 
both  the  accumulation  of  L-monoiodotyrosine-I^^^  in  the  intestinal  cells  and 
its  transport  across  the  intestine,  but  no  data  were  given  (Nathans  et  al., 
1960).  There  is  a  marked  difference  between  transintestinal  transport  and 
tissue  accumulation  of  triiodothyroacetate,  the  former  being  inhibited  much 


208  1.    MALONATE 

more  readily  by  a  number  of  substances;  malonate  at  10  mM  inhibits  up- 
take 16%  and  transport  70%  (Herz  et  al.,  1961). 

Everted  segments  of  the  rat  duodenum  transport  iron  from  the  mucosal 
to  the  serosal  surface  against  concentration  gradients  and  this  process  is 
dependent  on  oxidative  phosphorylations  (Dowdle  et  al.,  1960).  Malonate 
at  50  mM  reduces  the  inside/outside  ratio  of  Fe++  from  4.0  to  0.6.  Ca++  is 
also  transported  actively  and  malonate  at  20  mM  reduces  the  inside/outside 
ratio  of  Ca*^  from  5.0  to  2.5  (Schachter  and  Rosen,  1959).  Ca++  is  also  ac- 
cumulated by  the  intestinal  cells,  the  tissue/medium  ratio  being  5.8,  which 
is  decreased  by  20  mM  malonate  to  3.4  (Schachter  et  al.,  1960).  The  question 
of  the  chelation  of  Ca++  and  Fe+++  by  the  malonate  arises,  since  the  high 
malonate  concentrations  would  certainly  reduce  the  free  ions  appreciably. 
This  must  play  some  role  but  in  the  case  of  Ca++  cannot  explain  the  reduc- 
tion in  the  transport,  inasmuch  as  the  inside/outside  ratio  is  increased  as 
the  Ca++  concentration  is  lowered. 

Gastric  Acid  Secretion 

The  secretion  of  hydrochloric  acid  by  the  parietal  cells  is  dependent  on 
oxidations  and  the  formation  of  ATP,  since  it  is  strongly  inhibited  by  cya- 
nide, antimycin,  and  2.4-dinitrophenol.  However,  the  secretion  in  isolated 
rat  stomachs  is  not  affected  by  10  mM  malonate  (Patterson  and  Stetten, 
1949).  Injection  of  malonate  in  mice  subcutaneously  inhibits  the  accumu- 
lation of  p-aminohippurate  in  the  kidney  but  does  not  inhibit  acid  secretion: 
4.8  millimoles  of  malonate  reduces  the  kidney/medium  p-aminohippurate 
ratio  from  6.1  to  2.9  but  inhibits  the  secretion  of  hydrochloric  acid  only 
6%  (Davenport  and  Chavre,  1956).  This  is  near  the  fatal  dose  of  malonate 
and  many  of  the  mice  did  not  live  long  enough  to  perform  the  test.  Inas- 
much as  succinate  oxidase  activity  is  high  in  the  stomach  and  is  readily 
inhibited  by  malonate,  and  since  fluoroacetate  inhibits  acid  secretion,  the 
most  likely  explanation  for  the  lack  of  a  malonate  effect  is  a  failure  to 
penetrate  into  the  parietal  cells  sufficiently.  Some  evidence  for  the  partici- 
pation of  succinate  oxidase  in  acid  secretion  was  obtained  by  Vitale  et  al. 
(1956),  who  showed  that  stimulation  of  guinea  pig  or  human  gastric  mucosa 
with  histamine  leads  to  significant  increases  in  the  succinate  oxidase  activity, 
although  histamine  has  no  such  effect  in  liver  or  duodenum.  Furthermore, 
succinate  oxidase  is  concentrated  in  those  regions  of  the  stomach  where 
the  parietal  cells  are  abundant. 

Active  Transport  of  Ions  in  Various  Cells  and  Tissues 

The  effects  of  malonate  on  nerve  and  muscle,  to  be  discussed  in  the  fol- 
lowing sections,  depend  in  part  on  the  modification  of  the  active  transport 
of  ions  in  these  tissues.  Malonate  depresses  many  types  of  active  transport. 


CELLULAR    AND    TISSUE    FUNCTION  209 

as  we  have  seen  for  kidney  and  gastric  mucosa,  and  the  mechanism  may  be 
either  a  simple  reduction  in  the  energy  supply  or  a  more  direct  interference 
with  electron  transport  associated  with  ionic  movements  across  the  mem- 
brane. One  must  also  attempt  to  distinguish  between  effects  on  the  active 
transport  and  increases  in  permeability.  If  the  permeability  to  an  ion  is 
significantly  increased,  its  intracellular  accumulation  may  drop  because  the 
ion  pump  is  no  longer  able  to  maintain  the  normal  concentration;  such  an 
action  would  appear  superficially  to  be  an  inhibition  of  active  transport. 

Malonate  up  to  10  mM  has  no  effect  on  the  transport  of  Na+  and  K+ 
across  the  human  erythrocyte  membrane  (Maizels,  1951),  which  is  not  sur- 
prising since  the  principal  energy  source  in  such  cells  is  glycolytic.  In  ascites 
tumor  cells,  substrates  (for  example,  glucose  and  succinate)  increase  the 
efflux  of  Na+.  Cyanide  at  a  concentration  inhibiting  respiration  70%  has 
no  effect  on  either  the  influx  or  efiiux  of  Na+,  presumably  because  the  rate 
of  aerobic  glycolysis  is  simultaneously  doubled,  this  compensating  for  the 
oxidative  depression  (Maizels  et  al.,  1958).  Malonate  at  12.5  mM,  on  the 
other  hand,  inhibits  respiration  35%  but  produces  only  a  5%  increase  in 
the  glycolysis,  which  may  explain  why  the  rate  coefficient  for  Na+  efflux 
drops  from  6.5  to  5.1  hr-^  The  accumulation  of  intramitochondrial  K+  in 
preparations  from  rabbit  heart  is  dependent  on  oxidative  phosphorylation 
but  is  unaffected  by  0.2  mM  malonate  (Ulrich,  1960).  Inasmuch  as  a-keto- 
glutarate  was  the  substrate  used,  even  a  complete  block  of  succinate  oxida- 
tion might  not  be  expected  to  have  much  effect  on  ion  movements  because 
sufficient  ATP  may  be  generated  in  the  single-step  oxidation  of  a-keto- 
glutarate.  Ca++  uptake  and  binding  by  kidney  mitochondria  depend  on  an 
oxidizable  substrate  and  ATP;  it  is  depressed  75%  by  10  mM  malonate, 
which  suggests  interference  with  the  operation  of  the  cycle,  but  could 
relate  to  the  chelation  of  the  Ca++  by  the  malonate  (Vasington  and  Murphy, 
1962).  The  uptake  of  iodide  is  inhibited  by  rather  high  concentrations  of 
malonate  in  the  brown  alga  Ascophyllum,  nodosum  (79%  inhibition  at  25  mM) 
(Kelly,  1953),  the  rabbit  ciliary  body  (50%  inhibition  at  50  mM)  (Becker, 
1961),  and  the  rabbit  choroid  plexus  (50%  inhibition  at  20  mM)  (Welch, 
1962),  but  1  mM  malonate  has  no  effect  on  the  uptake  or  incorporation  of 
iodide  in  sheep  thyroid  particulate  fractions  (Tong  et  al.,  1957). 

The  accumulation  of  P,^^  in  the  roots  of  the  loblolly  pine  Pinus  taeda 
during  a  3  hr  incubation  is  inhibited  5%  at  pH  4.75  but  stimulated  54% 
at  pH  5.75  (Kramer,  1951).  Similar  results  are  obtained  in  mycorrhizal 
root  tips  but  the  inhibition  is  somewhat  greater.  It  is  possible  that  at  the 
concentration  (25  mM)  of  malonate,  the  stimulation  is  an  ionic  effect  which 
is  partially  counteracted  by  a  malonate  inhibition  at  the  lower  pH.  The 
uptake  of  K+  and  Br~  by  barley  roots  is  quite  strongly  inhibited  by  mal- 
onate at  pH  4.5  (see  accompanying  tabulation)  (Ordin  and  Jacobson,  1955). 
The  inhibition  is  overcome  to  some  extent  by  malate  and  fumarate;  sue- 


210  1.    MALONATE 

cinate,  however,  actually  increases  the  inhibition.  It  is  likely  in  this  tissue 
that  ion  accumulation  is  obligatorily  coupled  to  the  operation  of  the  cycle. 


Malonate 

%  Inhibition  of: 

(mM) 

K+  uptake 

Br-  uptake 

Respiration 

5 

10 

39 

92 

42 

70 

30 
55 

Effect  of  Malonate  on   Mitochondrial   Swelling 

Eat  liver  mitochondria  swell  quite  readily,  as  measured  by  changes  in 
light  scattering  or  optical  density,  when  treated  with  various  substances, 
and  the  effects  of  malonate  on  this  phenomenon  are  interesting.  Raaflaub 
(1953)  established  that  succinate  and  phosphate  promote  swelling.  This 
swelling  is  counteracted  by  ATP  in  both  cases,  but  malonate  prevents  the 
swelling  from  succinate  only.  This  was  confirmed  by  Tapley  (1956),  who 
extended  the  list  of  substances  causing  swelling  to  fumarate,  malate,  gluta- 
mate,  acetate,  and  a-ketoglutarate.  Swelling  is  prevented  by  citrate,  pyru- 
vate, and  oxalacetate,  as  well  as  malonate.  Since  malonate  can  prevent  the 
swelling  from  substrates  other  than  succinate,  there  is  some  question  as  to 
the  specificity  of  the  effect.  It  was  claimed  that  the  same  results  are  obtain- 
ed in  the  absence  of  oxygen  and  thus  that  the  swelling  is  not  related  to  the 
utilization  of  these  substrates.  Quite  different  conclusions  were  reached  by 
Chappell  and  GreviUe  (1958)  inasmuch  as  they  found  a  good  correlation 
between  swelling  and  utilizable  substrates.  Malonate  prevents  the  swelling 
from  succinate  but  not  from  a-hydroxybutyrate,  and,  in  general,  inhibitors 
blocking  oxidations  reduced  swelling.  Matters  were  further  complicated  by 
the  results  of  Keller  and  Lotspeich  (1959  b).  They  found  that  phlorizin 
caused  swelling  of  kidney  mitochondria  and  that  this  could  be  counteracted 
by  Mg++,  ATP,  2,4-dinitrophenol,  and  malonate.  Hunter  et  al.,  (1959  a,  b) 
considered  the  possibility  that  swelling  is  related  to  the  fraction  of  NAD  in 
the  oxidized  form,  since  amobarbital  prevents  oxidation  of  NADH  and 
prevents  swelling.  However,  succinate  in  the  presence  of  amobarbital  causes 
swelling  and  this  is  blocked  by  malonate.  Glutamate-induced  swelling  is  not 
prevented  by  malonate.  It  was  concluded  that  swelling  depends  on  electron 
flow  between  the  substrate  and  oxygen,  and  whether  or  not  an  inhibitor  will 
prevent  swelling  is  determined  by  where  in  the  electron  transport  chain 
the  substrate  and  the  inhibitor  act.  This  does  not  very  well  explain  the 
prevention  of  swelling  by  2,4-dinitrophenol  and  it  was  suggested  that  there 
are  at  least  two  different  types  of  mitochondrial  swelling.  Further  confusion 
was  introduced  by  Sabato  and  Fonnesu  (1959),  who  found  that  swelling  is 


CELLULAR    AND    TISSUE    FUNCTION  211 

prevented  by  oxidizable  substrates  such  as  succinate,  glutamate,  and  a- 
ketoglutarate,  and  that  malonate  counteracts  this  preventive  action.  These 
results  were  confirmed  by  Kaufman  and  Kaplan  (1960),  who  again  ob- 
served, in  contrast  to  the  earlier  workers,  that  succinate  inhibits  swelling 
and  malonate  reverses  this  protection.  They  believe  that  swelling  is  correl- 
ated with  the  mitochondrial  release  of  pyridine  nucleotides  (see  tabulation). 


Pyridine  nucleotide  released 
(//g/20  min) 


Optical  density 


No  substrate  64  —0.710 

Succinate  (20  mM)  14  —0.095 

Malonate  (20  mM)  68  -0.740 

Succinate  +  malonate  38  — 0.520 


Succinate  reduces  the  loss  of  the  pyridine  nucleotides  and  malonate  anta- 
gonizes this  effect.  One  must  conclude  that  there  must  be  different  mecha- 
nisms of  swelling  and  that  the  mitochondrial  behavior  perhaps  depends  on 
the  metabolic  state  and  the  nature  of  the  suspension  medium.  The  effects 
of  malonate  and  other  anions  on  the  concentrations  of  free  Ca++  and  Mg++ 
should  also  not  be  ignored,  inasmuch  as  EDTA  has  usually  been  shown  to 
modify  the  swelling. 

Conduction    and    Membrane    Potentials   of  Nerve 

Penetration  of  malonate  into  nerve  axons  in  the  physiological  pH  range 
must  be  very  poor.  This  may  account  for  the  failures  of  Shanes  and  Brown 
(1942)  to  observe  an  eft'ect  of  20  mM  malonate  on  the  resting  potential  of 
frog  nerve,  and  of  Greengard  and  Straub  (1962)  to  find  an  effect  of  10  mM 
malonate  on  nonmyelinated  nerve  posttetanic  hyperpolarization,  despite 
the  fact  that  this  phenomenon  is  quite  sensitive  to  other  inhibitors.  How- 
ever, Jenerick  (1957)  reported  some  effect  of  10  mM  malonate  on  frog  sciatic 
nerve  although  the  action  was  presumably  slow  in  developing.  When  the 
action  potential  spike  amplitude  is  reduced  by  80-90%,  the  threshold  for 
stimulation  begins  to  rise  rapidly.  Conduction  block  occurs  when  the  rest- 
ing potential  has  fallen  by  30-40%.  It  is  doubtful  if  the  decrease  in  external 
Ca++  concentration,  which  was  1.3  mM  initially,  resulting  from  chelation 
by  malonate  could  be  held  responsible  for  these  effects,  and  it  was  felt 
that  metabolic  interference  must  occur.  The  preganglionic  and  postgan- 
glionic action  potentials  in  preparations  of  cat  sympathetic  ganglia  are 
depressed  equally  (75-80%)  by  14  mM  malonate  and  transmission  through 
the  ganglia  is  reduced  (Larrabee  and  Bronk,  1952).  The  excitability  of  the 


212  1.    MALONATE 

isolated  cat  carotid  body  is  lowered  by  perfusion  with  malonate  (Anichkov, 
1953).  These  meager  data  are  all  we  have  to  understand  the  actions  of  mal- 
onate on  nerve  function.  Unfortunately,  little  has  been  done  on  junctional 
transmission,  inasmuch  as  it  might  be  predicted  that  the  synapses  would 
be  more  sensitive  to  malonate  than  are  the  axons,  because  of  both  a  higher 
permeability  of  such  regions  to  anions  and  a  greater  energy  requirement 
for  the  synthesis  of  acetylcholine. 

Skeletal  and  Smooth  Muscle  Function 

Essentially  nothing  is  known  of  the  effects  of  malonate  on  skeletal  muscle. 
Beckmann  (1934)  claimed  that  6.7  mM  malonate  causes  a  swelling  of  muscle, 
indicating  an  alteration  of  permeability.  This  was  termed  a  membrane- 
loosening  effect.  In  the  initial  work  of  Ling  and  Gerard  (1949)  with  intra- 
cellular microelectrodes,  it  was  observed  that  10  toM  malonate  drops  the 
resting  potential  of  frog  sartorius  muscle  from  78  mv  to  65.3  mv  over  a 
period  of  3  hr.  This  may  be  correlated  with  the  suppression  of  Na+  ex- 
trusion observed  by  Kernan  (1963)  in  the  same  muscle,  30%  inhibition 
being  produced  by  1  mM  malonate  over  2  hr,  an  effect  similar  to  that  oc- 
curring in  brain  slices  (Bilodeau  and  Elliott,  1963).  No  direct  work  on  the 
contractile  response  to  malonate  has  been  done. 

The  contractions  of  isolated  rabbit  intestine  are  not  inhibited  by  10  mM 
malonate,  whether  in  the  absence  of  substrate  or  in  the  presence  of  either 
acetate  or  glucose  (Weeks  and  Chenoweth,  1950;  Weeks  et  al.,  1950).  In- 
deed, there  is  a  tendency  for  malonate  to  increase  the  contractile  activity 
slightly,  especially  with  glucose  as  the  substrate.  There  is  also  no  interfer- 
ence with  the  recovery  of  substrate-depleted  strips  produced  by  the  ad- 
dition of  acetate  or  pyruvate.  It  was  suggested  that  a  lack  of  penetration 
of  malonate  into  the  smooth  muscle  cells  might  be  responsible.  Fluoroacetate 
is  quite  inhibitory  under  the  same  conditions  so  that  some  relationship  of 
the  contractility  to  the  cycle  is  likely.  The  contractile  properties  of  the 
vascular  smooth  muscle  in  the  cat  hind  limb  are  not  affected  by  1  mM 
malonate  (Hitchcock,  1946),  and  the  behavior  of  electrically  stimulated  pig 
carotid  artery  is  not  altered  by  10  mM  malonate  (Jacobs,  1950). 

It  would  be  important  to  know  more  about  the  possible  effects  of  mal- 
onate on  the  formation  and  release  of  the  neurohormones,  such  es  acetyl- 
choline and  the  catecholamines,  but  the  data  are  not  available.  It  is  in- 
teresting to  note,  however,  that  malonate  is  reasonably  effective  in  inhi- 
biting the  release  of  histamine  from  guinea  pig  lung  slices  during  an  ana- 
phylactic reaction  (Moussatche  and  Prouvost-Danon,  1958).  The  inhibition 
is  10%  at  20  mM,  40%  at  40  mM,  and  50%  at  60  mM.  The  inhibition  was 
attributed  to  the  effect  on  succinate  dehydrogenase.  Nevertheless,  malonate 
at  40-60  ToM  has  virtually  no  effect  on  the  release  of  histamine  brought 
about  by  the  application  of  the  histamine-releaser  Compound  48/80  (Mous- 


CELLULAR    AND    TISSUE    FUNCTION  213 

satche  and  Prouvost-Danon,  1957),  although  respiration  of  the  lung  slices 
is  depressed  fairly  strongly.  It  would  appear  that  malonate  interferes  with 
the  anaphylactic  release  of  histamine  by  a  mechanism  other  than  a  direct 
effect  on  the  formation  or  release  of  histamine.  It  may  be  noted  that  suc- 
cinate accelerates  the  oxygen  uptake  but  has  no  effect  on  the  release  of 
histamine. 

Cardiac  Membrane  Potentials  and   Function 

The  physiological  disturbances  produced  by  malonate  have  been  most 
thoroughly  studied  in  the  heart.  Although  the  effects  are  often  very  slight, 
despite  the  evident  importance  of  the  cycle  in  the  myocardium,  under  cer- 
tain conditions  the  responses  to  malonate  are  very  interesting.  The  earliest 
investigation  was  made  by  Forssman  and  Lindsten  (1946)  at  Lund,  who 
noted  a  marked  discrepancy  between  the  effects  of  malonate  in  the  whole 
animal  and  on  isolated  hearts.  Intravenous  injections  of  malonate  at  doses 
around  3.7-7.5  millimoles/kg  to  cats  and  rabbits  lead  to  an  increase  in  the 
venous  blood  pressure  and  a  fall  in  the  arterial  blood  pressure,  indicating 
cardiac  depression.  In  rabbits  the  cardiac  failure  begins  about  20  min  after 
the  injection  whereas  in  cats  the  changes  are  immediate.  In  rabbits  the 
heart  may  stop  after  40  min  but  in  cats  recovery  is  the  rule.  At  autopsy 
the  heart  is  found  to  be  dilated.  The  effects  of  malonate  on  the  isolated 
perfused  rabbit  heart,  however,  are  rather  small  and  inconsistent  (see  ac- 
companying tabulation).  Moreover,  succinate  at  the  same  concentrations 

Malonate  °''»  ^^^^^^  ^^ 


(mM) 

Amplitude 

Coronary  flow 

Rate 

10 

-14 

-23 

-  5 

20 

-19 

-16 

-  8 

40 

-34 

-29 

0 

acts  similarly.  These  are  the  immediate  effects  of  malonate  and  it  is  possible 
that  the  heart  would  recover  from  this  depression  after  several  minutes,  as 
do  rabbit  atria  (Webb,  1950).  It  is  doubtful  if  these  effects  are  related  to 
inhibition  of  succinate  oxidase;  they  are  more  likely  ionic  actions  on  the 
membrane.  The  reduction  in  the  coronary  flow  may  result  from  a  vascular 
constriction,  but  is  more  probably  the  response  to  the  decreased  functional 
activity. 

Isolated  rabbit  atria  are  depressed  only  by  high  concentrations  of  mal- 
onate, 30-40  mM  producing  a  30%  decrease  in  contractile  amplitude  and 
rate  at  2  min;  the  depression  is  temporary,  complete  recovery  being  ob- 
served after  8-10  min  (Webb,  1950b).  Atria  can  continue  to  beat  normally 


214  1.    MALONATE 

for  hours  in  50  mikf  malonate.  The  temporary  depression  brought  about 
by  malonate  is  not  counteracted  by  fumarate  added  either  before,  with,  or 
after  the  malonate.  In  fact,  fumarate,  along  with  pyruvate,  acetate,  suc- 
cinate, and  malate,  has  an  action  very  similar  to  that  of  malonate  on  the 
atria.  It  is  not  known  if  this  inhibition  and  recovery  are  related  to  the 
somewhat  slower  but  similar  time  course  of  ventricular  respiration  under 
the  influence  of  malonate  (page  181)  (Webb  et  al,  1949).  The  depression  is 
not  due  to  chelation  of  Ca++  or  Mg++  since  reduction  in  the  concentrations 
of  these  ions  produces  a  different  response.  Gardner  and  Farah  (1954)  con- 
firmed the  resistance  of  rabbit  atria  to  malonate,  finding  that  10-20  milf 
has  no  significant  effects  on  contractility,  spontaneous  rate,  excitability 
threshold,  refractory  period,  and  conduction  rate. 

The  effects  of  malonate  were  investigated  more  thoroughly  on  rat  atria 
(Webb  and  Hollander,  1959).  The  contractility  is  depressed  21%  imme- 
diately but  slow  recovery  occurs:  the  inhibition  is  13%  during  5-25  min, 
9%  during  25-45  min,  and  5%  during  45-60  min.  The  malonate  concen- 
tration used  was  15  mM.  The  addition  of  15  mM  NaCl  produces  a  rapid 
contractile  depression  about  half  as  great  as  from  malonate,  so  that  at  least 
part  of  the  initial  malonate  effect  is  attributable  to  the  Na+  ion.  A  slight 
initial  rise  in  the  magnitude  of  the  action  potential  is  observed  with  both 
malonate  and  NaCl,  but  in  the  case  of  malonate  this  is  soon  replaced  by  a 
small  depression.  There  is  also  some  shortening  of  the  action  potential  and 
a  decrease  in  its  area  after  the  first  5  min,  which  could  be  responsible  for 
the  fall  in  tension.  In  summary,  the  addition  of  15  ml/  malonate  produces 
a  rapid  initial  effect  attributable  mainly  to  the  Na+  and  this  is  progressively 
replaced  by  changes  due  to  the  malonate,  these  latter  changes  being  mod- 
erate depressions  of  the  action  potential  and  contractility.  The  importance 
of  the  cycle  in  the  atrial  function  is  indicated  by  the  marked  changes  brought 
about  by  fluoroacetate,  and  thus  the  resistance  to  malonate  is  probably  due 
to  a  low  intracellular  concentration  of  malonate.  Greater  effects  on  the  con- 
tractility of  rat  atrium  were  observed  by  Venturi  and  Schoepke  (1960), 
5  mM  depressing  22%,  10  mM  44%,  and  20  mM  90%.  It  was  stated  that 
NaCl  at  these  concentrations  does  not  alter  the  contractility.  Furthermore, 
succinate  is  as  inhibitory  as  malonate.  The  greater  depression  observed  here 
compared  to  my  work  is  difficult  to  explain.  Venturi  and  Schoepke  used 
Locke  solution  at  pH  7  whereas  I  used  Krebs-Ringer-bicarbonate  medium 
at  pH  7.4.  Part  of  the  larger  inhibition  seen  by  Venturi  and  Schoepke  thus 
might  be  due  to  the  lower  pH.  In  any  event,  these  effects  seem  to  be  un- 
related to  the  inhibition  of  succinate  oxidase  and  again  must  be  attributed 
to  some  action  directly  on  the  membrane.  Venturi  and  Schoepke  found  that 
increasing  Ca++  concentration  can  completely  overcome  the  depressant  ac- 
tions of  malonate,  succinate,  and  the  other  organic  anions  used,  leading 
them  to  suggest  that  the  negative  inotropic  action  is  due  to  the  chelation 


CELLULAR    AND    TISSUE    FUNCTION 


215 


of  Ca+"'".  However,  there  are  some  reservations  in  accepting  this  explanation 
completely.  In  the  first  place,  Ca++  stimulates  atrial  contractility  and  would 
be  expected  to  counteract  most  depressants  in  a  nonspecific  manner.  In 
the  second  place,  reducing  the  Ca++  from  1.22  mM  to  0.91  mM  does  not 
alter  the  contractility,  although  further  reduction  to  0.61  mM  depresses 
43%.  Malonate  at  15  mM  would  reduce  the  Ca++  from  1.22  mM  to  0.82  mM 
and  a  small  contractile  depression  may  result  from  this.  The  total  Ca++  in 
Locke  solution  is  2.16  mM  and  20  mM  malonate  would  reduce  the  free 
Ca++  to  1.31  mM,  which  alone  could  not  produce  the  90%  depression  seen 
by  Venturi  and  Schoepke.  In  the  third  place,  monocarboxylate  ions,  such 
as  acetate,  lactate,  and  pyruvate,  at  20  mM  depress  the  contractile  am- 
plitude 40-45%  and  these  do  not  deplete  the  Ca++. 

The  modifications  in  the  electrocardiogram  following  intravenous  injec- 
tions of  malonate  into  turtles  were  studied  by  Lenzi  and  Caniggia  (1953). 
At  a  dose  of  4.4  millimoles/kg  malonate  the  following  occurred:  brady- 
cardia, slowing  of  the  a-v  conduction,  a  tendency  for  the  shortening  of 
repolarization  and  electric  systoles,  with  eventually  a  total  a-v  block 
and  a  prolongation  of  the  depolarization  time  (see  accompanying  tabula- 
tion). Pacemaker  and  conduction  depression  are  thus  evident,  and  it  is 


Control 

Malonate 
at  30-35  min 

Control 

at 

Malonate 
57-65   min 

Rate 

50 

23 

78 

32 

pq    interval 

0.30 

0.465 

0.24 

block 

qrs   interval 

0.15 

0.18 

0.10 

0.18 

st-t  interval 

0.57 

0.895 

0.44 

1.06 

qt     interval 

0.72 

1 .  075 

0.54 

1.24 

quite  possible  that  similar  changes  would  be  observed  in  mammals,  con- 
sidering the  general  behavior  of  the  heart  in  cats  and  rabbits  treated  with 
malonate  (Forssman  and  Lindsten,  1946).  The  electrocardiogram  from  the 
embryonic  chick  heart  is  not  altered  by  4  mM  malonate  (Boucek  and  Paff, 
1961). 

In  contrast  to  the  depressant  effects  of  malonate  on  the  whole  heart  and 
isolated  atria,  the  rat  ventricle  strip  is  usually  strongly  stimulated,  as  first 
noticed  by  Masuoka  et  al.  (1952).  Substrate-depleted  and  hypodynamic  strips 
recover  to  almost  the  initial  contractile  amplitude  upon  addition  of  10  mM 
malonate  at  pH  6.2  (which  was  used  to  facilitate  penetration  of  the  mal- 
onate), and  simultaneously  the  stimulatory  action  of  succinate  is  blocked. 
The  ability  of  glucose  to  induce  recovery  is  augmented  by  malonate,  and 
that  of  pyruvate  is  slightly  reduced  (Berman  and  Saunders,  1955).  This 
interesting  positive  inotropic  action  was  analyzed  in  detail  because  it  was 


216  1.    MALONATE 

felt  that  such  an  action  might  have  bearing  on  the  mechanisms  whereby 
the  cardioactive  glycosides  stimulate  the  failing  heart.  Only  the  major 
results  will  be  summarized  here.  The  positive  inotropic  action  occurs  most 
strongly  when  glucose  is  present,  less  in  substrate-free  medium,  and  not  at 
all  with  pyruvate  or  a-hydroxybutyrate  as  substrate  (Covin  and  Berman, 
1956).  These  results  suggested  that  malonate  might  stimulate  the  Embden- 
Meyerhof  glycolytic  pathway,  resulting  in  an  accelerated  conversion  of  glu- 
cose and  glycogen  to  pyruvate.  If  this  were  so,  pyruvate  should  produce  a 
comparable  positive  inotropic  effect  and  it  does  in  both  substrate-depleted 
and  glucose-supplemented  strips.  Furthermore,  iodoacetate  at  0.2  mM 
blocks  the  stimulation  by  malonate,  whereas  it  does  not  affect  the  response 
to  pyruvate.  The  chelation  of  Ca++  was  shown  to  contribute  to  the  depression 
produced  by  malonate  at  high  concentrations  (20-50  mM),  and  it  probably 
reduces  the  amount  of  stimulation  seen  at  the  lower  concentrations  since 
lowering  the  Ca++  to  the  degree  calculated  to  occur  in  10  mM  malonate 
depresses  the  contractile  activity  23%.  The  effects  of  malonate  on  the  oxida- 
tion of  C^*-labeled  substrates  by  ventricle  strips  were  then  studied  in  cham- 
bers in  which  the  respiration  and  contractile  activity  could  be  determined 
simultaneously  (Rice  and  Berman,  1961).  Malonate  at  5.6  mM  under  con- 
ditions in  which  a  positive  inotropic  effect  is  observed  has  very  little  effect 
on  the  utilization  of  glucose- 1-C'^,  glucose-6-C^*,  and  pyruvate-2-C^*,  slight 
inhibition  of  glucose  oxidation  being  noted  although  this  is  possibly  not 
significant.  These  results  indicate  that  the  stimulatory  action  is  not  related 
to  (1)  acceleration  of  glucose  metabolism,  (2)  inhibition  of  the  cycle,  or  (3) 
stimulation  of  the  pentose-phosphate  pathway.  It  was  found  that  C^^Og  is 
produced  from  malonate-2-C^*  in  ventricle  strips,  and  possibly  part  of  the 
positive  inotropic  action  in  substrate-depleted  strips  is  related  to  the  oxida- 
tion of  malonate  via  formation  of  acetyl-CoA  and  its  incorporation  into 
the  cycle.  However,  the  explanation  for  the  greater  effect  of  malonate  in 
the  presence  of  glucose  and  the  inhibition  of  its  action  by  iodoacetate  is 
not  immediately  evident.  It  may  be  noted  that  other  metabolic  inhibitors, 
such  as  fluoride,  arsenate,  fluoroacetate,  and  dehydroacetate,  can  exert  po- 
sitive inotropic  actions  under  the  appropriate  conditions,  so  this  paradoxical 
effect  of  malonate  is  not  unique. 

Wenzel  and  Siegel  (1962)  determined  the  dose-response  curves  for  the 
positive  inotropic  effects  of  malonate  and  ouabain  on  the  rat  ventricle  strip, 
and  then  constructed  an  isobologram,  plotting  the  malonate  concentration 
against  the  ouabain  concentration  for  a  chosen  contractile  stimulation. 
Since  the  isobol  sags,  i.e.,  is  concave  upwards,  they  claimed  it  is  clear  that 
potentiation  occurs  and  that  this  indicates  the  sites  of  action  of  malonate 
and  ouabain  are  different.  There  is  some  doubt  that  a  moderately  sagging 
isobol  can  be  interpreted  as  potentiation,  inasmuch  as  pure  summation 
often  elicits  such  a  curve  (see  Figs.  1-10-7,  8). 


EFFECTS   OF   MALOXATE   IN   THE   WHOLE   ANIMAL  217 

The  response  of  the  heart  to  neurohormones  is  not  altered  by  malonate. 
The  positive  chronotropic  effect  of  epinephrine  on  the  frog  heart  is  not 
changed  by  0.1  roM  malonate  (Nickerson  and  Nomaguchi,  1950),  which  is 
not  surprising  considering  the  concentration.  A  brief  report  by  Ellis  and 
Anderson  (1951  a)  stated  that  malonate  does  not  affect  the  stimulation  by 
epinephrine  except  after  prolonged  treatment  when  the  frog  heart  is  de- 
pressed. The  malonate  concentration  was  not  given.  It  would  seem  likely 
that  any  severe  metabolic  disturbance  producing  marked  cardiac  depression 
would  prevent,  or  at  least  reduce,  any  type  of  stimulation,  since  the  ad- 
ditional functional  activity  would  demand  more  energy,  so  that  an  an- 
tagonism of  epinephrine  by  malonate  is  not  of  much  significance  unless  it 
occurs  when  the  heart  is  not  too  much  depressed.  Malonate  has  no  demon- 
strable effect  on  the  response  of  the  heart  to  acetylcholine,  with  respect  to 
reduction  of  either  rate  or  contractility  (Webb,  1950  b),  whereas  fluoroace- 
tate  alters  the  response  markedly,  indicating  again  that  the  cycle  is  of  im- 
portance in  these  myocardial  functions  but  that  malonate  does  not  reach 
sufficiently  high  intracellular  concentrations. 


EFFECTS   OF   MALONATE    IN   THE   WHOLE   ANIMAL 

A  summary  of  the  miscellaneous  results  relating  to  toxic  and  lethal  ef- 
fects of  malonate  is  given  in  Table  1-28.  One  may  conclude  that  in  mammals 
injected  doses  of  1.5  2.5  g/kg  (10  17  millimoles/kg)  of  sodium  malonate  are 
generally  lethal.  Such  doses,  especially  when  given  intravenously,  probably 
produce  plasma  levels  in  excess  of  10  inM  malonate  at  peak  concentra- 
tions. A  dose  of  12  millimoles/kg  subcutaneously  in  rats  leads  to  a  plasma 
concentration  of  4.5  mill  at  30  min,  and  another  similar  dose  raises  the 
plasma  level  to  around  8  mM  (Busch  and  Potter,  1952  a).  Intravenous 
injection  would  give  higher  peak  levels.  When  compared  in  the  same  ex- 
periments, the  acid  is  more  toxic  than  the  sodium  salt.  It  is  difficult  to 
know  if  this  is  due  to  a  nonspecific  acid  effect  or  to  better  penetration  into 
the  tissues.  It  would  appear  that  malonate  is  more  toxic  to  mice  at  38°  than 
at  30°  environmental  temperature  (Gruber  et  al.,  1949). 

The  sequence  of  symptoms  resulting  from  the  injection  of  malonate  into 
rats  and  mice  was  described  by  Gruber  et  al.  (1949)  as:  champing,  air  hunger, 
maintenance  of  the  head  in  a  dorsally  flexed  position,  and  clonic  convulsions. 
Busch  and  Potter  (1952  a)  found  dyspnea  and  convulsions  in  rats  follow- 
ing injection  of  toxic  doses.  The  cause  of  death  has  been  attributed  to  var- 
ious actions.  Forssman  (1941)  and  Forsmann  and  Lindsten  (1946)  believed 
that  death  is  due  to  cardiac  failure  (the  cardiovascular  effects  observed  were 
discussed  in  the  previous  section).  Handler  (1945)  also  favored  a  cardiac 
mechanism  for  death  and  found  the  succinate  oxidase  to  be  inhibited  50- 
75%  in  homogenates  prepared  from  poisoned  animals.  He  also  noted  that 


218 


1.    MALONATE 


"So 


ft 

CO 


lO    '« 


c3 

:;^ 

JU 

C  J 

r«^ 

05 

S3i 

^—^ 

(T) 

GO 

o 

H.^ 

CD 

TO 

Tt* 

05 

-1^ 

^ 

a 

^ 

T3 

S 

c 

eS 

T3 

t*. 

l>» 

> 

>> 

<» 

3    O 


^    t- 


(»      4)      o3 

W  W  P 


o 

p< 

(_;       tn       ^       ^ 

m  p^  Q  o 


Ci 

Oi 

C5 

xs 

"^ 

'IH 

•* 

-* 

ctf 

(-1 

02 

C5 

c: 

f/; 

ii 

" — ' 

" — ' 

^^ 

73 

o 

t**" 

^ 

^ 

^ 

w 

« 

« 

cS 

t3 
C 

«> 

01 

CS 

u 

b 

u 

01 

0) 

0) 

^ 

-O 

-O 

^ 

o 

D 

3 

3 

m 

;h 

t4 

3 

O 

O 

o 

S 

M 

^  ^  rW  "" 

a  a  ^  . 

2  ^  ^  §  "e 

cd  c^  c^ 


CO    ^    05 

Ift      /-T%      ^H 


f     "O       03       03       CD 
C^        ^        CO        M        «1 


f=^  O  ^ 


cs         -r?   c^ 


'T3   -C 


oj    o    d 


hJ  Z  H  h^  ^  S 


_     X 


o     o     o     cS 


Q  Q  Q  Q 

hj     ^-5     H-5     h^ 


o    <u    o    0)    c 
H  h^l  12;  hJ  M 


(N       |ci-*->*OOiO(M»OOOOOOCO«DTj< 


OG0OC0-Ht-rt(M(N-H 


O-hO-hO^OOO 


-t3     +3     ^o     -u     -u 


09        03        CO 


i  c^     cd  ^     ^     cd     cd     c6 

^   vj   vj   vj  _^   vj  _^   CO   CO  _^   CO   CO   CO   CO   CO 

'oo3cScS*ocS'oc3cS'Sc8(:3«Sc8cS 


'i^  O^ 


P 


Ph 


o  o 


>   CM 


PL|  O  O  Q  > 

h-i    CO    CO    W    M 


Ci2 


IS  ,o 


43    T) 

03      ^ 


bO 


to  C 
O  — ' 
P   ^o 


03 

>>  "a 
a 


W)     bO 


O 

O 

O 

03 

cS 

.2 

cS 
Xi 

o 

C 

o 

T3 

<D 

03      cS 


P    P 


EFFECTS    OF   MALONATE    IN   THE    WHOLE    ANIMAL  219 

the  heart  is  dilated  at  death  and  that  there  is  an  accumulation  of  ascitic 
and  pleural  fluid.  Forssman  and  Lindsten,  from  their  failure  to  obtain  appre- 
ciable direct  action  on  the  isolated  heart,  suggested  an  indirect  mechanism, 
possibly  mediated  through  the  effect  of  malonate  on  liver  metabolism.  The 
dyspnea  and  various  central  nervous  system  effects  may  result  from  a  direct 
action  of  malonate  but  could  also  arise  from  the  ionic  and  acid-base  im- 
balances produced.  Handler  (1945)  showed  that  malonate  causes  a  marked 
fall  in  CO2  capacity  in  the  blood  from  53  to  6  vol%.  Although  no  study  of 
the  alterations  in  the  plasma  electrolytes  undoubtedly  produced  by  mal- 
onate has  been  made,  Wick  et  at.  (1956)  noted  that  the  ability  of  the  blood 
to  coagulate  is  reduced,  and  attributed  many  of  the  actions  of  malonate  to 
the  chelation  of  divalent  cations.  Other  changes  in  the  blood  composition 
may  also  be  responsible  for  some  of  the  toxic  effects.  Handler  (1945)  re- 
ported that  1.6  g/kg  malonate  given  subcutaneously  to  rabbits  increases 
the  blood  glucose  368%,  blood  lactate  545%,  blood  pyruvate  163%,  serum 
inorganic  phosphate  262%,  and  serum  organic  phosphates  155%.  (The  keto- 
nemia  produced  by  malonate  was  discussed  previously).  There  is  thus  much 
opportunity  for  secondary  mechanisms  to  play  a  role  in  the  toxicity  of 
malonate.  At  the  present  state  of  our  knowledge  it  is  even  difficult  to  evaluate 
the  importance  of  succinate  oxidase  inhibition  in  these  effects. 

The  kidneys  have  the  highest  concentration  of  malonate  after  administra- 
tion and  therefore  the  renal  effects  and  nephrotoxicity  have  been  investigat- 
ed. Early  arguments  about  the  renal  toxicity  of  glutarate  were  published 
between  1907  and  1912.  Rose  (1924)  "reinvestigated  this  and  found  that 
glutarate  is  a  nephrotoxic  substance  in  rabbits,  as  indicated  by  the  increases 
in  nonprotein  nitrogen,  urea,  and  creatinine,  and  the  almost  complete  disap- 
pearance of  renal  function  as  measured  by  the  phenolsulfonphthalein  test. 
A  single  experiment  with  malonate  was  reported.  No  renal  damage  was 
observed  after  2  g  given  on  successive  days  and  3  g  2  days  later,  nor  was 
there  a  change  in  the  rate  of  dye  excretion.  Corley  and  Rose  (1929)  reported 
that  methylmalonate  and  ethylmalonate  are  slightly  toxic  to  the  kidneys 
at  doses  of  about  1  g/kg  in  rabbits,  there  being  a  definite  increase  in  the 
nonprotein  nitrogen  and  some  reduction  in  dye  secretion,  although  both 
effects  are  transitory.  Extensive  renal  damage  was  observed  by  Becker  and 
Rieken  (1954)  following  the  intraarterial  injection  of  20  mg  potassium  mal- 
onate (Fig.  1-19  a).  The  vessel  walls  become  edematous,  potocytosis  is  evi- 
dent, and  many  perinuclear  vacuoles  appear  in  the  loops  of  Henle.  However, 
it  requires  much  higher  doses  to  depress  the  respiration  of  kidney  slices 
prepared  from  injected  animals,  80  mg  potassium  malonate  giving  no  effect 
and  150  mg  inhibiting  30.5%.  Similar  histological  changes  occur  after  in- 
cubating kidney  tissue  in  vitro  with  110  mil/  potassium  malonate  for  20  min 
(Fig.  1-19  b),  vacuolization  being  intense.  Tinacci  (1953)  found  not  only 
kidney  damage  but  widespread  degenerative  changes  2-8  days  after  sub- 


220 


1.    MALONATE 


Fig.  1-19.  Renal  damage  resulting  from  (a)  intraarterial  injection  of  20  mg  potas- 
sium malonate  into  rabbits,  and  (b)  incubation  of  slices  with  110  mif  malonate  for 
20  min.   (From  Becker  and  Rieken,   1954.) 


EFFECTS  OF  MALONATE  ON  BACTERIAL  INFECTIONS        221 

cutaneous  injections  of  0.4-0.8  g/kg  of  sodium  malonate,  almost  all  organs 
being  involved.  Malonate  is  diuretic  when  subcutaneously  injected  into 
hydrated  rats  at  a  dose  of  11.5  millimoles/kg,  this  dose  being  sufficient  to 
inhibit  kidney  succinate  oxidase  (Fawaz  and  Fawaz,  1954).  However,  the 
effect  is  probably  osmotic  rather  than  due  to  enzyme  inhibition  by  the  mal- 
onate, since  KNO3  at  the  same  dosage  induces  the  diuresis,  and  also  be- 
cause diethylmalonate,  which  inhibits  kidney  succinate  oxidase,  has  no 
diuretic  activity.  Summarizing  these  results,  one  may  conclude  that  renal 
damage  may  occur  from  high  doses  of  malonate,  but  that  minimal  changes 
in  the  kidney  occur  after  the  usual  toxic  doses.  It  seems  unlikely  that  the 
renal  action  is  of  major  importance  in  poisoning  or  death  from  malonate. 

EFFECTS   OF   MALONATE   ON    BACTERIAL   INFECTIONS 

The  influence  of  enzyme  inhibitors  on  the  course  of  bacterial  infections 
is  of  great  interest  because  the  results  have  bearing  on  the  fundamental 
question  of  the  metabolic  basis  of  the  resistance  to  infection.  The  effects 
of  malonate  on  Salmonella  typhimurium  infection  in  mice  have  been  studied 
by  Berry  and  co-workers  at  Bryn  Mawr  in  a  series  of  excellent  investiga- 
tions. Mice  injected  intraperitoneally  with  a  Salmonella  suspension  show 
evidence  of  the  infection  on  the  third  day  and  most  succumb  by  the  sixth 
day.  If  mice  are  given  intraperitoneal  injections  of  20  mg  sodium  malonate 
in  saline  every  hour  for  8  hr,  they  die  much  more  rapidly  from  the  infection 
than  animals  given  only  saline  injections  (Berry  and  Mitchell,  1953  a). 
The  striking  effects  is  illustrated  in  Fig.  1-20.  Noninfected  mice  show  no 
effects  of  the  malonate.  Thus  sublethal  doses  of  malonate  are  able  either 
to  accelerate  bacterial  proliferation  or  to  decrease  the  resistance  of  the 
host  markedly.  These  are  the  basic  observations  and  the  later  work  attempts 
to  elucidate  the  mechanisms  by  which  these  effects  are  brought  about. 

The  reduction  of  the  survival  time  in  mice  infected  with  Salmonella  is 
not  unique.  Similar  effects  of  malonate  on  infections  with  Proteus  morganii, 
Staphylococcus  aureus,  Streptococcus  pyogenes  (Berry  and  Mitchell,  1954), 
Diplococcus  pneumoniae,  and  Klebsiella  pneumoniae  (Berry  et  al.,  1954  a) 
have  been  observed.  Furthermore,  reduced  survival  times  have  been  found 
in  Salmonella  infected  rats,  guinea  pigs,  and  chickens  (Berry  and  Beuzeville, 
1960).  Finally,  the  phenomenon  has  been  seen  with  other  inhibitors,  such 
as  fluoroacetate  and  arsenite  (Berry  et  al.,  1954  a,  b).  This,  then,  is  a  gen- 
eral effect  of  certain  types  of  inhibitor,  especially  those  affecting  the  cycle 
in  some  manner,  and  the  problem  is  thus  more  important  because  it  must 
relate  to  some  fundamental  metabolic  relationship  between  host  and  bacteria. 

We  shall  now  examine  in  greater  detail  some  of  the  characteristics  of 
the  malonate  effect.  Malonate  not  only  can  reduce  survival  time,  but  in 
some  instances  can  change  a  nonlethal  infection  into  a  lethal  one,  this  being 


222 


1.    MALONATE 


observed  with  Diplococcus,  Staphylococcus,  and  Proteus.  That  is,  an  infec- 
tion which  does  not  kill  plus  a  dose  of  malonate  which  is  non-toxic  may 
cause  the  death  of  all  the  animals  within  several  hours  (Berry  et  al.,  1954  a). 
This  is  a  true  case  of  synergism.  It  has  been  found  that  Salmonella  bactere- 
mia in  mice  is  much  greater  in  malonate-treated  animals  than  in  the  con- 
trols (Berry  and  Mitchell,  1953  b,  1954).  The  bacteremia  in  the  controls 
reaches  a  low  peak  value  soon  after  inoculation  and  then  falls  off,  whereas 
in  the  presence  of  malonate  it  continues  to  progress.  At  9  hr,  the  blood  of 


No.    10 
Dead 


/    Malonate 

/ 

j 

/Saline 

/  controls 

. 

y 

24       30       36        42       48 

Time    after     injection  (  Hours) 


Fig.  1-20  Effect  of  malonate  given  in  8  hourly  injections  of  20  mg 
sodium  malonate  on  the  mortality  of  mice  intraperitoneally  in- 
jected with  suspensions  of  Salmonella  typhimurium .  (From  Berry 
and  Mitchell,   1953  a.) 


the  controls  contains  10,000  to  20,000  bacteria/ml  whereas  the  blood  of 
malonate-treated  animals  has  a  count  of  around  3,000,000  bactcria/ml. 
Thus  the  bacteremia  is  over  100-fold  as  severe  in  the  treated  animals  as  in 
the  controls.  This  bacteremia  is  a  reflection  of  the  situation  throughout 
the  body.  The  total  number  of  bacteria  in  the  body  6.5  hr  after  the  injec- 
tions is  1.65  X  10^  in  the  controls  and  32.1  X  10^  in  the  malonate-treated 
mice  (Berry,  1955).  The  ratio  of  counts  in  the  treated  and  control  series  is 
20  for  the  whole  body  and  19  for  the  blood  at  this  time.  Now,  the  interesting 
thing  is  that,  at  the  time  of  death,  the  number  of  bacteria  in  the  body  is 
the  same  in  both  the  controls  and  malonate-treated  mice.  This  clearly  shows 
that  malonate  does  not  alter  the  susceptibility  of  the  mice  to  the  bacteria, 
but  reduces  the  time  required  for  the  bacteria  to  multiply  to  that  number 
necessary  to  kill. 


EFFECTS    OF    MALONATE    ON    BACTERIAL    INFECTIONS  223 

The  rapid  proliferation  of  bacteria  could  be  due  to  a  weakening  of  the 
body  defenses  for  disposing  of  the  bacteria.  This  does  not  seem  to  be  the 
case,  inasmuch  as  malonate  has  no  effect  on  the  uptake  of  thorotrast  by 
the  reticuloendothelial  system  (Berry,  1955)  nor  does  it  depress  phagocytosis 
except  at  very  high  concentrations  (Berry  and  Derbyshire,  1956).  Instead  it 
would  appear  that  malonate  disturbs  metabolism  in  such  a  way  that  it 
creates  a  more  favorable  environment  in  the  host  for  bacterial  growth. 
Malonate  itself  may  be  metabolized  slightly  by  Salmonella,  but  not  to  the 
extent  required  to  explain  the  explosive  proliferation  (Berry  and  Beuzeville, 
1960).  Growth  medium  was  prepared  from  the  eviscerated  carcasses  of  con- 
trol and  malonate-treated  animals  and  it  was  found  that  the  bacteria  grow 
more  rapidly  in  the  latter  (Berry,  1955).  It  has  also  been  shown  that  Sal- 
monella grows  more  rapidly  in  the  peritoneal  fluid  of  malonate-treated  mice 
than  in  the  controls  (Berry  and  Beuzeville,  1960).  Citrate  is  known  to  ac- 
cumulate following  the  administration  of  malonate.  This  was  confirmed  in 
mice  given  the  doses  of  malonate  capable  of  reducing  survival  times  of 
infected  animals  (Berry  et  al.,  1954  b).  Both  malonate  and  endotoxin  from 
Salmonella  increase  citrate  levels  in  most  tissues,  and  together  the  increases 
are  often  greater  than  with  either  alone  (see  accompanying  tabulation).  It 


Treatment 


Citrate  (//g/g)  in 


None 

42 

94 

76 

HI 

Malonate 

43 

245 

225 

120 

Endotoxin 

50 

273 

190 

187 

Malonate 

+  endotoxin 

70 

173 

173 

407 

Blood           Spleen         Kidney        Heart        Duodenum  Liver 

132  109 

115  170 

170  443 

543  240 


is  thus  possible  that  a  summation  of  effects  on  the  cycle  could  be  partially 
responsible  for  the  increased  mortality.  However,  Salmonella  infection  does 
not  increase  citrate  levels  (Berry  and  Beuzeville,  1960).  Could  the  increased 
citrate  be  favorable  to  the  growth  of  the  bacteria?  It  is  unlikely  that  this 
is  a  major  factor  because  malonate  is  the  most  potent  inhibitor  for  reducing 
survival  times  and  yet  both  arsenite  and  fluoroacetate  cause  greater  accu- 
mulations of  citrate.  The  primary  cause  of  the  augmented  bacterial  proli- 
feration has  not  been  found  but  the  range  of  possible  mechanisms  has  been 
narrowed.  Since  there  are  many  other  possible  substrates  for  Salmonella 
that  accumulate  during  malonate  inhibition,  it  will  be  necessary  to  examine 
these  in  mice  and  their  effects  on  the  growth  of  Salmonella. 

Some  work  on  this  problem  in  other  laboratories  may  be  briefly  mention- 
ed. Malonate  reduces  the  antibacterial  activity  of  guinea  pig  blood  toward 


224  1.    MALONATE 

Salmonella  enteritidis  but  this  is  not  due  to  a  decrease  in  the  number  of 
leucocytes  (Yamauchi,  1956).  The  survival  times  of  chicks  infected  with 
Salmonella  pullorum  are  reduced  by  500-800  mg/kg  of  malonate  injected 
simultaneously  or  shortly  after  the  bacterial  inoculation  (Gilfillan  et  al., 
1956).  On  the  other  hand,  the  diethyl  ester  of  malonate  increases  the  sur- 
vival time  of  mice  infected  with  Mycobacterium,  tvbercvlosis,  when  admin- 
istered daily  for  2  weeks  at  oral  doses  of  250-500  mg/kg  (Davies  et  al., 
1956).  Compounds  of  this  type  are  thus  considered  worthy  of  study  as 
chemotherapeutic  agents  in  tuberculous  infections. 


METABOLISM    OF    MALONATE 

Malonate  occurs  normally  in  many  types  of  organism  and  occasionally 
at  concentrations  possibly  inhibitory  to  succinate  dehydrogenase.  Many 
organisms  are  capable  of  metabolizing  malonate  by  various  pathways  and 
some  are  able  to  utilize  it  for  growth  or  cell  functions.  In  some  cases,  in- 
deed, it  is  difficult  to  demonstrate  the  inhibitory  action  of  malonate  in  the 
presence  of  its  own  oxidation.  The  metabolism  of  malonate  must  always  be 
considered  in  studies  of  the  effects  of  malonate  on  any  type  of  cellular  ac- 
tivity. It  is  often  impossible  to  detect  and  correct  for  the  metabolism  of 
malonate  without  using  labeled  malonate.  It  is  possible  that  many  studies 
of  the  inhibition  of  respiration  or  cycle  activity  by  malonate  have  been 
complicated  by  the  oxidation  of  the  malonate. 

Occurrence  of  Malonate 

Malonate  has  been  isolated  or  demonstrated  chromatographically  from  a 
number  of  microorganisms,  plants,  and  animals,  and  it  is  likely,  consider- 
ing the  recent  demonstration  of  its  role  and  the  role  of  malonate  derivatives 
in  fatty  acid  metabolism  that  its  occurrence  is  widespread.  The  accom- 
panying incomplete  tabulation  will  serve  to  illustrate  this.  Malonate  has 
also  been  found  in  winter  wheat,  barley,  oats,  alfalfa,  kidney  bean  leaves, 
clover,  pea  leaves,  vetch  leaves  (Soldatenkov  and  Mazurova,  1957),  sake 
(Kawano  and  Kawabata,  1953),  and  several  products  prepared  from  plants. 
Although  no  thorough  studies  of  animal  tissues  have  befen  made,  it  is 
evident  that  malonate  must  occur  in  rat,  dog,  and  human  tissues  to  some 
extent  if  it  is  found  in  the  urine.  Although  the  name  of  malonate  comes 
ultimately  from  the  Latin  mains,  it  has  never  been  identified  in  apples 
or  other  fruit. 

Methylmalonate  has  been  found  in  Propionibacterium  (Stjernholm  and 
Wood,  1961;  Wood  and  Stjernholm,  1961),  pigeon  liver  (Bressler  and  Wakil, 
1961),  pig  heart  (Flavin  et  al.,  1955),  mouse  adipose  tissue  (Feller  and  Feist, 
1957),  and  rat  and  human  urine  (Boyland  and  Levi,  1936;  Barness  et  al., 


METABOLISM    OF    MALONATE 


225 


Source  of  malonate 


Reference 


Achromobader  guttatum 
Nocardia  corallina 
Penicillium  funiculosum 
Aspergillus  niger 


Phaseolus  vulgaris  (bush  bean) 


Phaseolus  coccineus  (runner  bean) 
Wheat  plants 

Bunias  orientalis  (Cruciferae) 
Tobacco  leaves 


Lucerne  (green  alfalfa) 

Hevea  brasiliensis 

Helianthus  annus 

Umbelliferae  {Anthriscus  and  Apium) 

Leguminosae  (18  species) 

Rat  urine 

Dog  urine 
Human  urine 


Sguros  and  Hartsell  (1952a) 

i.ara  (1952) 

Igarasi  (1939) 

Challenger  et  al.  (1927)  Walker 

et    al.     (1927),     Subramanian 

et  al.  (1929) 
Young    and    Shannon    (1959), 

Rhoads  and  Wallace  (1957), 

Huffaker  and  Wallace  (1961) 
Bentley  (1952) 

Nelson  and  Hasselbring  (1931) 
Jermstad  and  Jensen  (1951) 
Wada  and  Kobashi  (1953),  Bel- 

lin  and  Smeby  (1958),  Vickery 

and    Palmer    (1956  b,    1957), 

Vickery  (1959) 
Turner  and  Hartman  (1925) 
Fournier  et  al.  (1961) 
Bentley  (1952) 
Bentley  (1952) 
Bentley  (1952) 
Stalder    (1958),    Thomas    and 

Stalder  (1958) 
Thomas  and  Kalbe  (1953) 
Stalder  (1958) 


1957;  Stalder,  1958;  Thomas  and  Stalder,  1958).  Ethylmalonate  has  been 
found  in  rat  and  human  urine  (Stalder,  1959).  Hydroxymalonate  (tartronate) 
occurs  in  Acetohacter  (Stafford,  1956)  and  malonic  semialdehyde  in  Pseudo- 
monas  (Nakamura  and  Bernheim,  1961). 

The  concentrations  of  malonate  in  plant  tissues  are  often  surprisingly 
high.  The  legumes  and  umbellifers  analyzed  by  Bentley  (1952)  contain 
0.5-2  mg/g  fresh  tissue.  These  values  correspond  to  7-30  raM  malonate  if 
distributed  uniformly  throughout  the  tissue  water.  The  stems  of  the  runner 
bean  {Phaseolus  coccineus)  contain  2.1  mg/g  and  the  expressed  juice  of  the 
stem  is  30  milf  in  malonate.  Since  20  raM  malonate  at  pH  4.5  inhibits  the 
respiration  of  these  stems  50%  and  causes  accumulation  of  succinate,  one 
would  expect  the  metabolism  in  these  plants  to  be  constantly  suppressed 
by  the  malonate,  unless  the  malonate  is  in  some  manner  segregated  from 
the  metabolic  systems.  Soldatenkov  and  Mazurova  (1957)  reported  similar 


226  1.    MALONATE 

values  in  legumes  (2-3%  of  the  plant  dry  weight)  and  that  in  kidney-bean 
and  clover  leaves  malonate  represents  45%  of  the  total  di-  and  tricarbox- 
ylates  present.  Bush-bean  {Phaseolus  vulgaris)  leaves  often  contain  as 
much  as  10  mg/g  dried  tissue  and  malonate  is  more  concentrated  than  fu- 
marate  or  succinate,  although  less  than  malate  and  citrate  (Young  and 
Shannon,  1959).  In  man  malonate  is  excreted  in  the  urine  at  an  average 
rate  of  0.0047  mg/kg/day  and  in  the  rat  at  10  times  this  rate  (Stalder,  1958). 
This  amounts  to  only  0.32  mg/day  in  man  (only  two  individuals  were  tested 
so  these  averages  are  not  accurate).  Since  malonate  is  apparently  metabol- 
ized in  mammals,  the  tissue  concentration  or  rate  of  excretion  will  reflect  a 
balance  between  formation  and  destruction.  In  other  words,  these  excretion 
values  do  not  necessarily  represent  the  rates  of  malonate  formation. 

Relatively  little  is  known  about  the  pathways  for  the  formation  of  mal- 
onate, but  the  miscellaneous  observations  make  it  likely  that  different  reac- 
tions are  involved  in  various  organisms.  Malonate  can  arise  from  many 
different  substrates  but  in  most  cases  the  pathways  are  complex  and  the 
immediate  precursors  are  not  known.  Malonate  can  be  formed  from  pyri- 
midines  and  barbiturates  in  the  mycobacteria  (Hayaishi  and  Kornberg, 
1952),  from  pyrimidines  in  Nocardia  (Lara,  1952),  from  acetate  in  Hevea 
hrasiliensis  (Fournier  et  al.,  1961)  and  avocado  (Mudd  and  Stumpf,  1961), 
from  citrate  in  Aspergillus  niger  (Challenger  et  al.,  1927),  from  succinate 
in  Aspergillus  niger  (Subramanian  et  al.,  1929),  from  asparagine  in  rats 
(Thomas  and  Stalder,  1959),  from  oxalacetate  in  pig  heart  extracts  catalyzed 
by  metmyoglobin  and  Mn++  (Vennesland  and  Evans,  1944;  Vennesland 
et  al.,  1946),  and  from  malonyl-CoA  in  Penicillium  cyclopium,  (Bentley  and 
Keil,  1961).  The  high  concentrations  of  malonate  in  bush-bean  plants  led 
Huffaker  and  Wallace  (1961)  to  study  the  mechanism  of  the  accumulation. 
They  found  that  the  malonate  synthesis  is  related  to  the  dark  COg  fixation 
in  the  roots,  phosphoenolpyruvate  being  carboxylated  to  oxalacetate  and 
this  going  to  malonate  with  the  help  of  one  or  more  enzymes.  The  addition 
of  phosphoenolpyruvate  and  Mg++  to  root  homogenate  leads  to  the  for- 
mation of  labeled  malonate  from  0^*02-  It  was  also  found  that  any  other 
reactions  utilizing  oxalacetate  decrease  the  yield  of  malonate.  It  is  very 
interesting  that  frogs  accumulate  malonate-C^*  from  C^Oa,  along  with  other 
dicarboxylates  (Cohen,  1963).  In  normal  strains  the  malonate  accounts  for 
only  0.3-0.5%  of  the  total  incorporation  but  in  hybrids  {R.  pipiens  X  R. 
sylvatica)  the  value  is  6-23%.  This  increased  accumulation  in  the  hybrids 
was  attributed  to  some  defect  in  the  metabolism  of  malonate. 

Methylmalonate  can  be  formed  from  propionate  (Flavin  et  al.,  1955)  in  a 
variety  of  tissues,  and  in  rat  liver  the  pathway  has  been  shown  to  go  through 
succinate  (Katz  and  Chaikoff,  1955).  The  feeding  of  isobutyrate  and  valine 
to  rats  leads  to  the  formation  of  methylmalonate  (Thomas  and  Stalder, 
1958)  and  the  feeding  of  isoleucine  leads  to  ethylmalonate  (Stalder,  1959). 


METABOLISM  OF  MALONATE  227 

The  first  reaction  appears  to  be  an  oxidative  deamination.  Since  methyl- 
malonyl-CoA  is  a  common  intermediate  in  many  tissues,  methylmalonate 
probably  arises  from  any  substance  forming  the  coenzyme  A  derivative. 
(This  will  be  discussed  in  greater  detail  in  the  following  sections). 

General   Occurrence  and    Nature  of  Malonate   Metabolism 

The  metabolism  of  malonate  by  many  organisms  and  tissues  has  been 
conclusively  demonstrated  by  a  variety  of  techniques.  The  ideal  method  is 
the  determination  of  C^^Og  or  other  labeled  products  formed  from  labeled 
malonate,  but  in  some  instances  good  evidence  is  provided  by  studies  of 
oxygen  uptake,  especially  when  the  endogenous  respiration  is  very  small, 
or  by  growth  in  media  containing  only  malonate  as  a  utilizable  substrate. 
In  other  cases,  the  evidence  is  more  indirect.  For  example,  a  stimulation  of 
growth  rate  or  a  rise  in  respiration  in  the  presence  of  other  substrates  may 
suggest  the  utilization  of  malonate  but  does  not  prove  it.  In  the  tabulation 
on  page  228  of  organisms  in  which  malonate  metabolism  has  been  claimed 
to  occur,  those  that  are  probable  and  based  on  indirect  evidence  are  des- 
ignated by  (P).  It  may  be  noted  in  addition  that  Shannon  et  al.  (1959)  found 
malonate  to  be  metabolized  by  the  excised  leaves  of  15  different  common 
plants  (such  as  fig,  peach,  eucalyptus,  azalea,  and  lantana)  and  30  other 
plants  representing  27  families,  from  which  it  must  be  concluded  that  plants 
are  generally  capable  of  utilizing  malonate. 

The  bacterial  oxidation  of  malonate  occasionally  shows  a  lag  period,  first 
observed  by  Lineweaver  (1933).  The  rate  of  oxidation  by  Azotobacter  is 
very  low  for  2-3  hr,  rises  to  a  maximum  around  6-8  hr,  and  falls  off  by 
10  hr.  The  malonate  is  eventually  99%  metabolized  with  an  R.Q.  of  1.6. 
The  theoretical  R.Q.  for  complete  oxidation: 

CHaiCOOH),  +  2  O2    -»    3  CO2  +  2  H,0 

is  1.5.  Lineweaver  postulated  that  two  separate  reactions  are  involved:  the 
decarboxylation  of  malonate  to  acetate,  and  the  oxidation  of  the  acetate. 
He  attributed  the  lag  phase  to  the  slow  decarboxylation,  which  was  sup- 
ported by  the  progressive  decrease  in  the  R.Q.  with  time.  This  interpreta- 
tion was  criticized  by  Karlsson  (1950)  because  Lineweaver  had  not  used 
cells  adapted  to  malonate.  Malonate-grown  Azotobacter  does  not  decarbox- 
ylate  malonate  appreciably  under  anaerobic  conditions,  so  Karlsson  con- 
cluded that  oxygen  is  required  for  the  initial  attack  on  malonate,  either 
because  malonate  must  be  oxidized  before  decarboxylation  or  because 
oxygen  may  be  required  for  some  activation  of  malonate  (for  example,  by 
a  phosphorylative  mechanism).  A  lag  phase  was  also  demonstrated  for 
Aerobacter  by  Barron  and  Ghiretti  (1953),  the  maximal  oxidative  rate  oc- 
curring around  2-3  hr  after  malonate  addition.  Only  40%  of  the  malonate 


228 


1.  MALONATE 


Organisms  and  tissues 
metabolizing  malonate 


Reference 


Pseudomonas  aeruginosa 
Pseudomonas  fluorescens 


Escherichia  coli  (P) 

Aerobacter  aerogenes 
Azotobacter  agilis 
Salmonella  typhimurium  (P) 
Mycobacterium  tuberculosis 


Mycobacterium  phlei 
Aspergillus  niger 

Aspergillus  (6  species)  (P) 

Penicillium  cyclopium  (P) 

Penicillium  (3  species)  (P) 

Streptomyces  olivaceus 

Pullularia  pullulans  (P) 

Avena  coleoptile  (P) 

Chlorella  pyrenoidosa  (P) 

Pollen  of  Camellia,  Thea,  and  Lilium 

Tobacco  leaves 

Bush  bean  leaves 

Peanut  mitochondria 

Euglena  gracilis 

Hymenolepis  diminuta  (cestode)  (P) 

Locusta  migratoria  fat  body 

Chickens  (P) 

Pigeon  liver 

Mice 

Rats 


Rat  liver 

Rat  ventricle 

Rabbits 

Dogs 

Dog  heart  and  muscle 

Human  placenta 

Human  prostate 


Gray  (1952) 

Hayaishi  (1953,  1954,  1955  a),  Wolfe 

and  Rittenberg  (1954),  Wolfe  et  al. 

(1954  a,  b,  1955) 
Grey  (1924),  Quastel  and  Whetham 

(1925),  Cook  (1930) 
Barron  and  Ghiretti  (1953) 
Lineweaver  (1933),  Karlsson  (1950) 
Berry  and  Beuzeville  (1960) 
Hayaishi     and     Kornberg     (1952), 

Bernheim  et  al.  (1953),  Horio  and 

Okunuki    (1954),    Kusunose   et   al. 

(1960) 
Miiller  et  al.  (1960) 
Walker  et  al.  (1927),  Challenger  et  al. 

(1927) 
Berk  et  al.  (1957) 
Bentley  and  Keil  (1961) 
Berk  et  al,  (1957) 
Maitra  and  Roy  (1961) 
Clark  and  Wallace  (1958) 
Albaum  and  Eichel  (1943) 
Eny  (1951) 
Okunuki  (1939) 
Vickery  (1959),  Vickery  and  Palmer 

(1957) 
Young  and  Shannon  (1959) 
Giovanelli  and  Stumpf  (1957) 
Danforth  (1953) 
Read  (1956) 
Tietz  (1961) 

Clementi  (1929),  Pupilh  (1930) 
Menon  et  al.  (1960) 
Lifson  and  Stolen  (1950) 
Lee  and   Lifson   (1951),   Busch  and 

Potter  (1952a),  Nakada  etal.  (1957), 

Thomas  and  Stalder  (1959) 
Menon  el  at.  (1960) 
Rice  and  Berman  (1961) 
Wick  et  al  (1956) 
Pohl  (1896) 
Menon  et  al.  (1960) 
Hosoya  and  Kawada  (1958),  Hosoya 

et  al.  (1960) 
Andrews  and  Taylor  (1955) 


METABOLISM   OF   MALONATE 


229 


is  oxidized  but  the  R.Q.  is  1.47,  close  to  that  for  complete  oxidation  to 
CO2  and  water.  Again,  no  decarboxylation  is  observed  in  nitrogen.  Horio 
and  Okunuki  (1954)  reported  a  30-60  min  lag  period  for  Mycobacterium  and 
also  found  that  decarboxylation  presumably  precedes  oxidation,  because 
CO2  formation  always  is  ahead  of  Oo  uptake,  and  acetate  can  be  demon- 
strated in  the  culture  after  2  hr.  The  explanation  of  this  behavior  is  found 
in  the  work  of  Gray  (1952)  on  Pseudomonas.  Unadapted  cells  show  a  lag 
period  of  2  hr  whereas  cells  cultured  for  70  hr  in  22  mM  malonate  are  able 
to  oxidize  malonate  immediately  (Fig.  1-21).  It  was  postulated  that  mal- 


Time  (  hours) 


Fig.  1-21.   Oxidation  of  malonate  by  normal   and 
adapted  Pseudomonas  aeruginosa.  (From  Gray,  1952.) 


onate  decarboxylase  is  an  adaptive  enzyme,  and  it  was  pointed  out  that 
the  resistance  of  certain  microorganisms  to  malonate  may  be  due  to  such 
an  enzyme  as  well  as  to  permeability  barriers.  Horio  and  Okunuki  showed 
that  streptomycin  does  not  directly  inhibit  the  decarboxylation  of  mal- 
onate or  the  oxidation  of  acetate,  but  inhibits  malonate  oxidation  in  un- 
adapted cells,  probably  by  preventing  the  synthesis  of  the  decarboxylase. 
The  oxygen  requirement  may  also  relate  to  the  adaptive  enzyme  syn- 
thesis. 


230 


1.    MALONATE 


Pathways  of  Malonate   Metabolism   in   Microorganisms 

An  analysis  of  the  pathways  of  malonate  oxidation  was  made  simultane- 
ously in  the  laboratories  of  Hayaishi  and  Rittenberg  between  1953  and  1955. 
The  work  was  done  on  Pseudomonas  fluorescens,  a  soil  isolate  capable  of 
utilizing  malonate  as  the  sole  carbon  source,  and  adapted  to  malonate  by 
culture  in  27-33  nvM  malonate  media.  Hayaishi  (1953)  observed  that  the 
decarboxylation  of  malonate  requires  ATP  and  CoA  and  postulated  that 
malonate  must  first  be  activated,  probably  to  malonyl-CoA.  Using  partially 
purified  extracts  from  Pseudomonas,  it  was  shown  that  malonate  is  quanti- 
tatively converted  to  COg  and  acetate,  no  other  products  being  detectable 
chromatographically.  The  proposed  scheme  (1-7)  may  be  represented  as 
follows  (Hayaishi,  1954,  1955  a): 


Malonate 


Acetate 


Malonate 


(1-7) 


Acetyl  — CoA 


The  cyclic  process  thus  involves  the  transfer  of  CoA  back  and  forth  be- 
tween the  acetyl  and  malonyl  groups.  Wolfe  et  at.  (1954)  also  ruled  out  the 
direct  decarboxylation  to  acetate  by  showing  a  dependence  on  ATP  and 
CoA,  and  in  later  work  (Wolfe  et  al.,  1954  b,  1955;  Wolfe  and  Rittenberg, 
1954)  proposed  the  following  scheme  (1-8)  based  mainly  on  chromatographic 
analyses  of  intermediates  and  products: 


Acetate  +  Malonyl-diCoA- 


Malonate 

ATP  +  CoA 


t 
Malonyl  -  CoA 


Acetyl  -  CoA 


>■ 


(1-8) 


The  principal  difference  between  the  two  schemes  is  the  participation  of 
malonyl-diCoA.  Hayaishi's  results  do  not  exlude  it  but  provide  no  evidence 
for  it,  while  Wolfe  et  al.  claim  to  have  detected  it  chromatographically. 
Although  the  decarboxylation  of  malonate  is  characterized  by  AF  =  —  7 
kcal/mole,  the  activation  energy  is  presumably  so  high  that  the  more  com- 
plex mechanisms  above  are  necessary.  The  malonate  decarboxylase  and 


METABOLISM  OF  MALONATE  231 

CoA-transferase  have  been  partially  purified.  The  acetyl-CoA  formed  from 
malonate  may  enter  the  cycle  directly  or  participate  in  other  reactions,  such 
as  the  formation  of  acetoacetate  if  the  cycle  is  blocked  by  malonate,  or 
transfer  its  CoA  to  malonate,  or  simply  be  hydrolyzed  to  acetate.  It  is  very 
interesting  that  a  lag  period  was  noticed  in  the  oxidation  of  malonate  by 
cell-free  extracts  (Wolfe  et  al.,  1955).  Little  oxygen  uptake  occurs  with 
malonate  until  it  is  all  decarboxylated;  during  the  period  of  the  most  rapid 
CO2  evolution,  the  respiration  is  low.  Yet  acetate  is  activated  and  oxidized 
rapidly.  One  possible  explanation  is  a  block  of  the  cycle  by  malonate  so 
that  acetyl-CoA  cannot  enter  until  most  of  the  malonate  has  been  me- 
tabolized. Another  explanation  involves  a  distribution  of  CoA  in  favor  of 
the  malonyl  derivatives  with  little  acetyl-CoA  present  during  the  active 
decarboxylation  reaction. 

Cryptococcus  terricolus  can  grow  on  malonate  as  the  sole  source  of  carbon 
and  malonate  stimulates  the  endogenous  respiration  and  CO2  formation  after 
a  lag  phase  (Pedersen,  1963).  It  would  appear  that  malonate  is  completely 
oxidized  to  CO2  and  water,  since  the  R.Q.  in  the  presence  of  malonate  is 
1.54.  In  other  microorganisms  where  the  oxidation  is  not  complete,  mal- 
onate may  be  incorporated  into  a  variety  of  substances,  especially  lipids 
(Bu'Lock  et  al.,  1962),  although  the  rate  of  incorporation  is  seldom  very 
rapid. 

Pathways  of  Malonate  Metabolism  In  Plants  and  Animals 

Despite  the  widespread  occurrence  of  malonate  metabolism  in  plant  tis- 
sues, little  is  known  of  the  reactions  involved,  although  perhaps  they  are 
not  significantly  different  from  those  described  for  microorganisms.  Peanut 
mitochondria  supplemented  with  ATP,  CoA,  and  other  factors  are  able  to 
oxidize  malonate  (Giovanelli  and  Stumpf,  1957).  Incubation  with  malo- 
nate-2-C^*  for  2  hr  leads  to  the  appearance  of  labeled  citrate,  malate,  and 
succinate,  indicating  the  sequence 

Malonate    ->■   malonyl-CoA   ->   acetyl-CoA    ->   CO2  +  HjO 

the  last  step  occurring  in  the  cycle.  The  participation  of  malonyl-CoA  in 
the  oxidation  of  propionate  by  peanut  mitochondria  is  suggested  by  the 
tracing  of  the  label  from  propionate- 1-C^*  (Giovanelli  and  Stumpf,  1958). 
The  following  sequence  involving  malonic  semialdehyde  may  be  formulated: 

Propionate    ->   propionyl-CoA    ->■   acrylyl-CoA    ->   |3-hydroxypropionyl-CoA    -> 

^-hydroxypropionate   ->   malonic  semialdehyde    ->   malonyl-CoA    -> 

acetyl-CoA    ->   CO^  +  H2O 

It  is  also  possible  that  CoA  derivatives  are  retained  throughout,  since  a 
malonic  semialdehyde-CoA  dehydrogenase  which  catalyzes  the  formation 


232  1.    MALONATE 

of  malonyl-CoA  has  been  found  in  Clostridium  kluyveri  (Vagelos,  1960). 
In  bush  bean  {Phaseolus  vulgaris)  leaves,  incubation  with  malonate-2-C^* 
leads  to  labeled  citrate,  isocitrate,  and  malate,  indicating  a  pathway  through 
acetyl-CoA  and  the  cycle  (Young  and  Shannon,  1959).  Malonate  is  incor- 
porated in  isolated  spinach  chloroplasts  at  about  half  the  rate  for  acetate, 
much  of  the  label  appearing  in  lipids  (Mudd  and  McManus,  1964). 

Metabolism  of  malonate  was  first  described  in  the  dog  by  Pohl  (1896), 
who  found  that  only  a  fraction  of  the  malonate  administered  to  the  animals 
can  be  recovered  in  the  urine.  This  subject  was  not  taken  up  again  until 
1950  and  since  that  time  much  has  been  learned  of  how  the  body  deals 
with  malonate,  and  of  the  role  of  malonate  and  its  derivatives  in  normal 
metabolism.  It  would  appear  from  the  limited  data  that  about  30%  of 
the  administered  malonate  is  metabolized  (Lifson  and  Stolen,  1950;  Busch 
and  Potter,  1952  a).  However,  the  rate  of  oxidation  is  relatively  slow  and 
in  rabbits  represents  less  than  2%  of  the  total  respiration  (Wick  et  al., 
1956).  The  rate  of  oxidation  may  be  in  part  limited  by  the  transfer  from 
the  extracellular  space  into  the  tissues,  since  this  is  slow. 

Various  mammalian  tissues  can  decarboxylate  malonate  and  utilize  the 
acetate  formed.  This  has  been  investigated  most  completely  in  rat  tissue 
slices  by  Nakada  et  al.  (1957),  who  determined  the  0^*0,  arising  from  mal- 
onate-1-C^^  during  1  hr  incubation  at  pH  7.4,  the  total  concentration  of 
malonate  being  5  xnM  (see  accompanying  tabulation).  Kidney,  liver,  and 


Tissue  Added  C^*  as  Qi^Oa  (%) 

Kidney  27.0 

Liver  18.0 

Heart  15.2 

Diaphragm  6 . 8 

Spleen  1 . 7 

Brain  1 . 5 

Lung  1 . 0 

Testis  0.6 


heart  are  particularly  active,  and  the  tissues  show  a  wide  range  of  decar- 
boxylative  ability,  part  of  which  may  be  due  to  different  rates  of  penetra- 
tion. The  variation  of  malonate  oxidation  with  concentration  is  shown  in 
Fig.  1-22  for  rat  kidney  slices,  and  an  inhibition  of  its  own  metabolism  is 
seen  at  concentrations  above  5  vaM.  The  inhibition  of  acetate- 1-C^*  oxida- 
tion is  shown  for  comparison.  The  relative  rates  of  activation  of  malonate, 
succinate,  and  glutarate  by  several  tissues  were  determined  by  measuring 
the  rates  of  formation  of  hydroxamic  acid  from  hydroxylamine  during  the 


METABOLISM   OF   MALONATE 


233 


incubations  (see  accompanying  tabulation)  (Menon  et  al.,  1960).  The  results 
show  not  only  differences  between  the  tissues,  but  also  that  the  activating 
system  for  malonate  is  different  from  that  for  succinate  and  glutarate. 


Tissue 

Hydroxamic 

acid  formed  (m/<moles/mg/30  min) 

Malonate 

Succinate 

Glutarate 

Dog  heart 

10.3 

7.4 

28.4 

Dog  muscle 

2.6 

8.2 

16.0 

Pigeon  liver 

4.7 

20.5 

31.7 

Rat  liver 

11.5 

62.9 

86.7 

The  metabolic  pathways  for  malonate  in  mammalian  tissues  appear  to 
be  very  similar  to  those  in  bacteria.  The  oxidation  of  malonate  requires 
ATP,  coenzyme  A,  and  Mg++  in  extracts  or  homogenates  of  rat  kidney 
(Nakada  et  al.,  1957),  human  placenta  (Hosoya  and  Kawada,  1958),  and, 
incidentally,  locust  fat  body  (Tietz,  1961).  However,  the  kinases,  decarbox- 


FiG.  1-22.  Effects  of  malonate  concentration  on  the  oxidation 

of  malonate  and  acetate  by  rat  kidney  slices  at  pH  7.4  and 

37°  during  2-hr  incubation.  (From  Nakada  et  al.,  1957.) 


ylases,  and  CoA-transferases  for  these  reactions  have  not  been  studied. 
An  important  transfer  of  coenzyme  A  between  acetoacetyl-CoA  and  mal- 
onate has  been  shown  to  occur  in  pig  heart  extracts  (Beinert  and  Stansly, 
1953).  Acetoacetyl-CoA  is  formed  by  the  condensation  of  two  acetyl-CoA 


234  1.    MALONATE 

molecules,  and  various  carboxylate  anions  can  accept  the  coenzyme  A  so 
that  acetoacetate  is  formed: 

2  Acetyl-CoA   :^   acetoacetyl-CoA  +  CoA 
Acetoacetyl-CoA  -|-  malonate   :;^    malonyl-CoA  +  acetoacetate 

succinate  and  butyrate  being  normally  the  most  active.  In  this  way  mal- 
onate can  give  rise  to  acetoacetate  as  well  as  by  its  block  of  the  cycle. 
Labeled  malonate  forms  labeled  acetoacetate  in  rat  liver  (Nakada  et  al., 
1957).  This  type  of  reaction  has  also  been  shown  to  occur  in  yeast,  and  dog 
heart  and  skeletal  muscle  (Menon  and  Stern,  1960).  The  enzyme,  succinyl- 
/3-ketoacyl-CoA  transferase,  was  purified  from  pig  heart  and  catalyzes  the 
transfer  to  both  malonate  and  glutarate.  The  following  reaction: 

Succinyl-CoA  +  malonate   :^   malonyl-CoA  +  succinate 

also  occurs.  It  is  interesting  to  speculate  that  some  of  the  succinyl-CoA 
formed  in  the  oxidation  of  a-ketoglutarate  transfers  its  coenzyme  A  to 
malonate  when  it  is  present;  if  the  malonyl-CoA  is  not  readily  metabolized, 
this  could  deplete  the  a-ketoglutarate  oxidase  of  coenzyme  A  and  slow  down 
the  reaction.  Condensation  of  malonyl-CoA  with  coenzyme  A  derivatives 
may  be  important  in  fatty  acid  synthesis.  In  pigeon  liver  and  carrot  roots 
there  is  an  enzyme  catalyzing  the  condensation  of  malonyl-CoA  with  either 
acetyl-CoA  or  butyryl-CoA;  although  the  product  is  unknown,  it  was  isolat- 
ed chromatographically  (Steberl  et  al.,  1960).  This  product  can  form  pal- 
mitate  with  other  enzyme  fractions.  When  malonyl-2-C^^-CoA  is  incubated 
with  extracts  from  various  rat  tissues,  various  Cig-Cig  acids  are  formed, 
depending  on  the  acyl-CoA  acceptor  used  (Horning  et  al.,  1960).  One  acyl- 
CoA  unit  is  incorporated  into  long-chain  fatty  acids  and  the  rest  of  the 
C-chain  in  supplied  from  malonyl-CoA.  Labeled  fatty  acids  are  also  formed 
from  malonate- 1-C^*  in  particle  suspensions  of  the  locust  fat  body  (Tietz, 
1961).  A  highly  purified  preparation  from  pigeon  liver,  which  converts 
malonyl-CoA  and  acetyl-CoA  to  palmitate  in  the  presence  of  NADPH,  has 
been  reported  (Bressler  and  Wakil,  1961).  In  the  absence  of  NADPH,  mal- 
onyl-CoA  and  acetyl-CoA  condense  to  form  an  unknown  product  (which 
is  not  acetoacetate,  butyrate,  or  /^-hydroxybutyrate).  The  conversion  of 
malonyl-CoA  to  fatty  acids  is  perhaps  mediated  through  such  condensa- 
tions to  Cg  acids,  forming  butyryl-CoA,  which  would  again  condense  with 
malonyl-CoA,  and  so  lengthen  the  chain. 

A  few  words  should  be  said  about  the  pathways  of  methylmalonate  me- 
tabolism since  it  has  been  recently  found  to  be  an  important  intermediate 
in  fatty  acid  |metabolism,  and  malonate  can  interfere  markedly  with  at 
least  one  of  these  reactions.  Methylmalonate  was  shown  to  be  an  interme- 
diate in  the  metabolism  of  propionate  in  various  tissues  (Flavin  et  al.,  1955; 
Katz  and  Chaikoff,  1955;  Feller  and  Feist,  1957).  In  the  course  of  this  work 


INHIBITORS    STRUCTURALLY    RELATED    TO    MALONATE  235 

a  novel  reaction  was  discovered,  namely,  the  interconversion  of  methyl- 
malonate  and  succinate.  The  enzyme,  which  has  been  called  a  methyl- 
malonyl-CoA  isomerase,  catalyzes  the  reaction: 

Methylmalonyl-CoA   :^   succinyl-CoA 

and  has  been  purified  from  ox  liver  (Stern  and  Freidman,  1960)  and  Pro- 
fionihacterium  shermanii  (Wood  and  Stjernhohn,  1961;  Stjernholm  and 
Wood,  1961).  At  equilibrium  the  ratio  (succinyl-CoA)/(methylmalonyl-CoA) 
is  10.5  (pH  7  and  25°).  This  reaction  functions  in  both  the  oxidation  and 
formation  of  propionate.  Another  reaction  of  importance  in  propionate 
metabolism  is 

Methylmalonyl-CoA  +  pyruvate   :^   propionyl-CoA  +  oxalacetate 

It  is  catalyzed  by  methylmalonyl-oxalacetate  transcarboxylase,  whereby  a 
carboxylation  may  be  effected  without  the  intervention  of  COg  or  the  ex- 
penditure of  energy  to  activate  COg.  It  may  finally  be  noted  that  the  feed- 
ing of  malonate  to  dogs  leads  to  a  marked  excretion  of  methylmalonate  in 
the  urine  (Thomas  and  Stalder,  1958).  This  does  not  necessarily  mean  that 
the  methylmalonate  is  formed  from  malonate,  since  malonate  inhibits  the 
interconversion  of  methylmalonyl-CoA  and  succinyl-CoA  very  strongly 
(Flavin  et  al.,  1955). 


INHIBITORS   STRUCTURALLY    RELATED   TO    MALONATE 

The  inhibition  of  succinate  dehydrogenase  by  various  dicarboxylate  ions 
was  treated  earlier  (page  34).  Maleate  will  be  dealt  with  in  Chapter  III-2 
and  oxalate  will  be  discussed  in  Volume  IV.  Inhibitors  related  to  malonate 
will  also  be  found  in  the  following  chapter  on  analogs.  It  then  remains  to 
take  up  the  esters  of  malonate,  hydroxy  malonate,  and  those  compounds 
in  which  the  carboxylate  groups  have  been  replaced  by  other  anionic  groups. 
As  mentioned  previously,  there  is  often  confusion  in  the  nomenclature;  we 
shall  adopt  the  following  (using  the  ethyl  derivatives  as  examples): 

COO"  Et^    /COO" 

Et— Hc:"       .  /C^       _ 

coo  Et  COO 

Ethylmalonate  Diethylmalonate 


/COO  /COO— Et 

HgC  Hz^s, 

^COO— Et  ^COO  — Et 


Malonic  Malonic 

monoethyl  ester  diethyl  ester 


236  1.    MALONATE 

Malonic  Esters 

Malonic  monoethyl  and  diethyl  esters  have  been  found  to  have  some  in- 
teresting actions.  They  can  increase  the  survival  period  of  mice  infected 
with  mycobacteria  (Davies  et  al.,  1956),  are  occasionally  carcinostatic  (Freed- 
lander  et  al.,  1956),  can  inhibit  the  breakdown  of  hexobarbital  in  liver  homo- 
genates  and  prolong  the  narcotic  action  (Kramer  and  Arrigoni-Martelli, 
1960),  and  inhibit  sporulation  of  Bacillus  cereus  (Nakata  and  Halvorson, 
1960).  However,  the  relation  of  these  effects  to  succinate  dehydrogenase 
inhibition  is  obscure.  In  most  cases,  malonic  esters  have  been  used  to  cir- 
cumvent the  permeability  barriers  to  malonate,  inasmuch  as  the  esters 
should  penetrate  into  cells  readily.  It  has  often  been  assumed  that  hydrolysis 
to  malonate  occurs  within  the  cells.  This  hydrolysis  must  be  enzymatic 
because  the  esters  are  quite  stable.  A  lipase  from  pig  liver  hydrolyzes  one 
ethyl  group  from  malonic  diethyl  ester  but  does  not  remove  the  other 
ethyl  group,  the  product  being  malonic  monoethyl  ester  (Christman  and 
Lewis,  1921).  I  have  been  able  to  find  no  direct  evidence  for  the  hydrolysis 
to  malonate.  Malonic  esters  are  neutral  at  physiological  pH's,  since  they 
are  very  weak  acids  with  pK„  values  around  15.75  (Rumpf  et  al.,  1955) 
and  ionize  very  slowly  with  a  rate  constant  of  1.8  X  10~^  min~^  (Pearson 
and  Mills,  1950). 

The  effects  of  the  malonic  esters  on  metabolism  will  now  be  discussed  in 
order  to  determine  if  there  is  any  indirect  evidence  for  the  intracellular 
hydrolysis  to  malonate  and  the  inhibition  of  succinate  dehydrogenase. 
When  injected  into  fluoroacetate-poisoned  rats,  malonate  and  malonic  di- 
ethyl ester  have  approximately  the  same  effects  on  the  accumulation  of 
citrate  in  the  heart  and  kidneys  (Fawaz  and  Fawaz,  1954).  Furthermore,  in 
kidney  slices,  the  diethyl  ester  inhibits  succinate  oxidation  92%  at  20  mM, 
and  malonate  at  the  same  concentrazion  inhibits  75%.  Less  inhibition  by 
the  ester  compared  to  malonate  is  observed  in  heart  slices.  Evidence  for 
hydrolysis  by  both  tissues  was  adduced  from  the  decreases  in  the  pH  ob- 
served. The  respiration  of  the  fungus  Zygorrhynchus  moelleri  is  not  inhibited 
readily  by  malonate  although  the  succinate  dehydrogenase  from  this  orga- 
nisms is  quite  sensitive,  indicating  a  failure  to  penetrate  (Moses,  1955). 
Malonic  diethyl  ester  was  tested  and  found  to  inhibit  the  oxidation  of  both 
glucose  and  acetate,  but  at  high  concentrations  (see  tabulation).  It  was  felt 


Malonic  diethyl 

ester 

%  Inhibition  of: 

{mM) 

Glucose  oxidation 

Acetate  oxidation 

10 

30 

100 

Stim  49 
30 
96 

10 

76 
98 

INHIBITORS    STRUCTURALLY    RELATED    TO    MALONATE  237 

that  the  greater  inhibition  of  acetate  oxidation  is  evidence  for  the  hydro- 
lysis of  the  ester  and  that  the  inhibition  is  exerted  by  malonate.  Malonic 
diethyl  ester  was  also  used  to  facilitate  penetration  into  Penicillium  chryso- 
genum,  since  even  100  roM  malonate  does  not  effect  acetate  metabolism 
(Goldschmidt  et  al.,  1956).  The  ester  at  20  milf  inibits  the  production  of 
C^*02  from  labeled  acetate  75-85%  and  simultaneously  decreases  the  in- 
corporation of  C^*  into  cellular  materials,  the  labeling  of  glutamate  being 
particularly  depressed.  The  utilization  of  acetate  by  Bacillus  cereus  is  also 
interfered  with  by  malonic  diethyl  ester,  so  that  acetate  and  p>Tuvate  ac- 
cumulate (Nakata  and  Halvorson,  1960).  Malonate  and  the  diethyl  ester 
have  been  compared  with  respect  to  their  effects  on  the  respiration  of 
Mycobacterium  pJdei  (Miiller  et  al.,  1960).  Malonate  stimulates  the  endoge- 
nous respiration,  presumably  through  its  oxidation,  and  the  ester  stimulates 
even  more  potently  at  1-10  raM,  although  at  100  mM  the  ester  inhibits 
and  malonate  still  stimulates.  The  respiration  with  glycerol  as  the  substrate 
behaves  similarly.  Finally,  malonic  diethyl  ester  markedly  stimulates  the 
endogenous  respiration  of  Chlorella  vulgaris  at  4-10  mM,  but  inhibits  the 
oxidation  of  glucose  and  acetate,  the  latter  more  strongly  (Merrett  and 
Syrett,  1960).  All  of  these  results  show  that  the  diethyl  ester  is  inhibitory 
but  certainly  do  not  constitute  conclusive  evidence  for  a  hydrolysis  to  mal- 
onate. This  is  a  subject  that  should  be  pursued  further  and  more  extensive 
tests  should  be  made  for  enzymes  hydrolyzing  the  esters.  Until  the  intra- 
cellular hydrolysis  can  be  established,  the  results  obtained  with  the  malonic 
esters  cannot  be  interpreted. 

Hydroxy  malonate  (Tartronate) 

The  substitution  of  a  methylene  hydrogen  of  malonate  by  any  group 
seems  to  reduce  rather  strongly  the  ability  to  inhibit  succinate  dehydrogen- 
ase. Even  the  small  hydroxyl  group  almost  abolishes  the  inhibitory  activity 
and  this  is  evidence  that  the  binding  of  malonate  to  the  active  center  of 
succinate  dehydrogenase  must  involve  severe  steric  restrictions.  Although 
tartronate  does  not  inhibit  succinate  dehydrogenase,  it  has  other  actions 
of  some  interest.  Quastel  and  Wooldridge  (1928)  found  that  71.4  roM  tar- 
tronate does  not  inhibit  succinate  oxidation  at  all  whereas  it  inhibits  the 
oxidation  of  lactate  90%,  as  measured  by  methylene  blue  reduction  in  tol- 
uene-treated E.  coli.  In  fact,  the  marked  differences  in  the  effects  of  mal- 
onate and  tartronate  led  Quastel  and  Wooldridge  to  postulate  the  specific 
structure  of  the  active  centers  of  enzymes.  Tartronate  inhibits  the  respira- 
tion of  rat  liver  slices  and  is  strongly  ketogenic,  acetoacetate  being  formed 
to  a  greater  extent  that  with  equivalent  concentrations  of  malonate  (Ed- 
son,  1936).  Pig  heart  malate  dehydrogenase  is  inhibited  24%  by  60  mM 
tartronate  (Green,  1936)  but  in  pigeon  liver  extracts  it  is  about  1000  times 
as  effective,  inhibiting  malate  oxidation  competitively  with  a  ^^  of  0.09  mM 


238  1.    MALONATE 

(Scholefield,  1955).  Tartronate  is  also  a  competitive  inhibitor  of  the  decar- 
boxylating  malate  dehydrogenase  (malic  enzyme)  of  pigeon  liver  with  a  K^ 
of  0.1  mM,  the  inhibition  being  stronger  than  with  malonate  (Stickland, 
1959  b).  The  carboxylation  of  pyruvate,  catalyzed  by  the  same  enzyme, 
is  also  strongly  inhibited,  but  there  is  a  large  noncompetitive  element 
(Stickland,  1959  a).  The  inhibition  of  lactate  and  malate  oxidations  and 
decarboxylations  is  not  surprising  since  tartronate  is  structurally  similar. 
Although  tartronate  occurs  in  plant  and  animal  tissues  to  the  extent  of 
8-15 //g/g  wet  weight  of  tissue,  which  would  correspond  to  about  0.1  xnM 
in  the  tissue  water  (Veitch  and  Brierley,  1962),  it  is  doubtful  if  it  could 
exert  a  regulatory  action  on  the  metabolism.  Tartronate  is  not  metabolized 
in  the  rat  and  consequently  is  near  8  raM  in  the  urine. 

The  respiration  of  guinea  pig  brain  slices  is  depressed  by  tartronate  at 
concentrations  of  67-75  rcvM  (Jowett  and  Quastel,  1937).  The  degree  of  inhi- 
bition depends  on  the  substrate  provided  and  is  maximal  with  lactate  and 
minimal  with  pyruvate.  Since  lactate  is  probably  metabolized  through 
pyruvate,  the  inhibition  here  may  be  mainly  on  lactate  dehydrogenase. 
However,  the  anaerobic  breakdown  of  pyruvate  and  anaerobic  glycolysis 
are  also  well  inhibited.  The  respiration  of  Mycohacterium.  phlei  with  lactate 
as  substrate  is  inhibited  65%  with  66  mM  tartronate,  although  the  endo- 
genous respiration  is  stimulated  and  the  oxidation  of  glucose  unaffected 
(Edson  and  Hunter,  1947).  This  relatively  specific  effect  on  lactate  dehydro- 
genase was  used  by  Fiume  (1960)  to  inhibit  aerobic  glycolysis  in  tumor  cells, 
inasmuch  as  lactate  dehydrogenase  is  involved.  It  was  postulated  that 
aerobic  glycolysis  might  be  inhibited  more  strongly  in  the  tumor  than  in 
normal  tissues,  depleting  the  ATP  supply  more  severely.  It  was  found  that 
tartronate  inhibits  aerobic  glycolysis  of  the  Yoshida  ascites  hepatoma  — 
26%  at  10  mM,  34%  at  20  mM,  and  58%  at  50  mM  -  but  comparisons 
with  normal  tissue  were  not  made. 

The  inhibition  of  phosphatases  should  perhaps  also  be  considered  in 
work  with  tartronate,  since  prostatic  acid  phosphatase  is  inhibited  com- 
petitively with  a  K^  of  around  50  mM.  The  inhibition  by  tartronate  is 
much  greater  than  by  malonate  and  about  twice  as  potent  as  by  ketomal- 
onate  (Kilsheimer  and  Axelrod,  1957). 

Aminomalonate 

This  substance  was  considered  to  be  a  possible  substrate  for  the  synthesis 
of  S-aminolevulinate  but  was  found  upon  examination  to  be  a  potent  inhib- 
itor of  S-aminolevulinate  synthetase  (Matthew  and  Neuberger,  1963).  The 
inhibition  of  the  enzyme  from  Rhodopseiidomonas  spheroides  and  chicken 
erythrocytes  is  competitive;  the  K,  for  the  bacterial  enzyme  is  0.0225  mM. 
Pyridoxal-P  is  a  coenzyme  in  these  systems  and  the  inhibition  depends  on 
its  concentration.  Aminomalonate  may  be  considered  to  be  an  analog  of 


INHIBITORS    STRUCTURALLY    RELATED    TO    MALONATE  239 

glycine  (it  is  carboxyglycine)  and  presumably  inhibits  8-aminolevulinate 
synthetase  by  binding  to  the  glycine  site  and  complexing  with  pyridoxal-P 
as  well.  Aminomalonate  condenses  with  aldehydes  nonenzymatically  in 
the  presence  of  pyridoxal-P.  Furthermore,  other  pyridoxal-P-dependent 
enzymes  are  inhibited,  e.g.  serine  hydroxy methyltransferase,  whereas  en- 
zymes involved  in  glycine  metabolism  but  not  requiring  pyridoxal-P  are  not 
inhibited.  Aminomalonate  can  be  formed  in  the  tissues  by  transamination 
between  ketomalonate  and  glutamate,  and  can  be  decarboxylated  by  an 
enzyme  found  in  silkworm  larvae  and  rat  heart  and  liver  to  glycine.  This 
derivative  of  malonate  is  a  good  illustration  of  how  a  simple  change  in  the 
structure  can  create  an  inhibitor  with  quite  different  properties  and  inhib- 
itory spectrum. 

Substituted   Malonates 

Although  the  alkylmalonates  are  not  particularly  interesting  as  inhibi- 
tors, there  are  two  malonate  derivatives  that  may  warrant  further  investi- 
gation. Fluoromalonate  was  studied  by  Chari-Bitron  (1961)  on  the  principle 
that  if  malonate  is  metabolized  through  acetyl-CoA,  fluoromalonate  might 
follow  the  same  pathway  and  enter  the  cycle  as  fluoroacetyl-CoA,  produc- 
ing the  same  effects  as  fluoroacetate,  namely,  a  block  of  the  cycle  at  the 
aconitase  step.  The  toxicity  of  fluoromalonate  is  a  good  deal  less  than 
fluoroacetate  but  the  ester  is  as  toxic  in  mice  (see  accompanying  tabulation). 


LD50  (mg/kg) 

Animal  

Fluoromalonate        Fluoromalonic  diethyl  ester  Fluoracetate 


Mouse 

80 

Rat 

60 

Guinea  pig 

2 

15  15 

70  5 

—  0.25 


Death  is  associated  with  a  marked  accumulation  of  citrate  in  the  tissues 
and  it  differs  strongly  from  malonate  in  this  respect.  Accumulation  of  cit- 
rate also  occurs  in  kidney  mitochondria  with  fluoromalonate  at  concentra- 
tions around  1  mM.  It  was  further  established  that  decarboxylation  of 
fluoromalonate  occurs  in  kidney  preparations.  Finally,  fluoromalonate  is 
only  about  one  tenth  as  effective  as  malonate  in  the  inhibition  of  succinate 
dehydrogenase.  The  results  thus  conform  quite  well  to  the  predicted  mech- 
anism. Difluoromalonate  and  its  amide  inhibit  quite  readily  the  oxida- 
tions of  succinate  and  fumarate  by  Pseudomoyias  (around  70%  inhibition 
at  0.7  mM),  but  there  is  no  inhibition  of  succinate  dehydrogenase  in  soni- 
cates; the  mechanism  is  unknown  (Bernheim,  1963). 


240  1.    MALONATE 

Inasmuch  as  the  active  center  of  succinate  dehydrogenase  possesses  a 
sulfhydryl  group  close  to  the  cationic  binding  sites,  this  being  the  basis 
for  the  inhibition  by  mercurials  and  other  sulfhydryl  agents,  the  possibility 
of  combining  a  sulfhydryl  inhibition  with  malonate  presents  itself.  Mercuri- 
malonamide  and  mercurimalonic  diethyl  ester  have  been  prepared  (Naik 
and  Patel,  1932)  and  these  compounds,  or  particularly  the  hydrolyzed 

^COO-Et 
^C(X>-Et 

forms  if  they  are  stable,  might  be  interesting  to  examine  as  inhibitors  of 
succinate  dehydrogenase. 

Acetylene-dicarboxylate  and    Propane-tricarboxylate 

Succinate  dehydrogenase  is  inhibited  competitively  by  acetylene-dicar- 
boxylate (Dietrich  et  at.,  1952).  The  order  of  addition  of  the  succinate  and 
acetylene-dicarboxylate  is  important;  for  example,  if  acetylene-dicarboxy- 
late is  added  20  min  after  the  succinate,  the  rate  of  oxygen  uptake  de- 
creases slowly  and  does  not  become  equal  to  that  observed  when  the  suc- 
cinate is  added  after  the  inhibitor  until  5  hr  (Thomson,  1959).  Acetylene- 
dicarboxylate  is  about  as  effective  as  malonate  on  long  incubation  with  the 
enzyme.  The  constants  obtained  on  rat  kidney  succinate  dehydrogenase 
are  ir,„  =  4.12  mM  and  K^  =  0.81  roM  when  substrate  and  inhibitor  are 
added  together;  after  18  hr  incubation  with  the  inhibitor,  iC,  =^  0.171  vaM. 
It  is  difficult  to  understand  why  the  rate  of  inhibition  is  so  slow.  Succinate 
dehydrogenase  from  pig  heart  is  less  readily  inhibited,  K^  being  1.4  mM 
with  A'^-methylphenazine  as  electron  acceptor  and  16.5  when  ferricyanide  is 
the  acceptor  (Hellerman  et  al.,  1960).  Possibly  insufficient  time  for  equilib- 
rium was  allowed.  The  inhibitions  of  succinate  and  pyruvate  oxidations 
by  acetylene-dicarboxylate  in  suspensions  of  rate  heart  mitochondria  are 
shown  in  Fig.  1-23  (Montgomery  and  Webb,  1956  b).  Acetylene-dicarboxy- 
late is  less  potent  than  malonate  against  succinate  oxidation  and  more  po- 
tent against  pyruvate  oxidation,  indicating  that  an  inhibition  is  exerted 
at  some  other  point  in  the  cycle. 

Propane-tricarboxylate  is  a  rather  weak  inhibitor  of  succinate  dehydro- 
genase from  rat  heart,  5  mM  inhibiting  around  30%  when  the  succinate  is 
also  5  mM.  The  oxidation  of  a-ketoglutarate  is  inhibited  60%  under  the 
same  conditions,  suggesting  some  effect  on  the  a-ketoglutarate  oxidase. 
The  oxidation  of  pyruvate  is  inhibited  even  more  strongly  (Fig.  1-24).  It 
might  be  thought  that  propane-tricarboxylate  would  inhibit  aconitase  or 
isocitrate  dehydrogenase,  but  very  little  inhibition  is  noted,  either  of  cit- 
rate oxidation  or  the  ability  of  citrate  to  function  as  a  source  of  oxalac- 
etate  for  the  oxidation  of  pyruvate. 


INHIBITORS    STRUCTURALLY    RELATED    TO    MALONATE 


241 


Traws-l,2-Cyclopentane-dicarboxylate  inhibits  the  succinate  dehydroge- 
nase of  Tetrahymena  geleii  homogenates,  around  50%  inhibition  being  ob- 
served when  the  inhibitor/substrate  ratio  is  0.5  (Seaman  and  Houlihan, 
1950).  On  the  other  hand,  in  intact  cells  this  substance  increases  the  utili- 
zation of  pyruvate,  acetate,  and  succinate.  The  uptake  of  acetate  is  in- 


80 

- 

Pyruvote  +  Malate 

^^ 

%60 
INH 

- 

/ 

40 

■ 

/ 

20 

- 

1                    1                    i 

Succinate 

I  5  10  50 

(Acetylene  -  Dicarboxylate) 

Fig.  1-23.  Effects   of  acetylene-dicarboxylate    (in  milf) 

on  the  oxidations  of  p>Tuvate  -f  malate  and  succinate 

by    rat    heart    mitochondria.    (From    Montgomery    and 

Webb,   1956  b.) 

creased  as  much  as  50%  by  ^ra??5-l,2-cyclopentane-dicarboxylate  and  in 
its  presence  succinate  is  taken  up  and  oxidized  whereas  succinate  does  not 
enter  the  cells  normally.  It  was  postulated  that  trans-1, 2  cyclopex\':ine- 
dicarboxylate  increases  the  permeability  of  the  membrane  to  these  sub- 
strates, possibly  by  an  action  on  the  metabolic  systems  involved  in  the 


)5  I  2  5 

(  Propone  -  Tncarboxylote) 

Fig.    1-24.    Effect    of   propane -tricarboxylate    (in 

inM)  on  the   oxidation  of  pyruvate  +  malate  by 

rat  heart    mitochondria.    (From   Montgomery   and 

Webb,  1956  b.) 


242  1.    MALONATE 

inward  transport.  It  is  not  known  if  such  an  effect  is  observed  with  other 
dicarboxylate  anions. 

Sulfonate,  Phosphonate,  and  Arsenate  Analogs  of  Succinate 

Succinate  and  malonate  are  bound  to  the  active  center  of  succinate  de- 
hydrogenase in  part  through  electrostatic  forces  involving  the  negatively 
charged  carboxylate  groups.  It  might  be  expected  that  substances  in  which 
the  carboxylate  groups  are  replaced  by  other  anionic  groups  would  also  be 
inhibitory.  The  structures  for  some  of  the  compounds  and  the  intercharge 
distances  are  given  in  Table  1-1.  Klotz  and  Tietze  (1947)  first  demonstrated 
the  inhibition  of  succinate  dehydrogenase  (rat  liver)  by  1,2-ethanedisulfo- 
na'te  and  /5-sulfopropionate,  50%  inhibition  being  observed  in  both  cases 
when  the  inhibitor  is  13  mM  and  succinate  is  20  mM,  these  inhibitions 
being  approximately  equivalent  to  those  of  malonate.  Intact  E.  coli  cells 
are  not  affected  by  1,2-ethanedisuIfonate  but  the  oxidation  of  succinate 
by  cell-free  preparations  is  well  inhibited,  corresponding  to  the  results  with 
malonate  (Ajl  and  Werkman,  1948).  Some  inhibitor  constants  for  these 
substances  are  shown  in  Table  1-29. 

There  appears  to  be  definite  differences  in  the  susceptibility  of  the  suc- 
cinate dehydrogenases  from  different  sources.  Seaman  (1952)  noted  that 
/?-phosphonopropionate  inhibits  the  Tetrahymena  enzyme  more  strongly 
than  does  malonate,  and  yet  no  inhibition  is  observed  on  the  enzymes 
from  rat  heart,  liver,  brain,  or  muscle.  The  inhibitions  are  competitive 
wherever  they  have  been  tested  and  there  is  no  evidence  that  the  mecha- 
nism is  in  any  way  different  from  that  of  malonate.  Since  these  groups  are 
larger  than  the  carboxylate  group,  the  interchange  distances  are  greater 
than  in  malonate  or  succinate,  and  this  must  play  some  role  in  determining 
their  binding  to  the  enzyme.  However,  the  distances  for  the  malonate  ana- 
logs, methanedisuffonate  and  arsonoacetate,  are  between  those  for  malon- 
ate and  succinate,  so  that  binding  should  be  as  tight  as  for  these  latter 
substances  if  this  were  the  only  factor.  Another  factor  of  importance  is  the 
total  net  charge  on  these  inhibitors,  inasmuch  as  each  of  the  sulfonate, 
phosphonate,  and  arsonate  groups  can  ionize  more  than  once.  In  other 
words,  a  disulfonate  can  exist  in  five  different  forms  with  charges  0,-1, 
—  2,  —3,  and  —4.  Furthermore,  the  third  and  fourth  ionization  constants 
are  in  the  physiological  range  of  pH  (see  accompanying  tabulation).  The 


vK„  vK„ 


Arsonoacetate 

7.7 

— 

1 ,2-Ethanediphosphonate 

6.84 

8.17 

1 ,4-Butanediphosphonate 

7.28 

9.05 

Pyrophosphate 

5.69 

7.76 

INHIBITORS    STRUCTURALLY    RELATED    TO    MALONATE 


243 


Table  1-29 

Inhibitor  Constants  for  Sulfonate,  Phosphonate,  and  Arsonate  Analogs  of 

Substrates " 


Inhibitor 

Preparation 

K, 

Reference 

Malonate 

Tetrahymena  succinate 

6.67 

Seaman  (1952) 

/S-Phosphonopropionate 

dehydrogenase 

1.51 

Arsonoacetate 

9.25 

1,2-Ethanedisulfonate 

Rat  liver  succinate 

2.73 

Klotz  and  Tietze 

/?-  Sulfopropionate 

dehydrogenase 

3.69 

(1947) 

Malonate 

Mouse  liver  succinate 

0.19 

Tietze  and  Klotz 

1 ,2-Ethanedisulfonate 

dehydrogenase 

26.5 

(19.52) 

o-Sulfobenzoate 

11.7 

/5-Phosphonopropionate 

No  inh 

Arsonoacetate 

No  inh 

Methionate 

20.6 

Malonate 

Beef  heart  succinate 

0.014 

Rosen  and  Klotz 

Methanediphosphonate 

dehydrogenase 

0.5 

(1957) 

Phosphonoacetate 

0.15 

Pyrophosphate 

0.0011 

Hs^pophosphate 

0.14 

1,2-Ethanediphosphonate 

9.6 

1,4-Butanediphosphonate 

8.9 

Arsonoacetate 

15.3 

Malonate 

Fumarase 

40 

Massey  (1953b) 

a-Hydroxy-/3-sulfo- 

propionate 

16.5 

"  The  A',  values  for  rat  and  mouse  liver  succinate  dehydrogenases  were  cal- 
culated assuming  A'^  =  6  X  10~^  M,  and  for  beef  heart  succinate  dehydrogenase 
K^  =  4  X  10"*  M.  Although  these  values  are  undoubtedly  inaccurate,  the  Aj  values 
so  calculated  are  useful  for  comparisons  since  within  each  experiment  the  relative 
values  are  reliable. 


rather  poor  inhibition  produced  by  the  arsono  and  phosphono  derivatives 
was  postulated  by  Tietze  and  Klotz  (1952)  as  possibly  due  to  the  extra 
negative  charge  carried  bj^  these  substances.  In  the  more  recent  work  of 
Rosen  and  Klotz  (1957),  the  ionization  was  taken  into  account  and  the 
inhibitor  constants  calculated  on  the  basis  of  the  concentration  of  doubly 
charged  anion  present.  Evidence  that  too  high  a  charge  reduces  the  inhib- 
itory potency  is  provided  by  the  falling  off  of  the  inhibition  as  the  pH  is 
raised.  On  the  other  hand,  it  is  difficult  to  understand  why  there  should 


244  1-    MALONATE 

be  less  affinity  between  cationic  groups  on  the  enzyme  and  doubly  charged 
anions,  compared  to  singly  charged  anions,  and  one  cannot  help  but  wonder 
if  the  alterations  in  pH  in  the  experiments  of  Kosen  and  Klotz  were  not 
affecting  the  ionizable  groups  on  the  enzyme.  These  workers  suggested  that 
the  binding  of  these  substances  involves  an  iron  ion  on  the  enzyme  and 
correlated  affinities  with  the  ability  to  chelate  iron.  This  is  certainly  a  type 
of  binding  that  should  be  kept  in  mind,  but  it  is  difficult  to  reconcile  with 
the  fact  that  iron-chelating  agents  such  as  1,10-phenanthroline  and  2,2'-bi- 
pyridine  do  not  interfere  with  the  binding  of  succinate  to  the  dehydrogenase 
even  though  their  attachment  to  the  enzyme  can  be  demonstrated  spectro- 
scopically. 

An  interesting  succinate  dehydrogenase  inhibitor,  not  fundamentally  re- 
lated to  the  compounds  previously  discussed,  is  3-nitropropionate  (hiptage- 
nate),  found  to  be  the  toxic  principle  of  Indigofera  endecaphylla  (Morris 
et  al.,  1954).  It  inhibits  succinate  dehydrogenase  competitively  with  K^  == 
0.19  n\M  (Hylin  and  Matsumoto,  1964,)  although  no  inhibition  of  the 
enzyme  is  found  after  administration  of  the  substance  to  animals,  so  that 
it  is  not  possible  to  correlate  the  toxicity  with  an  effect  on  this  enzyme.  It 
is  rather  surprising  that  3-nitropropionate  binds  so  well  to  succinate  de- 
hydrogenase, lacking  two  anionic  groups,  and  it  would  be  interesting  to 
know  more  of  the  nature  of  the  interaction  of  the  nitro  group  with  the 
enzyme. 

With  respect  to  inhibitors  related  to  malonate  in  one  way  or  another, 
it  must  be  concluded  that  none  possesses  particular  advantages  over  mal- 
onate for  the  specific  inhibition  of  succinate  dehydrogenase,  although  cer- 
tain derivatives  have  interesting  properties  and  metabolic  actions.  Progress 
could  be  made  by  the  finding  of  forms  of  malonate,  or  related  inhibitors, 
that  are  uncharged,  reasonably  stable  outside  the  cell,  and  easily  split 
to  the  active  inhibitor  intracellularly.  It  would  also  be  valuable  to  have  an 
inhibitor  which  would  initially  bind  to  the  succinate  dehydrogenase  at  the 
substrate  site,  because  of  its  complementary  configuration,  and  then  react 
chemically  with  an  adjacent  group,  so  that  the  inhibition  of  this  enzyme 
would  be  not  only  specific  but  slowly  reversible. 


CHAPTER  2 

ANALOGS  OF  ENZYME  REACTION 
COMPONENTS 


An  enzyme-catalyzed  reaction  involves  the  combination  of  the  com- 
ponents with  specific  sites  on  the  apoenzyme  protein  surface,  these  areas 
possessing  the  particular  molecular  configuration  and  the  electrical  field 
distribution  required  for  the  attachment  and  the  electronic  displacements 
characterizing  the  activated  complex.  If  one  of  these  components  is  modified 
in  any  way,  its  behavior  in  the  system  will  usually  be  altered,  due  primarily 
to  the  new  pattern  of  interaction  between  the  modified  component  and 
the  enzyme.  The  development  and  study  of  such  analogs  of  substrates  and 
coenzymes  have  been  very  active  fields  during  the  past  few  years  for  several 
reasons.  First,  the  determination  of  the  relative  aflfinities  of  analogs  that 
are  substrates  or  inhibitors  for  enzymes  is  one  of  the  most  effective  means 
for  analyzing  the  topography  of  active  centers  and  establishing  the  types 
of  interaction  involved  in  the  catalysis.  Second,  it  is  hoped  that  inhibitors 
more  specific  for  blocking  certain  enzymes  than  the  inhibitors  previously 
available  will  be  found  and  this  has  been  justified  to  a  certain  extent. 
Third,  it  has  been  realized  that  analog  inhibition  has  direct  bearing  on  the 
important  phenomenon  of  feedback  control  of  metabolic  sequences  and  on 
the  general  regulation  of  cellular  metabolism.  Last,  it  is  anticipated  that 
the  use  of  proper  analogs  may  be  useful  in  the  specific  correction  of  certain 
abnormal  metabolic  patterns  and  growth  processes,  such  as  occur  in  here- 
ditary enzyme  defects  or  neoplastic  changes. 

The  previous  chapter  is  concerned  with  malonate,  a  classic  analog  inhib- 
itor, and  in  the  present  chapter  it  is  proposed  to  extend  this  principle  to 
a  variety  of  enzymes  in  order  to  establish  some  basic  concepts.  There  are 
a  number  of  inhibitors  which  act,  at  least  occasionally,  because  they  are 
structurally  related  to  some  enzyme  reaction  component,  but  which  for 
various  reasons  will  be  discussed  in  separate  chapters.  Such  are  carbon 
monoxide,  fluoroacetate  and  fluorocitrate,  parapyruvate,  arsenate,  pyro- 
phosphate, monoamine  oxidase  inhibitors,  certain  inhibitors  of  cholinester- 
ase,  and  various  drugs.  Furthermore,  it  is  necessary  to  point  out  that  no 
attempt  will  be  made  to  review  the  vast  literature  on  the  depression  of 

245 


246  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

growth  and  proliferation  exerted  by  many  analogs,  inasmuch  as  my  pur- 
pose here  is  to  restrict  the  discussion  to  enzymic  and  metabolic  levels. 

TERMINOLOGY 

The  term  analog  is  defined  very  broadly  as  any  substance  that  is  in  some 
way  structurally  related  to  a  substrate,  coenzyme,  or  cofactor.*  It  may 
either  participate  in  the  enzyme  reaction  to  a  greater  or  lesser  extent  than 
the  normal  components,  or  inhibit  the  reaction  by  interfering  with  the 
functioning  of  these  normal  components.  The  commonly  used  term  anti- 
metabolite generally  implies  that  the  substance  is  a  biologically  abnormal 
compound  synthesized  in  the  laboratory  and  capable  of  interfering  with  the 
reactions  of  some  cellular  metabolite.  We  shall  in  this  chapter  frequently 
be  concerned  with  inhibitions  produced  by  substances  such  as  carbohy- 
drates, amino  acids,  purines,  and  nucleotides,  which  naturally  occur  in  most 
cells,  and  thus  the  more  general  term  analog  is  preferred.  The  use  of  the 
term  isostere  has  generally  been  restricted  to  a  substance  produced  by  the 
substitution  of  an  atom  or  group  in  the  normal  compound  by  another  atom 
or  group  with  similar  electronic  or  steric  properties.  A  homolofj  is  a  member 
of  a  series  in  which  some  part  or  property  of  a  basic  chemical  type  is  progres- 
sively varied,  as  is  the  case  when  an  aliphatic  chain  is  lengthened  by  adding 
successive  methylene  groups.  Most  of  the  analogs  to  be  discussed  therefore 
fall  in  to  one  or  more  of  these  latter  categories,  but  it  is  felt  that  there  is 
little  benefit  to  be  derived  from  using  these  more  specific  terms. 

POSSIBLE  SITES  AND  MECHANISMS  OF  INHIBITION 

The  most  common  mechanism  of  inhibition  is  a  competition  between 
the  analog  and  the  normal  reactant  for  a  specific  site  on  the  enzyme  surface. 
However,  the  frequently  made  assumption  that  this  is  the  only  mechanism 
involved  is  often  unjustified.  Other  mechanisms  which  should  be  borne  in 
mind  will  be  mentioned  here;  they  will  be  illustrated  and  discussed  in  greater 
detail  later  in  the  chapter.  Let  us  first  consider  the  mechanisms  which  may 
apply  particularly  to  the  inhibition  of  pure  enzymes. 

(A)  Binding  of  the  analog  to  the  enzyme  sites  for  substrate,  coenzyme, 
or  activator  by  interactions  which  are  at  least  in  part  those  involved  in 
the  binding  of  the  normal  reactant  and  which  allow  reversibility. 

(B)  An  irreversible,  or  practically  irreversible,  reaction  with  the  enzyme 

*  The  definition  of  analog  must  be  imprecise  because  it  is  impossible  to  limit  accu- 
rately how  much  structural  deviation  can  occur  before  the  derivative  can  no  longer 
be  thought  of  as  related  to  the  parent  compound. 


POSSIBLE    SITES    AND    MECHANISMS    OF    INHIBITION  247 

site  subsequent  to  binding.  Inasmuch  as  substrates  occasionally  form  a 
temporary  chemical  bond  with  the  enzyme,  an  analog  may  do  likewise  but 
fail  to  complete  the  reaction,  remaining  chemically  attached  to  the  site. 

(C)  Binding  of  the  analog  to  an  enzyme  site  other  than  that  with  which 
the  normal  reactant  interacts.  Such  binding  may  be  simply  fortuitous  or 
the  site  may  be  specifically  for  the  purpose  of  allowing  feedback  inhibition 
by  a  product  formed  in  the  sequence  in  which  the  enzyme  participates. 
Regions  outside  the  catalytic  areas  with  which  inhibitors  can  react  are  of- 
ten called  allosteric  sites. 

(D)  The  analog  may  be  a  substrate  of  the  enzyme  and  will  inhibit  the 
reaction  of  the  normal  substrate  to  a  degree  dependent  on  the  relative 
binding  affinities  and  reaction  rates. 

{E)  Binding  of  the  analog  to  a  complex  of  the  enzyme  with  the  normal 
substrate,  coenzyme,  or  activator. 

(F)  The  formation  of  a  molecular  complex  of  the  analog  with  the  normal 
reactant  as  a  result  of  their  structural  complementarity.  Although  such 
complexes  are  probably  uncommon  and  have  seldom  been  considered  in 
work  with  analogs,  we  shall  see  that  examples  of  this  mechanism  are  known. 

(G)  Inhibition  by  a  mechanism  only  indirectly  related,  or  completely 
unrelated,  to  the  structural  similarity  of  the  analog  to  the  normal  substra,te. 
An  analog  may,  for  example,  possess  chelating  properties  not  exhibited 
by  the  substrate,  or  it  may  react  with  SH  or  carbonyl  groups. 

When  one  is  investigating  more  complex  systems,  particularly  cellular 
preparations,  a  number  of  other  mechanisms  for  analog  inhibition  may  be 
proposed,  and  these  should  be  added  to  the  above  list. 

(H)  The  analog  may  interfere  with  the  transport  mechanism  by  which  the 
normal  substance  is  taken  through  the  cell  membrane,  since  the  two  sub- 
stances may  both  combine  with  some  membrane  carrier  or  enzyme  system 
required  for  efficient  transport  or  accumulation. 

(/)  The  analog  may  not  be  the  actual  inhibitor,  but  may  be  transformed 
through  a  metabolic  sequence  into  a  substance  which  blocks  a  later  reaction, 
a  process  frequently  termed  lethal  synthesis.  In  certain  instances  the  analog 
may  complete  a  long  and  complex  metabolic  journey  to  terminate  as  a 
component  in  some  important  cellular  product.  The  incorporation  of  pyrim- 
idine  and  purine  analogs  (e.g.,  the  5-halouracils,  2-thiouracil,  8-azagua- 
nine,  and  8-azathymine  into  RNA  and  DNA)  and  amino  acid  analogs 
(e.g.,  tryptazan,  7-azatryptophan,  ethionine,  p-fluorophenylalanine,  and 
/5-2-thienylalanine  into  proteins)  has  been  frequently  demonstrated.  The 
products  containing  the  analogs  may  be  so  abnormal  as  to  fail  to  function 
properly  in  the  cells,  thereby  producing  far-reaching  and  complex  dis- 
turbances. 


248  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

(J)  The  analog  may  act  not  on  the  enzyme  attacking  the  normal  sub- 
strate but  on  an  enzyme  involved  in  the  formation  of  this  substrate,  since 
the  precursors  of  the  substrate  will  usually  be  structurally  similar  to  it. 
In  the  linear  sequence: 

E,         Ej 
X  -^S->  P 

an  analog  of  S  may  have  been  designed  to  inhibit  Eg  but  actually  in  cellular 
metabolism  acts  primarily  on  Ej  to  reduce  the  rate  of  formation  of  P. 


KINETICS  OF  ANALOG   INHIBITION 

The  kinetics  of  competitive  inhibition  have  been  presented  in  Chapter 
1-3,  and  the  graphical  analyses  for  the  proof  of  competition  and  the  deter- 
mination of  the  constants  discussed  in  Chapter  1-5.  Type  A  plots  of  l/v 
against  1/(S)  have  almost  invariably  been  used  to  demonstrate  the  types 
of  inhibition  produced  by  analogs,  but  other  types  of  plotting  may  be  more 
satisfactory  in  certain  situations,  especially  when  the  inhibition  is  not 
clearly  and  completely  competitive.  It  is  my  opinion  that  many  kinetic 
analyses  of  inhibition  would  be  improved  if  several  types  of  plotting  proce- 
dure were  used,  allowing  comparison  of  the  results  and  more  accurate  cal- 
culations of  the  constants. 

It  is  quite  often  the  case  that  the  inhibition  by  an  analog  is  not,  by  the 
usual  methods  of  analysis,  competitive  with  the  substrate  or  coenzyme 
to  which  the  analog  is  structurally  related,  and  such  results  have  puzzled 
many  workers.  The  plotting  may  indicate  noncompetitive,  coupling,  mixed, 
or  indeterminable  inhibition  mechanisms.  It  has  been  pointed  out  (Chap- 
ter 1-3)  that  true  noncompetitive  inhibition  must  be  rather  rare  and  this 
is  particularly  true  for  inhibition  by  analogs.  Coupling  or  uncompetitive 
inhibition  is  perhaps  more  common  among  analogs  than  with  other  in- 
hibitors, especially  with  regard  to  coenzymes  or  cofactors,  inasmuch  as 
the  substrate  combines  with  enzyme-coenzyme  or  enzyme-cofactor  com- 
plexes in  many  reactions.  It  is  important  to  determine  in  any  case  if  the 
inhibition  is  really  competitive  and  the  kinetics  modified  to  obscure  this, 
or  whether  the  mechanism  is  actually  other  than  competitive,  in  those 
instances  in  which  the  graphical  analysis  does  not  demonstrate  the  typical 
and  expected  competitive  picture. 

There  are  several  reasons  why  an  analog  inhibition  may  not  turn  out  to 
be  competitive  by  the  usual  plotting  procedures,  and  it  may  be  useful  to 
list  them  at  this  point. 

(A)  The  analog  may  be  acting  by  some  mechanism  other  than  specific 
attachment  to  an  active  site  on  the  enzyme,  that  is,  by  one  of  the  mecha- 
nisms listed  in  the  previous  section. 


KINETICS    OF    ANALOG    INHIBITION  249 

(B)  The  analog  may  be  bound  very  tightly  to  the  enzyme,  in  which  case 
the  order  of  addition  of  the  substrate  and  the  analog  may  be  of  great  im- 
portance. If  the  analog  is  added  to  the  enzyme  and  the  mixture  is  incubated 
before  the  substrate  is  added,  the  inhibition  may  be  very  marked  and  the 
substrate  may  be  unable  to  displace  the  inhibitor  from  the  enzyme  in  a 
reasonable  time.  On  the  other  hand,  if  both  substrate  and  analog  are  added 
together,  the  inhibition  may  be  very  low  initially  but  progress  slowly,  due 
to  the  relatively  small  fraction  of  the  active  centers  available  for  reaction 
with  the  inhibitor.  In  either  case  secondary  changes  in  the  enzyme  may 
occur  and  complicate  the  kinetics  (Chapter  1-12).  The  basic  problem  in 
the  interpretation  of  such  inhibitions  is  the  inability  experimentally  to 
achieve  satisfactory  equilibrium  conditions,  and  it  must  be  stressed  that 
the  usual  inhibition  equations  and  plotting  procedures  apply  only  to  equi- 
librium conditions.  There  has  been  confusion  between  noncompetitive  and 
irreversible  inhibitions.  The  arsenicals,  the  mercurials,  and  iodoacetate, 
for  example,  are  often  thought  to  inhibit  succinate  dehydrogenase  non- 
competitively,  and  yet  succinate  and  these  inhibitors  react  at  the  same  site 
on  the  enzyme,  as  shown  by  the  protection  afforded  by  the  presence  of 
succinate  when  it  is  added  with  the  inhibitors.  These  inhibitions  are  indeed 
competitive  under  certain  conditions  and  during  specific  time  intervals, 
but  once  the  inhibition  has  been  established  it  is  difficult  to  demonstrate 
a  competitive  effect.  Several  examples  of  this  situation  using  analog  inhi- 
bitors will  be  encountered  in  this  chapter. 

(C)  The  concentration  of  free  inhibitor  may  be  depleted  due  to  its  com- 
bination with  the  enzyme  or  other  materials  and  one  is  then  dealing  with 
a  mutual  depletion  system  (Chapter  1-3).  The  quantitative  aspects  of  com- 
petition may  be  modified  quite  markedly  in  such  cases.  This  behavior  must 
be  looked  for  particularly  when  one  is  working  with  very  potent  analog 
inhibitors. 

(D)  An  irreversible  or  semi-irreversible  change  in  the  configuration  or 
properties  of  the  active  center  may  occur  following  reaction  with  the  inhibi- 
tor, so  that  even  after  dissociation  of  the  inhibitor  from  the  enzyme  the 
affinity  of  the  enzyme  for  the  substrate  is  altered.  The  active  centers  of 
certain  enzymes  appear  to  be  flexible  and  adapt  in  some  way  to  the  in- 
teracting molecules,  and  it  is  possible  that  such  a  change  would  not  be  readily 
reversible.  The  active  center,  or  at  least  the  immediately  adjacent  region, 
of  the  penicillinase  of  Bacilltis  cerevs  is  altered  by  combination  with  the 
competitive  analog  of  benzylpenicillin,  in  that  there  is  a  marked  increase 
in  the  sensitivity  to  iodination,  and  this  is  prevented  by  the  presence  of 
substrate  (Citri  and  Garber,  1961).  Although  this  alteration  is  presumably 
reversible,  since  the  ability  to  hydrolyze  benzylpenicillin  after  removal  of 
the  inhibitor  is  unimpaired,  it  is  easy  to  imagine  changes  only  slowly  re- 
versible. 


250  2.  ANALOGS  OF  ENZYME  KE ACTION  COMPONENTS 

(E)  111  cellular  systems,  or  possibly  in  subcellular  preparations  of  some 
structural  complexity,  the  failure  of  the  analog  to  penetrate  to  the  site  of 
inhibition  as  readily  as  the  substrate  would  obviously  distort  the  kinetics, 
although  the  inhibition  itself  is  truly  competitive. 

{F)  If  the  substrate  possesses  several  groups  through  which  it  is  bound 
to  the  enzyme,  an  analog  which  has  only  one  of  these  groups  (perhaps  an 
analog  comprising  only  a  part  of  the  normal  substrate  molecule)  may  block 
off  one  enzyme  attachment  point  but  allow  the  substrate  to  bind  through 
the  remaining  groups.  In  many  cases  this  will  prevent  catalysis,  but  in 
others  it  could  only  reduce  the  rate  at  which  the  substrate  is  reacted. 
Simultaneously  there  will  be  a  reduction  in  the  affinity  of  the  enzyme  for 
the  substrate.  The  plots  will  give  evidence  of  a  mixed  inhibition,  which  is 
the  actual  situation,  but  nevertheless  competition  of  a  type  is  occurring. 

Analogs  of  coenzymes  often  present  a  special  problem  in  this  connection 
since  the  coenzyme  may  be  bound  quite  tightly  to  the  apoenzyme.  The 
addition  of  analog  to  the  complete  enzyme  may  not  result  in  significant 
displacement  of  the  coenzyme  and  little  inhibition  will  be  observed.  How- 
ever, if  the  enzyme  is  resolved  into  its  components  by  dissociating  the  coen- 
zyme in  some  manner,  competition  between  the  coenzyme  and  analog  may 
be  demonstrated  in  recombination  experiments  in  which  the  analog  reduces 
the  ability  of  added  coenzyme  to  reactivate  the  enzyme. 

The  type  of  inhibition  may  depend  on  the  concentration  of  the  analog. 
It  has  been  noted  several  times  that  an  inhibition  may  be  competitive  at 
low  analog  concentrations  and  partially  or  completely  noncompetitive  at 
higher  concentrations.  The  noncompetitive  elements  of  the  inhibition 
probably  reflect  the  increasing  unselectivity  of  action  that  is  a  common 
property  of  all  inhibitors  when  the  concentration  is  raised  beyond  a  cer- 
tain level. 

An  analog  of  a  substrate  will  occasionally  undergo  reaction  in  the  presence 
of  the  enzyme  and,  since  both  substrate  and  analog  bind  at  the  same  site 
on  the  enzyme,  competition  will  occur.  The  behavior  in  such  a  situation 
was  discussed  in  Chapter  1-3  (page  96).  The  inhibition  observed  will  depend 
on  what  is  determined.  If  we  designate  the  substrate  by  S  and  its  analog 
by  S': 

E  +  S    ^ES— >E  +  P  (2-1) 

E  +  S'  ;?t  ES'  ^  E  +  P'  (2-2) 

and  the  individual  rate  expressions  may  be  written  as: 

F„.(S) 
{S)+K,    1  +  Vv 

A, 


KINETICS    OF    ANALOG    INHIBITION  251 

F.'(S') 


(S')  +  K, 


1+^ 


(2-4) 


If  the  disappearance  of  S  or  the  formation  of  only  P  is  measured,  the  inhi- 
bition by  S'  will  be  typically  competitive,  but  the  inhibitor  constant  de- 
termined by  the  plotting  procedures,  K^,,  wiU  not  necessarily  be  a  true  dis- 
sociation constant.  The  result  wiU  depend  on  the  ratio  kg.jkg,  which  in 
most  cases  will  be  considerably  less  than  unity.  If  P  and  P'  are  identical 
and  determined  together,  the  inhibition  wiU  be  less  than  in  the  previous 
case.  It  is  important  in  work  with  analogs  to  establish  whether  they  are 
catalytically  reacted  and,  if  so,  to  plan  the  inhibition  experiments  accord- 
ingly. It  is  also  necessary  in  many  instances  to  determine  initial  rates,  since 
the  concentration  of  the  analog  may  be  reduced  significantly  because  of 
its  conversion  to  the  product. 

The  product,  or  one  of  the  products,  of  the  enzyme-catalyzed  reaction 
may  generally  be  considered  as  an  analog  of  the  substrate,  or  of  part  of 
the  substrate  molecule.  Inhibition  by  products  was  taken  up  in  Chapter 
1-4  (page  140)  and  it  was  pointed  out  there  that  the  inhibition  is  not 
necessarily  competitive,  since  the  product  can  react  with  the  enzyme  or 
other  components  of  the  reaction  in  a  variety  of  ways.  Several  instances 
of  product  inhibition  will  be  encountered  in  this  chapter  and  in  none  of 
these  is  the  inhibition  due  to  a  simple  reversal  of  the  forward  reaction, 
but  to  actual  combination  with  the  enzyme  at  or  near  the  active  center. 

A  somewhat  more  complex  situation,  in  which  the  analog  is  transformed 
into  a  product  that  inhibits  a  subsequent  reaction  of  the  substrate,  is  fairly 
common  and  warrants  some  discussion  although  a  complete  kinetic  analysis 
is  difficult.  The  simplest  system  may  be  represented  by: 

A  --^  B  A  C  (2-5) 

E.     t 

A'  — ^  B'  (2-6) 

The  substrate,  A,  and  its  analog,  A',  both  are  reacted  by  E^  and  the  analog 
product,  B',  inhibits  E,.  There  are  several  ways  in  which  the  rate  can  vary 
with  time  and  the  behavior  of  the  system  will  depend  on  many  factors. 
In  the  first  place,  the  presence  of  A'  may  slow  down  reaction  1  whereby 
A  is  transformed  to  B,  this  being  a  case  of  competing  substrates.  The  for- 
mation of  C  may  thus  be  initially  slowed  for  this  reason.  The  concentra- 
tion of  B'  will  progressively  rise  and  the  inhibition  on  Ej  increase.  However, 
the  concentration  of  A'  will  decrease  and  the  competition  w^th  A  be  reduced, 
leading  to  a  relative  acceleration  of  reaction  1.  The  change  in  the  rate  of 
formation  of  C  will  depend  on  the  balance  between  these  two  types  of 
inhibition.  For  example,  if  A'  is  reacted  fairly  rapidly  but  B'  is  not  a  very 


252  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

potent  inhibitor  of  Eg,  the  rate  may  first  rise  as  the  inhibition  on  E^  is 
released,  and  then  fall  later  when  the  concentration  of  B'  rises  sufficiently. 
On  the  other  hand,  if  A'  is  not  readily  depleted  and  B'  inhibits  well,  the  in- 
hibition will  steadily  increase.  It  is  clear  that  the  kinetics  may  not  indicate 
competitive  inhibition  when  the  rate  of  formation  of  C  is  measured.  The 
concentration  of  the  intermediate  B  may  rise  and  fall  in  a  complex  manner, 
as  discussed  in  Chapter  1-9  (page  438).  This  is  actually  the  simplest  case 
of  lethal  synthesis  complicated  by  competition  in  the  initial  reaction. 


MEANS  OF  EXPRESSING  RESULTS 

The  results  in  the  study  of  analogs  have  usually  been  given  in  terms  of 
inhibitions  at  different  concentrations  of  substrate  and  analog.*  Such 
results  are  often  valuable  but  rather  difficult  to  interpret,  especially  when 
different  analogs  at  different  concentrations  must  be  compared.  It  would 
be  more  valuable  if  inhibitor  constants  were  calculated  and  presented, 
along  with  Michaelis  or  substrate  constants,  and  fortunately  this  practice 
is  becoming  more  common.  With  values  of  K,„  and  K^  it  is  possible  to  de- 
termine the  inhibition  expected  at  any  combination  of  substrate  and  analog 
concentrations.  Furthermore,  it  is  possible  to  calculate  relative  interaction 
energies  from  a  series  of  K-q  obtained  for  a  group  of  analogs,  and  thereby 
put  the  inhibition  on  a  more  molecularly  interpretable  basis. 

There  are  fundamentally  two  types  of /i",.  The  actual  dissociation  constant 
for  the  EI  complex  may  be  called  the  true  inJiihitor  constant  and  refers  to 
a  particular  free  and  active  form  of  the  inhibitor.  The  experimentally 
determined  K^,  on  the  other  hand,  often  differs  from  the  true  Ki  and  may 
be  termed  the  apparent  inhibitor  constant.  This  apparent  K^  may  depend 
on  a  number  of  factors,  the  most  important  of  which  is  usually  the  pH 
since  many  inhibitors  are  weak  acids  or  bases.  This  problem  has  been 
discussed  in  some  detail  in  Chapter  1-14.  If  the  inhibitor  is  a  weak  acid 
only  one  form  may  be  the  inhibitor,  say  the  ionized  I~  form,  in  which  case 
the  apparent  K^  will  vary  with  pH  in  the  range  around  the  p^^  of  the 
inhibitor.  The  true  K^,  which  refers  to  the  equilibrium 

E  +  I-  ^  EI- 

will  not  vary  with  the  pH  because  of  changes  in  the  ionization  of  the  inhi- 
bitor, although  it  may  for  other  reasons.  The  apparent  K^,  in  fact,  will 
depend  on  any  type  of  equilibrium  between  active  and  inactive  forms  of 
the  inhibitor,  for  example  an  enol-keto  isomerism,  or  on  the  binding  of 
the  inhibitor  to  nonenzyme  components  of  the  preparation.  The  Kj's  given 

*  The  substrate  concentration  has  been  omitted  in  some  reports  and  this  vitiates 
the  results  and  makes  them  quantitatively  meaningless. 


MEANS   OF   EXPRESSING   RESULTS  253 

in  this  chapter  will  in  almost  all  cases  be  apparent  inhibitor  constants  be- 
cause the  true  K/s  have  generally  not  been  calculated  in  the  reports, 
although  in  many  instances  they  must  be  very  close  to  the  true  K/s  because 
of  the  nature  of  the  analog  or  the  conditions  of  the  experiments.  I  have 
taken  the  liberty  in  certain  cases  where  the  data  are  adequate  of  calculating 
the  K/s  from  the  inhibitions  reported.  In  every  instance  where  plotting 
procedures  have  been  used  to  determine  the  type  of  inhibition,  it  would 
have  been  possible  to  determine  the  appropriate  constants,  but  these  con- 
stants have  been  seldom  reported. 

An  especially  unsatisfactory  means  of  expressing  the  results  is  to  give  the 
inhibition  for  a  certain  ratio  of  inhibitor  to  substrate  concentrations, 
(I)/(S),  which  has  often  been  called  the  inhibition  index.  If  the  actual  con- 
centrations are  not  given,  it  is  not  possible  to  visualize  the  inhibition  quan- 
titatively, because  (I)/(S)  is  not  constant  for  a  certain  degree  of  inhibition 
even  when  the  inhibition  is  competitive.  This  has  been  clearly  pointed  out 
in  Chapter  1-3  (page  106)  but  it  is  important  to  re-emphasize  it  here.  The 
ratio,  (I)/(S),  is  constant  and  meaningful  only  at  high  substrate  concen- 
trations which  saturate  the  enzyme.  This  may  be  seen  from  the  following 
expression,  obtained  by  rearranging  the  equation  for  competitive  inhibition: 


1         1 

~k7  ^'Wi 


1 


(2-7) 


It  has  sometimes  been  assumed  tliat  for  50%  inhibition,  (I)/(S)  =  KJKg, 
but  this  is  not  necessarily  true,  which  can  be  easily  seen  by  rewriting  Eq. 
2-7  for  i  =  0.5  and  (S)  =  nK,: 

(I) 


l:[v  +  ^l 


,   -.  (2-8) 

(S)          Ks     '  ' 

(I)/(S)  =  KJK^  only  when  the  substrate  concentration  is  high  relative  to 
Kf.  (i.e.,  when  n  is  much  greater  than  unity).  This  can  also  be  seen  in  another 
way:  If  K,  =  1,  K^  =  0.3,  and  (I)/(S)  is  kept  constant  at  0.3,  the  inhibition 
will  vary  with  the  absolute  magnitudes  of  the  concentrations  as  shown  in 
the  following  tabulation: 


(S) 

(I) 

i 

0.1 

0.03 

0.083 

0.3 

0.09 

0.19 

0.5 

0.15 

0.25 

1 

0.3 

0.33 

2 

0.6 

0.40 

5 

1.5 

0.45 

10 

3 

0.48 

20 

6 

0.49 

254  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

It  is  obvious  that  a  statement  that  the  inhibition  was  a  certain  value  at 
(I)/(S)  =  0.3  would  be  of  little  significance. 

When  a  series  of  analogs  is  tested,  quantitative  expression  of  the  relative 
affinities  of  the  enzyme  for  the  various  analogs  is  desirable  when  possible. 
The  term  affinity  implies  no  units  and  has  a  vague  meaning,  having  been 
used  in  a  variety  of  ways.  A  common  method  of  expressing  the  affinity 
is  to  equate  it  tolIK^  in  the  case  of  an  inhibitor,  and  to  HKg  for  a  substrate. 
The  ratio  of  the  affinities  of  an  inhibitor  and  a  substrate  is  thus  often  ex- 
pressed as  KJKi.  If  K,  =  1  mM,  and  K,  =  0.001  mM,  it  would  be  stated 
that  the  affinity  of  the  enzyme  for  the  inhibitor  is  1000  times  that  for  the 
substrate.  This  may  sound  dramatic  but  is  misleading  in  a  way.  It  would 
seem  that  affinity  might  be  better  expressed  in  terms  of  binding  energies. 
The  ratio  of  the  free  energies  of  binding  of  inhibitor  and  substrate  is 
given  by: 


AF,         vKs 


(2-9) 


In  the  hypothetical  case  above,  AFJAF^  =  2,  and  the  inhibitor  is  bound 
twice  as  tightly  as  the  substrate,  which  is  a  more  reasonable  way  of  desig- 
nating the  relative  affinities. 

One  might  consider  three  types  of  AF  for  the  binding  of  an  inhibitor  to 
an  enzyme.  There  is  the  true  AF  corresponding  to  the  true  K^^  and  an  ap- 
parent AF  corresponding  to  the  apparent  experimental  K^.  In  addition 
there  is  the  theoretical  AF  corresponding  to  the  interaction  energy  of  the 
enzyme  and  inhibitor  in  a  vacuum,  uncomplicated  by  solvent  and  ions.  In 
order  to  compare  different  analogs  with  respect  to  their  affinities  for  the 
enzyme,  it  is  actually  this  last  AF  one  would  wish  in  most  cases,  but  it  is 
impossible  to  obtain.  Lacking  this,  one  must  use  the  true  AF  values,  but, 
as  pointed  out  above,  true  K/s  are  not  often  available.  It  may  be  quite 
misleading  to  compare  calculated  AF  values  for  a  series  of  analogs  if  these 
analogs  have  different  p-ff^'s  and  the  experimental  pH  is  in  the  region  of 
these  p-fi'./s.  The  differences  in  the  AF's  may  reflect  mainly  the  different 
degrees  of  ionization  rather  than  the  differences  in  binding  energies  of  the 
inhibitory  forms.  In  several  instances  in  this  chapter,  I  have  calculated  the 
AF  values  for  such  series  of  analogs  and  it  must  be  remembered  that  the 
validity  of  comparing  these  values  is  sometimes  questionable.  It  is  some- 
times impossible  to  calculate  even  an  apparent  AF  and  one  must  be  satisfied 
with  values  that  are  relative  to  some  chosen  compound;  these  will  be  called 
relative  AF  values.  Although  the  AF  itself  may  not  be  meaningful,  occa- 
sionally the  difference  of  the  AF  values  for  two  inhibitors  will  be  significant 
in  attributing  interaction  energies  to  certain  groups,  as  discussed  in  Chapter 
1-6  (page  268),  since  all  of  the  other  factors  involving  the  solvent  and  ionic 
atmosphere  may  remain  relatively  constant  for  the  inhibitors.  The  relative 
binding  energies  may  be  calculated  in  some  cases  even  though  the  K^'a 


IMPORTANT  TYPES  OF  MOLECULAR  ALTERATION  255 

are  not  known,  since  for  two  inhibitors: 

AF,  —  AF,  =  1.422  log  ['  il  ~  \'l  (2-10) 

ii  (1  —I2) 

if(Ij)  =  (l2)- 

It  is  easy  to  derive  an  expression  for  the  relationship  between  the  true 

AF  difference  and  the  apparent  JF  difference  for  two  inhibitors  when  the 

sole  factor  involved  is  the  ionization  of  the  inhibitors.  If  we  designate  the 

true  free  energy  of  binding  as  AF  and  the  apparent  free  energy  of  binding 

as  AF',  the  difference  in  the  true  binding  energies  for  the  two  inhibitors 

will  be  AFi  —  AF^,  and  the  difference  in  the  apparent  binding  energies 

will  be  AF^  —  AF^'.  The  relationship  between  these  for  analogs  that  are 

singly  ionizing  weak  acids  is: 

(AF,  -  AF,)  =  {AF,'  -  AF,')  -  1.422  log  \  +  [|h^)/^?]        ^^'^^^ 

It  is  therefore  possible  to  correct  for  the  pH  effect  if  the  p/C^'s  of  the  inhibi- 
tors are  known  and  the  true  interaction  energy  difference  may  be  obtained. 

IMPORTANT  TYPES  OF  MOLECULAR  ALTERATION 
PRODUCING  INHIBITING  ANALOGS 

A  substrate  or  coenzyme  might  generally  be  considered  to  have  three 
different  types  of  molecular  region:  (1)  groups  involved  primarily  in  the 
binding  to  the  enzyme,  (2)  groups  involved  in  the  catalytic  reaction,  and 
(3)  groups  or  regions  not  directly  involved  in  either  binding  or  reaction. 
There  is  overlap  between  these  in  some  cases,  of  course,  because  the  groups 
undergoing  chemical  change  usually  participate  to  a  certain  extent  in  the 
binding.  Furthermore,  some  substrates,  such  as  succinate,  do  not  possess 
the  third  type  of  group,  all  of  the  molecule  being  directly  involved  in  bind- 
ing and  reaction.  The  properties  of  a  substance  produced  by  altering  a  single 
group  or  region  of  a  substrate  wiU  depend  on  the  type  of  group  or  region 
that  is  modified,  that  is,  its  function  in  the  reaction  of  the  substrate  with 
the  enzyme.  A  change  in  a  binding  group  will  usually  alter  the  interaction 
energy  in  the  combination  of  the  substance  with  the  enzyme  and  may  or 
may  not  affect  the  susceptibility  to  chemical  reaction,  whereas  a  change  in 
a  group  directly  involved  in  the  catalysis  will  generally  reduce  the  reactivity 
without  necessarily  modifying  the  binding  to  the  enzyme.  It  is  also  evident 
that  a  change  in  the  third  type  of  neutral  group  will  not  be  so  likely  to 
alter  the  behavior  of  the  substrate,  unless  such  a  change  in  some  way 
secondarily  modifies  the  interactions  of  the  other  groups.  The  aim  in  the 
design  of  analogs  for  enzyme  inhibition  is  to  produce  a  compound  which 
will  bind  reasonably  tightly  to  the  enzyme  (preferably  more  tightly  than 


256  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

the  substrate),  but  which  is  resistant  to  chemical  reaction  in  its  complex 
with  the  enzyme.  It  would  appear  that  the  most  effective  inhibitors  would 
result  from  modifications  of  the  reactive  groups,  or  of  groups  adjacent 
to  the  reactive  region,  rather  than  changes  in  binding  groups.  Malonate 
illustrates  this  in  a  simple  way  because  here  the  — CH2CH2 —  group  of 
succinate  has  been  altered  to  the  nonoxidizable  — CHg —  group,  while 
the  binding  — COO"  groups  remain.  Amidation  of  the  — C00~  groups  of 
either  succinate  or  malonate,  thereby  eliminating  the  negative  charges, 
produces  a  substance  that  is  neither  a  substrate  nor  an  inhibitor  because 
the  affinity  for  the  enzyme  has  been  lost. 

The  total  interaction  energy  between  a  substrate  or  an  inhibitor  and  the 
enzyme  active  center  is  the  result  of  all  the  forces  of  attraction  and  repulsion 
summed  over  all  the  participating  groups.  Every  atom  or  group  of  a  sub- 
strate or  its  analog  contributes  to  some  degree  to  the  interaction  energy 
but,  practically,  the  binding  may  be  attributed  usually  to  two  or  three 
groups  that  serve  to  orient  the  molecules  on  the  enzyme  surface.  The  sub- 
traction, addition,  or  alteration  of  substrate  groups  may  change  the  binding 
energy  in  various  ways.  Modifying  a  region  vicinal  to  a  binding  group  may 
sterically  interfere  with  the  normal  approach  of  this  group  to  the  enzyme 
group  with  which  it  interacts,  or  it  may  by  inductive  or  resonance  effects 
alter  the  properties  of  the  binding  group,  as  discussed  in  Chapter  1-6 
(page  304).  It  is  worthwhile  emphasizing  again  that  a  change  in  a  particular 
region  of  a  molecule  may  produce  variations  in  the  electronic  configurations 
throughout  the  entire  molecule,  and  that  a  change  in  the  interaction  energy 
cannot  generally  be  attributed  solely  to  this  altered  region.  Furthermore, 
the  volume  and  configuration  of  a  substrate  and  its  analog  may  involve 
water  of  hydration,  so  that  a  group  change  can  secondarily  affect  the  bind- 
ing by  modifying  the  disposition  of  the  bound  water  molecules.  The  in- 
troduction of  so-called  neutral  groups,  such  as  hydrocarbon  chains,  can 
bring  about  an  increase  in  the  binding  energy  through  nonspecific  van  der 
Waals'  interactions,  providing  these  groups  do  not  interfere  sterically  with 
the  approach  of  the  important  binding  groups  to  the  enzyme  surface. 

Most  analogs  of  substrates  have  less  affinity  for  the  enzymes  than  do  the 
natural  substrates,  which  is  reasonable  in  view  of  the  enzyme  active  center 
conformation  to  the  substrate  configuration.  However,  occasionally  an 
analog  will  exhibit  a  much  tighter  binding  than  the  substrate,  the  extra 
binding  energy  being  more  than  could  be  attributable  simply  to  a  new  group 
introduced  into  the  molecule.  In  such  cases  it  is  likely  that  a  qualitative 
change  in  the  binding  is  involved.  A  substrate  frequently  forms  a  covalent 
bond  with  the  enzyme  during  the  catalytic  reaction,  and  normally  this 
constitutes  only  an  intermediate  state  in  the  sequence  of  changes.  Certain 
analogs  may  be  able  to  form  this  type  of  bond  but  are  unable  to  complete 
the  sequence,  so  that  the  analogs  remain  firmly  attached  to  the  enzyme. 


IMPORTANT  TYPES  OF  MOLECULAR  ALTERATION  257 

This  is  known  to  occur  in  the  case  of  diisopropylfluorophosphate  and  re- 
lated cholinesterase  inhibitors,  as  well  as  with  monoamine  oxidase  inhibi- 
tors such  as  iproniazid. 

An  analog  may  be  related  to  the  corresponding  substrate  in  one  of  two 
general  ways:  it  may  be  either  an  isomer  of  tne  substrate,  or  a  substance 
obtained  by  the  replacement  of  one  or  more  groups  on  the  substrate. 
An  isomeric  analog  may  be  a  geometric  isomer  (e.g.,  one  of  a  cis  and  trans 
pair),  optical  isomer,  or  any  stereoisomer  of  the  substrate.  It  might  be 
thought  that  such  analogs  would  often  be  specific  and  useful  inhibitors  but 
actually,  except  for  certain  optical  isomers  (see  page  268),  this  is  seldom 
the  case,  the  reason  being  that  the  configuration  of  the  analog  is  more 
important  than  a  simple  equivalence  of  all  the  atoms  and  groups.  A  sub- 
stitution analog  can  result  from  a  variety  of  molecular  changes  in  the  sub- 
strate. What  is  often  called  addition  of  groups  is  usually  only  a  substitution 
of  the  new  group  for  a  H  atom  (e.g.,  the  replacement  of  a  H  atom  with  a 
F  atom  or  a  CH3  group),  and  what  is  called  deletion  of  groups  is  usually 
a  substitution  of  a  H  atom  for  the  group  that  is  removed.  On  the  other  hand, 
an  important  type  of  analog  is  derived  by  the  substitution  of  one  functional 
group  with  another  group  (e.g.,  the  replacement  of  an  OH  group  with  a 
SH  group,  or  of  a  CH3  group  with  a  CI  atom).  Some  commonly  interchangea- 
ble groups  might  be  put  in  the  following  families: 

(a)  — NH2         —OH         — SH         — CH3         —CI         — F         — H 

(b)  —COO-        — SO3-        — ASO3H-        — PO3- 

(c)  — CONH2         — SO2NH2 

{d)      — S—        —0—        — NH—        — CH=CH—        — CH2— 

(e)       — Phenyl         — benzyl         — pyridyl         — pyrimidyl         — cyclohexyl 

A  good  deal  has  been  written  about  isosteric  and  isomorphic  groups  in  the 
production  of  analogs  (for  an  excellent  review  see  Schatz,  1960),  especially 
with  respect  to  the  development  of  new  drugs,  but  this  has  limited  bearing 
on  the  elaboration  of  enzyme  inhibitors.  The  replacement  of  substrate 
groups  with  isosteric  and  approximately  isomorphic  groups  usually  leads 
to  substances  that  are  also  substrates.  It  is  generally  necessary  to  alter 
the  proper  region  of  the  substrate  molecule  significantly  in  order  to  produce 
an  effective  inhibitor.  There  are  actually  at  the  present  time  no  general 
rules  for  the  most  efficient  procedures  to  be  used  for  the  modification  of 
substrates  to  produce  inhibitors.  Various  enzymes  exhibit  quite  different 
reaction  mechanisms  and  an  effective  transformation  of  the  substrate  in 
one  case  will  not  work  for  other  enzymes.  For  this  reason  the  most  important 
thing  to  establish  initially  is  the  nature  of  the  particular  enzyme  mechanism, 
if  an  attempt  is  to  be  made  to  design  analogs  rationally.  The  binding  groups, 
reactive  groups,  and  relatively  neutral  groups  in  the  substrate  must  be 


258  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

determined  so  that  modifications  in  the  structure  may  be  made  in  the  proper 
regions.  The  examjjles  discussed  in  this  chapter  will  clearly  demonstrate 
that  many  analog  inhibitors  are  not  isosteric  or  isomorphic  with  the  sub- 
strate and  that,  indeed,  many  of  the  most  useful  inhibitors  appear  to  dif- 
fer quite  markedly  from  the  substrate.  In  this  connection  it  is  necessary 
to  call  attention  to  the  danger  of  visualizing  molecules  on  the  basis  of  their 
classic  two-dimensional  formulas.  One  must  also  realize  that  usually  not 
all  of  a  substrate  molecule  is  involved  in  the  binding  and  reaction  with  the 
enzyme;  the  side  of  the  molecule  more  distant  from  the  enzyme  may  be 
relatively  less  important  than  the  rest  of  the  molecule  and  modifications 
on  this  side  may  give  the  appearance  of  producing  radically  different  sub- 
stances, whereas  from  the  standpoint  of  the  enzyme  surface  these  sub- 
stances may  be  very  similar  to  the  substrate.  Conversely,  just  because  two 
substances  look  alike  when  written  in  the  usual  structural  formulas  is  not 
enough  to  ensure  that  they  will  exhibit  comparable  interactions  with  the 
active  center  of  an  enzyme.  Ideally,  one  should  learn  to  conceive  enzyme 
reactions  on  an  electronic  and  molecular  level  and  to  visualize  analogs 
with  three-dimensional  molecular  imagination,  in  other  words,  to  approach 
the  problem  of  analog  design  from  the  point  of  view  of  a  rational  active 
center. 

Some  analogs  are  not  directly  inhibitory  but  are  metabolically  trans- 
formed into  inhibitory  substances  that  block  a  sequence  at  a  later  step. 
Such  analogs  are  often  very  interesting  and  useful,  being  in  many  cases 
specific  and  potent.  The  design  of  an  analog  to  be  an  inhibitor  precursor 
presents  a  slightly  different  problem  than  in  the  general  case.  If  the  analog 
is  to  enter  into  the  metabolic  sequence  it  must  be  a  substrate  of  the  initial 
enzymes  and,  hence,  structurally  similar  to  the  natural  substrate  in  the 
region  of  the  reactive  groups.  Isosteric  and  isomorphic  substitution  is  often 
useful  in  this  situation.  This  probably  accounts  for  the  popularity  of  fluo- 
rine as  a  replacement  for  a  H  atom  in  the  design  of  this  type  of  analog. 
Many  F-substituted  analogs  enter  into  metabolic  sequences  and  interfere 
at  a  more  distal  region,  for  example,  fluoroacetate,  5-fluorouracil,  2>-fluoro- 
phenylalanine,  and  6-deoxy-6-fluoro-D-glucose.  The  F  atom  is  the  smallest 
of  the  common  atoms  that  may  be  substituted  for  a  H  atom,  and  approaches 
the  H  atom  in  size  (van  der  Waals'  radii  for  H  and  F  atoms  being  1.2  and 
1.35  A,  respectively)  (see  Table  1-6-8).  The  F  atom  is  also  relatively  un- 
reactive  and  forms  a  stable  bond  to  carbon.  However,  it  is  strongly  electro- 
negative and  alters  the  electronic  configuration  in  comparison  to  the  parent 
compound.  The  dipole  moment  of  the  C — F  bond  will  be  quite  different 
from  the  C — H  bond,  this  altering  neighboring  bonds  as  well  as  inducing 
an  ability  to  form  hydrogen  bonds  with  the  enzyme.  Thus  the  F  analog 
not  only  may  be  much  like  the  substrate  in  over  all  size  and  configuration, 
allowing  it  to  be  metabolically  reactive,  but  eventually  may  be  transformed 


THE    CONCEPT    OF    INHIBITION    BY    ANALOGS  259 

into  a  substance  in  which  the  F  atom,  because  of  its  electronegative  char- 
acter, interferes  in  some  way  with  the  catalytic  reaction.  Another  consid- 
eration arises  when  phosphorylation  of  the  substrate  is  an  initial  step  in 
the  metabolic  sequence.  Here  one  must  be  careful  not  to  alter  the  groups 
involved  in  the  phosphorylation,  but  to  modify  groups  that  are  reactive  in 
a  later  enzymic  step. 

DEVELOPMENT  OF  THE  CONCEPT  OF  INHIBITION 
BY  ANALOGS 

This  important  concept,  which  today  plays  a  major  role  in  the  fields 
of  enzymology  and  biochemistry  and  is  becoming  more  and  more  important 
in  pharmacology,  chemotherapeutics,  and  pathology,  has  an  interesting 
history  illustrating  a  typical  growth  pattern  of  a  scientific  idea.  The  de- 
velopment of  the  general  concept  has  been  traced  by  Martin  (1951),  Wool- 
ley  (1952),  and  Albert  (1960),  so  that  here  it  is  necessary  only  to  present  a 
cursory  exposition  related  particularly  to  enzyme  inhibition.  However, 
it  must  be  realized  that  many  fields  of  study  —  including  immunology, 
drug  antagonism,  and  microbial  growth  inhibition  among  others  —  con- 
tributed in  one  way  or  another  to  this  concept. 

Despite  the  fact  that  numerous  examples  of  competitive  inhibition  by 
analogs  had  been  demonstrated  since  1910,  and  that  the  analog  concept 
had  been  quite  clearly  stated  around  1930,  general  recognition  of  the  basic 
principles  did  not  occur  until  after  1940.  Wohl  and  Glimm  (1910)  reported 
the  inhibition  of  amylase  by  glucose  and  galactose,  as  well  as  by  the  reac- 
tion product,  maltose,  and  shortly  Michaelis  described  the  inhibition  of 
/5-fructofuranosidase  by  fructose  and  a-methylglucoside  (Michaelis  and 
Pechstein,  1914),  and  the  inhibition  of  a-glucosidase  by  glucose  (Michaelis 
and  Rona,  1914).  Isolated,  unpremeditated,  and  unrecognized  discoveries, 
such  as  the  inhibition  of  arginase  by  ornithine  (Gross,  1921)  and  the  inhi- 
bition of /?-fructosidase  by  /^-glucose  but  not  by  cf-glucose  (Kuhn,  1923)  un- 
fortunately gave  no  impetus  to  the  formulation  of  a  general  concept.  Even 
the  classic  work  of  Quastel  on  malonate  inhibition,  described  in  the  previous 
chapter,  wherein  competitive  inhibition  by  structural  analogs  was  considered 
in  a  modern  fashion,  apparently  did  not  activate  anyone  outside  Cambridge, 
where  Bernheim  (1928)  studied  the  aconitate  inhibition  of  citrate  oxidation, 
Murray  (1929)  and  Murray  and  King  (1930)  applied  the  principles  of 
analogs  (although  in  a  somewhat  naive  way)  to  the  inhibition  of  lipase  by 
various  ketones  and  alcohols,  Richter  (1934)  proved  the  inhibition  of  ca- 
techol oxidase  by  resorcinol  to  be  competitive,  Keilin  and  Hartree  (1936) 
reported  potent  competitive  inhibition  of  uricase  by  the  methylurates,  and 
Green  (1936)  extended  the  malonate  inhibition  of  succinate  dehydrogenase 
to  the  inhibition  of  malate  dehydrogenase  by  several  dicarboxylates. 


260  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Several  accidental  observations  were  made  of  the  toxicity  of  metabolite 
analogs  while  they  were  being  tested  for  activity  in  animals,  for  example, 
the  discovery  that  ethionine  is  toxic  to  rats  (Dyer,  1938)  and  that  pyridine- 
3-sulfonate  is  lethal  to  dogs  suffering  from  nicotinic  acid  deficiency  (WooUey 
et  al.,  1938),  but  the  importance  was  not  immediately  recognized.  Appar- 
ently the  stimulus  for  the  formulation  of  the  general  theory  had  to  come 
from  a  discovery  of  clinical  importance,  and  this  was  provided  by  the  find- 
ing of  the  antagonism  by  p-aminobenzoate  of  the  action  of  the  sulfona- 
mides by  Woods  (1940).  Between  1940  and  1943  many  examples  of  analog 
antagonism  were  reported,  and  from  1943  to  1946  it  was  shown  that  vi- 
tamin deficiency  symptoms  could  be  produced  in  animals  by  the  administra- 
tion of  the  appropriate  analogs  of  all  the  known  vitamins.  It  was  then 
possible  to  return  to  the  enzyme  level  and  apply  the  principles  formulated 
long  before  to  the  new  results.  Since  1946  there  has  been  a  steadily  increas- 
ing interest  in  analog  inhibitors,  which  is  reflected  in  the  large  numbers  of 
publications  on  this  subject  each  year  (Fig.  2-1).  At  the  present  time  a  paper 
on  analog  enzyme  inhibition  as  covered  in  this  chapter  appears  every  other 
day;  this  does  not  include  all  of  the  investigations  on  analog  inhibitors 
that  are  to  be  treated  separately  (such  as  fluoroacetate,  arsenate,  carbon 
monoxide,  parapyruvate,  and  others),  or  analogs  commonly  used  clinically 


Fig.  2-1.  Curve  showing  the   annual   number  of  publications  on  analog  inhibition 

of  enzymes. 


ANALOG   INHIBITION    OF    MEMBRANE    TRANSPORT  261 

(sucli  as  the  cholinesterase  and  monoamine  oxidase  inhibitors),  or  antime- 
tabolites used  in  microbial  growth  suppression  or  tumoristasis. 

A  few  general  references  treating  aspects  of  the  subject  not  included 
in  the  present  work  are  suggested  for  additional  information:  Welch  (1945), 
Work  and  Work  (1948),  Woolley  (1950  b,  1952),  Martin  (1951),  Rhoads 
(1955),  Matthews  (1958),  Albert  (1960),  Schueler  (1960),  Schatz  (1960), 
Kaiser  (1960),  and  the  Symposium  on  Antimetabolites  sponsored  by  the 
National  Vitamin  Foundation  (1955).  A  great  deal  of  information  on  the 
biological  actions  of  many  types  of  analog  will  be  found  in  Volume  I  of 
"Metabolic  Inhibitors"  edited  by  Hochster  and  Quastel  (1963),  and  this 
aspect  of  the  subject  will  be  mainly  omitted  in  the  present  work  so  that 
the  enzymic  effects  may  be  discussed  in  sufficient  detail. 


ANALOG   INHIBITION   OF  MEMBRANE  TRANSPORT 

Before  discussing  specific  enzyme  inhibitions  by  analogs,  we  must  turn 
our  attention  to  the  possibility  that  the  depression  of  the  utilization  of 
some  substrate  or  metabolite  by  an  analog  is  not  due  to  an  action  on  any 
enzyme  involved  in  the  metabolism,  but  is  the  result  of  a  specific  inter- 
ference with  the  transport  of  the  substrate  or  metabolite  into  the  cell. 
It  is  quite  probable  that  some  of  the  actions  of  analogs  on  metabolism,  at- 
tributed to  competition  at  the  enzyme  level,  are  actually  exerted  at  the 
cell  membrane;  indeed,  certain  instances  will  be  discussed.  It  is  becoming 
more  and  more  evident  that  many  substrates  and  metabolic  precursors 
are  taken  up  by  cells  by  processes  other  than  simple  diffusion,  and  this 
applies  particularly  to  certain  carbohydrates,  amino  acids,  coenzymes, 
or  coenzyme  precursors.  It  is  not  necessary  that  the  transport  be  active 
to  be  influenced  by  analogs;  any  movement  of  a  substance  through  a  mem- 
brane which  involves  a  carrier,  a  special  size  or  configuration  of  pores,  or  a 
specific  type  of  mechanism,  active  or  passive,  can  be  slowed  in  the  presence 
of  an  analog  of  this  substance.  It  is  possible,  o^  course,  that  the  membrane 
trasport  is  mediated  through  an  enzyme  reaction,  such  as  a  phosphoryla- 
tion, and  the  competition  by  the  analog  is  then  truly  an  enzymatic  one. 

It  is  sometimes  difficult  to  determine  if  there  is  inhibition  of  a  transport 
process.  The  demonstration  of  a  reduced  uptake  of  a  substrate  from  the 
medium  is  generally  not  sufficient  evidence,  inasmuch  as  a  decreased  in- 
tracellular utilization  might  also  be  responsible.  The  best  procedure  is  to 
determine  directly  the  concentration  of  the  substrate  within  the  cells  in 
the  absence  and  presence  of  the  analog,  but  sometimes  the  concentration 
is  too  low  to  measure  accurately,  especially  when  the  substrate  is  rapidly 
metabolized.  The  demonstration  of  typical  competitive  inhibition  with 
respect  to  the  action  of  an  analog  on  some  metabolic  process  is  also  not 
sufficient  evidence  for  an  enzymic  site  of  action,  because  the  inhibition  of 


262  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

membrane  transport  may  exhibit  competitive  kinetics.  The  kinetics  will 
often  depend  on  whether  the  transport  is  limiting  the  metabolic  utilization 
of  the  substrate  or  not.  Membrane  transport  involving  a  carrier  (C)  molecule 
can  often  be  represented  by: 

S  +  C  ^  SCo  ->  SC,  -►  C  +  S  (2-12) 

where  the  subscripts  refer  to  the  outside  and  inside  of  the  membrane.  The 
later  reactions  may  be  written  in  one  direction  only  because  of  the  utili- 
zation of  the  substrate  as  it  enters  the  cell.  If  an  analog  also  combines 
with  the  carrier: 

S'  +  C  ^  S'C  (2-13) 

whether  it  is  transported  into  the  cell  or  not,  typical  competitive  behavior 
will  be  observed  since  the  forms  of  Eqs.  2-12  and  2-13  are  the  same  as 
those  for  competitive  enzyme  inhibition. 

Plots  of  transport  rates  against  the  external  concentrations  of  the  trans- 
ported substance  usually  yield  hyperbolic  curves,  and  double  reciprocal 
plots  are  often  linear,  allowing  the  calculation  of  a  constant  which  corres- 
ponds to  the  Michaelis-Menten  constant  in  enzyme  kinetics.  It  is  frequently 
assumed  that  this  is  the  dissociation  constant  for  the  complex  of  the  sub- 
stance with  the  carrier,  but  it  is  not  necessarily  true  for  the  same  reasons 
that  K^,^  is  not  always  K,.  The  kinetics  of  transport  inhibition  are  likewise 
commonly  similar  to  those  observed  with  enzymes  and  values  of  K^  may 
be  determined  by  appropriate  plotting,  this  constant  representing  the  dis- 
sociation constant  of  the  carrier-analog  complex.  The  kinetics  of  carrier 
transport  and  its  inhibition  have  been  elaborated  by  Wilbrandt  and  Rosen- 
berg (1961)  and  Rosenberg  and  Wilbrandt  (1962)  for  facilitated  diffusion 
and  certain  restricted  types  of  active  transport.  It  is  interesting  in  connec- 
tion with  certain  types  of  inhibition  work  to  note  that  the  accumulation 
ratio,  (X),/(X)^  =  VJkK„j,  where  F^„  is  the  maximal  transport  rate,  K,„ 
is  the  Michaelis-Menten  constant  for  transport,  and  k  is  the  passive  diffu- 
sion constant  for  the  membrane.  Thus  the  cell/medium  ratio  may  be  altered 
by  the  inhibitor  as  a  result  of  changes  in  any  of  these  three  parameters. 

Carbohydrate  Transport 

Competition  between  sugars  for  entrance  into  cells  has  been  observed 
in  many  tissues  but  has  seldom  been  studied  quantitatively,  so  that  in  most 
cases  it  is  impossible  to  know  if  true  competition  kinetics  are  followed. 
Occasionally  a  reduction  in  the  inhibition  with  increase  in  the  concentration 
of  the  transported  substrate  has  been  noted;  for  example,  the  active  ac- 
cumulation of  D-galactose  by  rabbit  kidney  cortex  slices  is  inhibited  61% 
by  5.6  rciM  glucose  when  D-galactose  is  0.1  mM  and  only  28%  when  d- 
galactose  is  0.2  mM  (Krane  and  Crane,  1959),  but  these  results  do  not  fit 


ANALOG   INHIBITION    OF    MEMBRANE    TRANSPORT  263 

a  simple  competitive  formulation.  Indeed,  the  exact  site  of  inhibition  has 
not  been  established  in  any  case.  The  entrance  of  L-arabinose  into  rat 
heart  cells  is  inhibited  92%  by  glucose  at  equimolar  concentration  and, 
since  L-arabinose  is  not  metabolized,  the  inhibition  is  presumably  on  some 
phase  of  membrane  transport  (Morgan  and  Park,  1958).  In  the  same  tissue 
the  competition  between  glucose  and  3-methylglucose  at  the  outer  but 
not  the  inner  surface  of  the  cell  membrane  leads  to  a  net  outward  transport 
of  3-methylglucose  against  a  concentration  gradient,  and  it  was  concluded 
that  there  are  stereospecific  combining  sites  at  both  surfaces.  A  compari- 
son between  cellular  and  subcellular  preparations  can  occasionally  indicate 
a  membrane  site  for  an  inhibition.  Galactose  markedly  inhibits  the  utili- 
zation of  fructose  by  intact  ascites  tumor  cells  but  not  in  homogenates, 
pointing  to  a  competition  before  the  hexokinase  step  and  probably  in 
transport  (Nirenberg  and  Hogg,  1957).  The  effects  of  2-deoxy-D-glucose 
on  the  fermentation  of  glucose,  fructose,  and  mannose  by  yeast  (to  be 
discussed  in  more  detail  on  page  391)  led  Scharff  (1961)  to  assume  a  trans- 
port system  which  is  stable,  since  it  still  functions  in  acetone-dried  cells, 
and  perhaps  bound  to  a  complex  or  "bundle"  of  fermentation  enzymes 
somewhere  in  the  outer  regions  of  the  yeast  cells. 

Although  the  nature  of  the  transport  mechanisms  and  the  site  of  inhi- 
bition are  generally  not  known,  it  is  clear  that  the  interference  is  quite 
specific  and  dependent  on  the  molecular  configurations  of  the  sugars.  The 
transport  of  D-galactose  (5  mM)  by  hamster  jejunum  is  inhibited  to  varying 
degrees  by  other  sugars  and  derivatives  at  25  nxM  (see  tabulation)  (Wilson 


Sugar  %  Inhibition 

D-Mannose  Stim  4 

D- Xylose  7 

3-0-Methyl-D-glucose  59 

a-Methyl-D-glucoside  95 

et  al,  1960),  and  the  uptake  of  glucose  by  lymph  node  cells  is  likewise 
inhibited  differently  by  several  sugars  at  9  mM  (see  tabulation)  (Helm- 
reich  and  Eisen,  1959),  both  observations  pointing  to  stereospecific  trans- 


Sugar  %  Inhibition 

D-Arabinose  7 

D-Galactose  9 

D-Fructose  40 

D-Mannose  61 


264  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

port  systems  presumably  involving  carriers  in  the  membranes.  The  pat- 
terns of  competition  between  various  sugars  have  often  led  to  the  assumption 
of  two  or  more  different  transport  mechanisms  for  carbohydrates  and  that 
these  systems  can  operate  simultaneously  and  independently.  Competition 
for  transport  can  be  demonstrated  in  rat  diaphragm  muscle  for  members 
of  the  group  including  D-glucose,  D-mannose,  D-xylose,  D-arabinose,  l- 
arabinose,  D-lyxose,  and  3-0-methyl-D-glucose,  but  not  between  these 
and  D-galactose,  D-fructose,  maltose,  a-methyl-D-glucoside,  and  /?-methyl- 
D-glucoside  (Battaglia  and  Randle,  1960),  so  that  different  sites  for  entry 
were  postulated.  A  number  of  sugars  are  transported  into  erythrocytes: 
all  the  aldoses  penetrate  by  a  transport  system  characterized  for  glucose, 
while  the  ketoses  penetrate  according  to  a  pattern  of  passive  diffusion 
(LeFevre  and  Davies,  1951).  The  aldoses  compete  with  each  other,  e.g., 
the  uptake  of  glucose  is  strongly  inhibited  by  mannose,  but  any  aldose 
delays  the  entrance  of  a  ketose,  e.g.,  the  uptake  of  fructose  is  prevented 
by  glucose  and  galactose,  while  the  ketoses  do  not  perceptibly  alter  the 
transport  of  the  aldoses.  One  carrier  system  in  the  erythrocyte  has  been 
characterized  as  reacting  only  with  those  monosaccharides  in  which  the 
pyranose  ring  tends  to  assume  the  "chair"  configuration,  which  illustrates 
a  unique  type  of  stereospecificity  (LeFevre  and  Marshall,  1958).  On  the 
other  hand,  in  hamster  intestine  only  a  single  transport  mechanism  or  car- 
rier seems  to  be  present,  mutual  inhibition  occurring  between  D-glucose 
and  D-galactose,  D-glucose  and  1,5-anhydro-D-glucitol,  D-galactose  and  1,5- 
anhydro-D-glucitol,  D-glucose  and  6-deoxy-D-glucose,  and  6-deoxy-D-glucose 
and  1,5-anhydro-D-glucitol,  whereas  sugars  that  are  not  transported  do  not 
interfere  with  those  that  are  (R.  K.  Crane,  1960). 

Amino  Acid  Transport 

The  situation  here  is  very  much  the  same  as  in  sugar  transport  and  there 
is  good  evidence  for  stereospecific  systems  and  multiple  pathways.  There 
is  competition  between  L-leucine  and  DL-isoleucine  for  renal  tubular  resorp- 
tion in  the  dog,  these  being  well  resorbed  amino  acids,  and  there  is  also 
competition  between  the  poorly  resorbed  L-arginine  and  L-lysine  (Beyer 
et  at.,  1947).  However,  no  interference  is  observed  between  L-leucine  and 
L-arginine.  It  is  possible  to  classify  the  amino  acids  into  groups  with  respect 
to  their  mutual  interference  in  resorption.  In  the  rat  kidney,  dibasic  amino 
acids  (arginine,  lysine,  and  cystine)  are  actively  accumulated  and  there 
is  mutual  inhibition  of  the  transport  (Rosenberg  et  al.,  1962).  Monobasic 
amino  acids  (alanine,  phenylalanine,  and  histidine)  do  not  interfere  with 
the  uptake  of  the  dibasic  amino  acids,  nor  does  arginine  depress  the  uptake 
of  the  monobasic  amino  acids.  It  seems  likely  that  separate  transport  sys- 
tems are  present.  The  transport  system  for  basic  amino  acids  in  the  hamster 
intestine  is  distinct  from  that  for  other  amino  acids  and  similar  to  the  renal 


ANALOG   INHIBITION  OF   MEMBRANE    TRANSPORT 


265 


system  (see  accompanying  tabulation)  (Hagihira  et  al.,  1961).  Arginine 
and  cystine  interfere  more  with  lysine  transport  than  with  glycine  trans- 
port, whereas  methionine  behaves  in  the  reverse  manner,   and  there  is 

/-I  J.    A.-  %  Inhibition  of  transport  of : 

,  ,  .,  .  Concentration 


inniDitor 

{raM) 

Glycine 

L-Lysine 

L-Arginine 

2 

16 

89 

L-Cystine 

0.8 

0 

45 

L-Methionine 

1 

73 

32 

L-Lysine 

1 

14 

— 

Glycine 

1 

— 

0 

little  or  no  interference  between  glycine  and  lysine.  Proline,  histidine,  and 
glycine  are  actively  transported  across  the  intestinal  wall  and  methionine 
at  eqiiimolar  concentration  completely  inhibits  this  (Wiseman,  1954).  This 
intestinal  transport  carrier  is  limited  to  monoamine-monocarboxylates  and 
they  compete  with  each  other.  In  addition  to  a  common  carrier,  there  may 
be  an  additional  carrier  for  glycine  and  proline  (Newey  and  Smyth,  1964), 
and  it  was  pointed  out  that  although  each  carrier  system  would  conform 
to  Michaelis-Menten  kinetics,  the  total  transport  with  two  or  more  carriers 
involved  would  not  necessarily. 

Intestinal  transport  systems  may  react  with  only  the  l-  or  the  D-form 
of  an  optically  isomeric  pair.  D-Methionine  is  accumulated  and  transported 
against  a  concentration  gradient  by  the  rat  intestine  and  this  is  blocked 
completely  by  equimolar  concentrations  of  L-methionine  ( Jervis  and  Smyth, 
1960).  Similarly,  the  transport  of  L-I^^^-monoiodotyrosine,  which  involves 
active  accumulation  of  the  amino  acid  in  the  gut  wall,  is  inhibited  by 
many  L-amino  acids  (the  most  effective  being  L-tryptophan,  L-methionine, 
L-leucine,  and  L-isoleucine)  but  scarcely  at  all  by  any  of  the  four  D-amino 
acids  tested  (Nathans  et  al.,  1960).  L-Tryptophan,  for  example,  at  10  mM 
reduces  the  tissue/medium  ratio  from  8.85  to  1.55  and  the  inhibition  is 
apparently  competitive. 

The  active  cumulative  uptake  of  amino  acids  by  ascites  carcinoma  cells 
comprises  several  transport  systems,  each  with  a  specific  range  of  substrates. 
The  uptake  of  glycine-1-C^*  {K,,,  =  6.4  mM)  is  most  strongly  inhibited  by 
1-aminocyclopentanecarboxylate  {K,  =  1.47  mM),  and  this  is  competitive, 
while  the  uptake  of  DL-methionine-S^^  {K„,  =  1.7  mM)  is  inhibited  best 
by  allylglycine  {K^  =  0.86  mM).  Glycine  transport  is  moderately  inhib- 
ited by  aUylglycine  and  less  readily  by  furylglycine  and  thienylglycine 
(Scholefield,  1961).  On  the  other  hand,  the  uptake  of  DL-leucine-1-C^^ 
is  stimulated  by  most  of  these  inhibitors.  Transport  of  L-tryptophan  in 


266  2.  ANALOGS  OF  ENZYME  EEACTION  COMPONENTS 

ascites  cells  occurs  by  both  diffusion  and  active  transport  (Jacquez,  1961). 
Certain  amino  acids  accelerate  this  at  lower  concentrations  (1  mM)  and 
competitively  inhibit  at  higher  (5  mM),  while  other  amino  acids  (such 
as  L-alanine,  L-lysine,  and  L-arginine)  only  inhibit.  Oxender  and  Christen- 
sen  (1963)  thoroughly  studied  the  effects  of  many  amino  acids  on  the  up- 
take of  neutral  amino  acids  by  ascites  cells  and  found  they  fall  into  two 
overlapping  clusters,  the  transport  systems  apparently  not  being  very 
specific. 

The  penetration  of  amino  acids  into  the  brain  is  probably  important 
for  the  metabolism  and  function  of  that  tissue,  and  there  appear  to  be 
several  transport  systems  available.  Tyrosine  enters  the  brain  readily  in 
vivo  and  a  specific  transport  is  probably  involved,  since  L-tyrosine  pene- 
trates more  rapidly  than  D-tyrosine  and  the  entry  is  potently  inhibited  by 
certain  other  amino  acids,  particularly  L-tryptophan,  L-leucine,  L-valine, 
/5-fluorophenylalanine,  and  L-histidine  (Chirigos  et  al.,  1960).  In  phenyl- 
ketonuria the  blood  levels  of  phenylalanine  are  high  due  to  the  inability 
of  the  tissues  to  metabolize  it  to  tyrosine.  It  is  possible  that  these  high 
concentrations  can  interfere  with  the  entry  of  other  amino  acids  into  the 
brain  and  partially  account  for  the  central  nervous  system  disturbances. 
The  uptake  of  five  amino  acids  by  rat  brain  slices  is  indeed  inhibited  by 
L-phenylalanine  (see  accompanying  tabulation)  and  it  was  felt  that  such 

Amino  acid  (2  raM)         %  Decrease  of  concentration  gradient 


L-Proline  9 

L-Histidine  42 

L-Arginine  46 

L-Ornithine  47 

L-Tvrosine  70 


could  occur  in  vivo  (Neame,  1961).  The  transport  of  L-histidine  is  inhib- 
ited by  neutral  aliphatic  amino  acids  and  short-chain  diamino  acids  to  a 
degree  dependent  on  the  length  of  the  carbon  chain  (Neame,  1964).  Inhi- 
bition by  the  dicarboxylic  amino  acids  is  not  dependent  on  the  chain  length. 
In  general,  the  L-isomers  inhibit  more  potently  than  the  D-isomers.  It  was 
suggested  that  histidine  is  transported  by  a  system  which  transports  most 
other  amino  acids,  but  with  different  affinities,  since  the  inhibitions  are  all 
competitive.  The  synthetic  amino  acid,  1-aminocyclopentanecarboxylate, 
is  not  metabolized  but  is  actively  transported  in  brain  slices  and  ascites 
cells,  and  the  system  involved  must  be  the  same  as  for  methionine  since 
it  is  affected  similarly  by  the  same  competitive  amino  acids  (Ahmed  and 
Scholefield,  1962).  Such  transported  but  nonmetabolized  amino  acids  may 
well  be  of  use  in  studying  transport  inhibition,  since  effects  can  be  clearly 


ANALOG    INHIBITION  OF  MEMBRANE    TRANSPORT  267 

distinguished  from  possible  inhibitions  of  amino  acid  incorporation  in  the 
cell.  Reference  should  be  made  to  the  very  complete  investigation  of  amino 
acid  transport  in  brain  slices  by  Abadom  and  Scholefield  (1962),  in  which  the 
many  competitive  inhibitions  established  point  to  several  separate  amino 
acid  transport  systems. 

The  entrance  of  valine,  proline,  and  hydroxyproline  into  the  human 
erythrocyte  is  not  mutually  competitive,  but  is  inhibited  markedly  by 
certain  sugars,  such  as  glucose,  galactose,  and  xylose,  although  not  by 
fructose  (Rieser,  1961).  These  results  would  indicate  that  some  amino  acids 
and  sugars  follow  the  same  transport  pathway.  If  this  is  a  general  phenome- 
non, one  must  consider  in  the  use  of  analogs  of  these  substances  the  possi- 
bility that  an  amino  acid  analog  might  depress  glucose  uptake,  and  thereby 
secondarily  interfere  with  the  transport  by  suppressing  energy  generation. 

The  accumulation  of  L-histidine  by  the  parasitic  fungus  Botnjtis  fabae 
is  inhibited  by  most  other  amino  acids,  and  one  transport  system  seems  to 
be  available  to  all  the  amino  acids  (Jones,  1963).  Substitution  at  the  NH2 
or  COOH  groups  lessens  or  abolishes  the  inhibitory  activity,  indicating 
that  the  binding  to  the  carrier  is  at  least  partly  electrostatic. 

Miscellaneous  Transport   Systems 

The  oxidation  of  protocatechuate  by  a  Flavohacterium  is  competitively 
inhibited  by  ^-aminosalicylate  and  one  might  conclude  that  mutual  in- 
teraction with  some  enzyme  is  responsible.  However,  59-aminosalicylate 
does  not  affect  the  rate  or  extent  of  the  oxidation  in  extracts,  measured  in 
different  ways  (Hubbard  and  Durham,  1961).  These  results  thus  point  to 
competition  for  a  transport  system  in  the  membrane,  rather  than  the  more 
usual  explanation.  The  active  transport  of  biotin  across  the  hamster  in- 
testine is  inhibited  by  various  analogs  (e.g.,  biocytin,  desthiobiotin,  di- 
aminobiotin,  and  biotin  methyl  ester),  but  the  nature  of  the  inhibition  was 
not  investigated  so  it  may  not  be  competitive  (it  is  not  with  lipoate)  (Spen- 
cer and  Brody,  1964).  The  active  influx  of  urate  into  erythrocytes  is  com- 
petitively inhibited  by  hypoxanthine  with  K^  =  0.1  mM,  whereas  the 
efilux  consists  of  two  components,  one  sensitive  to  hypoxanthine  (Lassen 
and  Overgaard-Hansen,  1962). 

There  are  several  instances  in  which  the  transport  of  inorganic  ions  is 
inhibited  by  other  related  ions.  The  transfer  and  exchange  of  phosphate 
across  the  membrane  of  S.  aureus  are  inhibited  by  chlorate,  borate,  and  ar- 
senate (Mitchell,  1954),  although  only  arsenate  is  able  to  substitute  com- 
pletely for  phosphate  in  the  exchange.  The  uptake  of  sulfate  by  yeast  is 
competitively  inhibited  by  thiosulfate  (Kleinzeller  et  al.,  1959).  Nitrate 
inhibits  quite  well  the  accumulation  of  iodide  in  the  rabbit  ciliary  body, 
50%  reduction  of  the  tissue/medium  ratio  occurring  at  3  vaM,  but  it  is 
not  known  if  this  is  truly  competitive  (Becker,  1961). 


268  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

The  examples  chosen  to  ilhistrate  transport  inhibition  do  not  always 
involve  analogs,  except  in  the  most  general  sense,  but  clearly  demonstrate 
the  importance  of  considering  such  a  type  of  interference  whenever  analogs 
are  used  in  cellular  preparations. 

ANALOGS  WHICH  ARE  ISOMERS  OF  SUBSTRATES 

The  behavior  of  the  isomers  of  normal  metabolites,  particularly  optical 
isomers,  constitutes,  in  a  way,  a  special  field  and  therefore  some  of  the  more 
interesting  results  will  be  discussed  in  this  section  rather  than  under  the 
specific  enzymes  that  are  involved.  The  concept  that  a  proper  fit  of  a  sub- 
strate to  the  enzyme  surface  is  necessary  for  reaction  implies  that  enzymes 
will  usually  be  stereospecific,  and  this  has  been  demonstrated  many  times. 
Since  an  enzyme  commonly  attacks  only  one  form  of  an  isomeric  pair, 
the  unreactive  form  may  be  either  an  inhibitor  or  completely  inert.  In  most 
cases  the  unreactive  form  does  not  bind  to  the  enzyme  at  all,  as  might  be 
expected  from  the  different  spatial  configurations  of  most  isomeric  pairs, 
and  is  not  inhibitory. 

Enantiomeric  Analogs 

These  analogs  are  related  to  the  corresponding  substrates  on  the  basis 
of  molecular  asymmetry  and  the  most  common  examples  are  optically 
active  due  to  an  asymmetric  carbon  atom.  Unnatural  enantiomers  often 
exhibit  no  affinity  for  the  enzymes.  Indeed,  it  has  been  generally  found 
that  D-amino  acids  do  not  interfere  with  the  microbial  growth-promoting 
activity  of  L-amino  acids,  although  there  are  exceptions.  Some  examples 
of  a  lack  of  inhibition  by  optical  isomers  may  be  mentioned.  The  oxidation 
of  L-phenylalanine  by  the  L-amino  acid  oxidase  of  Neurospora  is  not  in- 
hibited by  D-phenylalanine,  even  when  the  latter  is  present  at  500  times 
the  concentration  of  the  substrate  (Burton,  1951  b),  and  D-tryptophan 
does  not  inhibit  E.  coli  L-tryptophanase  even  though  it  interferes  with 
growth  (Gooder  and  Happold,  1954).  D-Malate  is  not  oxidized  by  the  malate 
dehydrogenase  of  Mycobacterium  tuberculosis  nor  does  it  inhibit  the  oxida- 
tion of  L-malate  (Goldman,  1956  b).  A  rather  unusual  case  is  presented  by 
potato  tyrosinase  in  that  both  l-  and  D-tyrosine  are  attacked  at  the  same 
rate,  but  there  is  no  evidence  of  mutual  inhibition,  perhaps  because  of  the 
limited  range  of  concentrations  used  (Spencer  et  al.,  1956).  Occasionally 
a  slight  inhibition  is  noted  but  one  which  would  not  be  of  any  practical 
significance,  as  in  the  just  detectable  inhibition  by  D-leucine  of  the  oxida- 
tion of  L-leucine  by  the  L-amino  acid  oxidase  of  the  hepatopancreas  of 
Cardium  tuberculatum,  an  inhibition  actually  much  less  than  exerted  by 
other  amino  acids  and  hence  probably  not  specific  (Roche  et  al.,  1959). 


ANALOGS  WHICH  ARE  ISOMERS  OF  SUBSTRATES  269 

However,  the  primary  purpose  of  this  section  is  to  discuss  instances  in 
which  significant  inhibition  by  optical  isomers  is  observed. 

The  asparaginase  of  Mycobacterium  phlei  attacks  only  L-asparagine  and 
this  deamidation  is  quite  well  inhibited  by  D-asparagine  (Grossowicz  and 
Halpern,  1956  a).  It  is  a  particularly  clear  and  straightforward  instance 
of  stereomeric  inhibition  and  a  1/v  —  1/(S)  plot  shows  it  to  be  completely 
competitive.  Ai'ound  75%  inhibition  is  produced  by  40  mM  D-asparagine 
when  L-asparagine  is  10  mM.  On  the  other  hand,  the  asparaginases  of 
Bacillus  coagulans  and  B.  stearothermophilus  are  inhibited  by  D-asparagine 
but  not  competitively  (Manning  and  Campbell,  1957).  The  type  of  inhi- 
bition is  difficult  to  designate  since  the  double  reciprocal  plots  intersect 
to  the  right  of  the  ordinate,  that  is,  the  inhibition  does  not  tend  toward 
noncompetitive  kinetics.  A  yet  more  complex  situation  is  presented  in  the 
depression  of  the  formation  of  a-amylase  by  D-aspartate  in  Pseudomonas 
saccharophila,  0.2  mM  blocking  the  protein  synthesis  completely  (Eisen- 
stadt  et  a].,  1959).  The  inhibition  is  readily  reversed  by  L-aspartate  and  is 
characterized  by  a  fairly  long  lag  period  before  inhibition  is  observed. 
The  inhibition  was  assvimed  to  be  on  the  reaction: 

L- Aspartate  +  IMP  +  GTP  ->  fumarate  +  AMP  +  GDP  +  P 

thus  producing  an  impairment  in  AMP  synthesis,  this  secondarily  disturbing 
the  formation  of  ATP  and  the  activation  of  amino  acids  for  protein  synthesis. 
Examination  of  this  reaction  in  cell-free  extracts  showed  that  D-aspartate 
does  indeed  inhibit  competitively.  Another  type  of  inhibition  is  exhibited 
by  pea  glutamine  synthetase  with  L-glutamate,  ammonia,  and  ATP  as 
substrates  (Varner,  1960).  D-Glutamate  inhibits  the  formation  of  L-gluta- 
mine.  However,  D-glutamate  is  also  a  substrate  and  actually  has  a  lower 
K„^  although  the  reaction  rate  is  slower  than  with  L-glutamate  (^,„  for 
L-glutamate  is  10  mM  and  for  D-glutamate  is  2  mM).  This  is  then  an  exam- 
ple of  a  competitive  inhibition  between  isomeric  substrates.  It  may  also 
be  mentioned  that  D-glutamate  inhibits,  although  quite  weakly,  the  de- 
carboxylation of  L-glutamate  by  bacterial  glutamate  decarboxylase  (Ro- 
berts,  1953). 

The  splitting  of  L-histidine  by  rat  liver  histidase  is  inhibited  by  D-histi- 
dine,  11%  inhibition  being  given  by  2  mM,  36%  by  12  mM,  55%  by  24 
mM,  and  85%  by  48  mM  when  the  L-histidine  is  12  mM  (Edlbacher  et  al., 
1940).  The  formation  of  L-histidine  from  L-histidinol  occurs  in  two  steps: 

L-Histidinol   -^  L-histidinal  ->  L-liistidine 

The  enzymes  catalyzing  both  these  reactions  are  inhibited  by  D-histidinol 
and  D-histidinal  competitively  (Adams,  1955).  The  Ki  for  D-histidinol  is 
0.05  mM  for  both  oxidations  by  a  yeast  preparation  and  it  is  possible  that 
only  a  single  enzyme  is  involved. 


270  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Germination  of  Bacillus  cereus  spores  is  induced  by  certain  amino  acids, 
such  as  L-alanine,  and  it  would  appear  in  this  case  that  L-alanine  dehydro- 
genase is  essential  for  the  activation  process  (O'Connor  and  Halvorson, 
1961  b).  D- Alanine  and  some  other  analogs,  such  as  D-a-amino-w-butyrate, 
inhibit  the  germination  when  it  is  induced  by  L-alanine,  and  also  inhibit 
the  oxidative  deamination  of  L-alanine  and  to  a  lesser  extent  other  amino 
acids,  there  being  a  good  correlation  between  these  two  inhibitory  actions 
in  the  series  of  analogs  used  (see  accompanying  tabulation).  The  inhibition 


%  Inhibition  of  deamination 
Substrate  ^^  D-alanine  (100  mi/) 


L-Alanine  61 

L-a-Amino-?i-butyrate  59 

L-Norvaline  48 

L-Serine  28 

L-Valine  24 

L-Cysteine  15 

L-Isoleucine  12 

L-Leucine  0 

L-Phenylalanine  0 


by  D-alanine  is,  however,  by  no  means  specific  for  L-alanine.  It  is  possible 
that  several  enzymes,  with  different  susceptibilities  to  D-alanine,  are  in- 
volved in  the  deamination  of  the  various  amino  acids. 

An  interesting  illustration  of  optical  specificity  is  provided  by  the  0- 
phosphoserine  phosphatase  from  chicken  liver  (Neuhaus  and  Byrne,  1960), 
Both  L-phosphoserine  and  D-phosphoserine  are  substrates  and  both  l- 
serine  and  D-serine  inhibit.  The  L-serine  is  a  much  more  potent  inhibitor 
(see   accompanying   tabulation),    l- Alanine   also   inhibits,   being   between 


K, 

(railf) 

Substrate 

L-Serine 

D-Serine 

L-Phosphoserine 
D-Phosphoserine 

0.68 
0.70 

27 
29 

l-  and  D-serine  in  potency,  but  D-alanine  does  not  inhibit.  The  inhibition 
by  L-serine  seems  to  be  uncompetitive  from  a  double  reciprocal  plot,  but 


ANALOGS  WHICH  ARE  ISOMERS  OF  SUBSTRATES  271 

this  is  apparent  only  and  the  type  of  inhibition  does  not  fit  into  the  usual 
classical  categories.  The  following  scheme  was  proposed: 

E  +  PS  ^  EPS  :^  EP  +  S 

E  +  P 

where  PS  is  phosphoserine.  It  was  shown  to  fit  the  data  kinetically  if 
k_Jki  is  small.  This  example  shows  well  the  danger  of  uncritically  accepting 
the  usual  interpretation  of  a  plotting  procedure  since,  as  emphasized 
previously,  there  are  types  of  inhibition  different  from  those  included 
in  the  classic  formulations. 

a-Chymotrypsin  hydrolyzes  the  L-isomers  of  various  tryptophanamides 
and  tyrosinamides,  and  these  reactions  are  usually  inhibited  by  the  D-iso- 
mers  (Huang  and  Niemann,  1952;  Manning  and  Niemann,  1958).  When 
the  substrate  is  nicotinyl-L-tryptophanamide  (Kg  =  2.7  mM),  the  reaction 
is  inhibited  by  nicotinyl-D-tryptophanamide  (K^  =1.4  mM)  and  a  number 
of  other  derivatives  of  D-tryptophanamide.  The  hydrolysis  of  several  de- 
rivatives of  L-tyrosinamide  is  similarly  inhibited  by  the  D-isomers  (see  ac- 
companying tabulation).  It  is  to  be  noted  that  in  every  case  the  D-isomer 
is  bound  more  tightly  than  the  L-isomer,  assuming  that  K^  does  indeed 
represent  a  dissociation  constant.  The  possible  forces  binding  these  sub- 
stances to  the  enzyme  will  be  discussed  in  a  later  section  (pages  370-375). 

Tyrosinamide  Kg  for  L-isomer         K^  for  D-isomer 


Nicotinyl- 

12 

9 

Chloroacetyl- 

27 

6.5 

Trifluoroacetyl- 

26 

20 

Acetyl- 

32 

12 

Anomeric  Analogs 

Michaelis  and  Pechstein  (1914),  in  their  early  work  on  /5-fructofuranosi- 
dase,  and  Michaelis  and  Rona  (1914),  studying  yeast  maltase  inhibition, 
concluded  that  the  configuration  around  carbon  1  of  carbohydrates  (i.e., 
a-  and  /5-anomers)  is  of  importance  in  determining  the  affinity  of  these 
substances  for  the  enzymes,  since  a-methylglucoside  inhibits  both  enzymes 
quite  potently  whereas  /?-methylglucoside  inhibits  very  little  or  not  at 
aU.  The  splitting  of  phenol-/?-glucosides  by  taka-/?-glucosidase  is  also  in- 
hibited by  phenol-a-glucoside  (Ezaki,  1940).  The  synthesis  of  polysaccha- 
ride from  a-D-glucose-1 -phosphate  by  muscle  phosphorylase  is  not  inhib- 
ited by  /?-D-glucose-l-phosphate;  however,  it  is  interesting  that  a-methyl- 


272  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

glucoside  inhibits  while  /5-methylglucoside  does  not,  indicating  the  impor- 
tance of  the  carbon  1  configuration  (Campbell  et  al.,  1952).  In  general,  the 
/5-anomers  cannot  act  as  either  substrates  or  inhibitors  of  phosphorylase. 
The  /^-glucuronidase  of  mouse  liver  is  also  stereospecific,  since  menthyl- 
/?-glucoronide  is  a  substrate  but  menthyl-a-glucuronide  only  a  very  weak 
inhibitor  (Levvy  and  Marsh,  1952). 

Positional  Analogs 

Isomers  in  which  a  ring  group  is  moved  from  one  position  on  the  ring 
to  another  are  generally  not  inhibitory  due  to  the  fairly  marked  structural 
changes  involved.  There  are  exceptions,  however,  and  one  of  the  most 
striking  is  the  inhibition  of  the  oxidation  of  p-hydroxyphenylpyruvate  to 
homogentisate  by  m-hydroxyphenylpyruvate  in  preparations  from  dog  li- 
ver (La  Du  and  Zannoni,  1955).  A  depression  of  50%  is  seen  with  0.2  mM 
and  over  90%  with  0.5  mM  m-hydroxyphenylpyruvate  when  the  substrate 
concentration  is  presumably  1.2  raM,  indicating  a  tighter  binding  to  the 
enzyme  of  the  m-isomer.  Other  examples  of  positional  isomers  will  be 
encountered  in  later  sections. 

Geometric  Isomeric  Analogs 

One  would  not  expect  that  potent  inhibitors  would  be  found  in  cis 
and  trans  pairs  because  of  the  different  molecular  configurations.  We  have 
already  seen  that  fumarate  and  maleate  differ  markedly  in  their  reactions 
with  succinate  dehydrogenase  (page  34).  The  outstanding  exception  to 
this  rule  is  the  well-known  inhibition  of  aconitase  by  fraws-aconitate.  This 
enzyme  catalyzes  the  interconversion  of  the  tricarboxylates: 

Citrate  ^  cts-aconitate  ;fi  isocitrate 

although  perhaps  cis-aconitate  is  not  an  obligatory  intermediate  between 
citrate  and  isocitrate.  Bernheim  (1928),  impressed  by  the  results  obtained 
by  Quastel  with  malonate,  tested  the  effect  of  frans-aconitate  on  liver 
"citric  dehydrogenase"  (the  reduction  of  the  methylene  blue  used  in  this 
system  was  actually  due  to  the  oxidation  of  isocitrate  formed  from 
citrate  via  aconitase)  and  found  definite  inhibition.  He  believed  the  inhi- 
bition to  be  related  to  the  structural  similarity  between  citrate  and  trans- 
aconitate,  stating,  "The  curve  obtained  seems  to  indicate  that  the  aconitic 
acid  is  adsorbed  on  the  enzyme  so  that  part  of  the  surface  is  unavailable 
for  citric  acid."  Twenty  years  later  a  thorough  study  of  this  inhibition  was 
made  by  Saffran  and  Prado  (1949),  using  aconitase  from  pigeon  breast 
muscle.  Both  the  conversion  of  m-aconitate  to  citrate  and  the  disappearance 
of  citrate  are  inhibited  by  irans-aconitate.    However,   trans-acomta.te  is 


ANALOGS   WHICH  ARE  ISOMERS  OF  SUBSTRATES  273 

not  bound  as  tightly  to  the  enzyme  as  are  the  substrates.  When  citrate  is 
3.3  mM,  50%  inhibition  is  found  with  16  xnM  trans-a.comta.te;  since  K^^ 
for  citrate  is  roughly  1  mM,  K,  is  approximately  4  mM.  The  inhibition  is 
competitive  although  the  Ijv  —  1/(S)  plots  are  not  ideal,  perhaps  because 
of  some  enzyme  inactivation  or  failure  to  achieve  equilibrium.  The  inhibi- 
tion of  rat  mammary  gland  aconitase  seems  to  be  somewhat  more  potent, 
since  equimolar  concentrations  of  citrate  and  ^raws-aconitate  lead  to  around 
50%  inhibition  (Abraham  et  al.,  1960).  Studies  on  the  stereospecificity  and 
deuterium  transfer  during  reactions  catalyzed  by  aconitase  (Speyer  and 
Dickman,  1956;  Englard  and  Colowick,  1957)  point  to  a  three-point  at- 
tachment of  the  tricarboxylates  (and  perhaps  an  intermediate  carbonium 
ion)  to  the  apoenzyme  and  Fe++.  The  carboxyl  groups  in  trans-aconitate 
would  not  appear  to  be  in  such  a  favorable  position  as  in  m-aconitate  for 
the  formation  of  this  complex  and  this  might  explain  the  relatively  weaker 
binding.  The  effects  of  trans-acomtate  on  other  enzymes  have  been  little 
investigated  but  it  has  been  found  to  be  a  fairly  potent  inhibitor  of  fu- 
marase,  Z^  being  0.63  mM  at  pH  6.35  (Massey,  1953  b).  The  K,  increases 
with  rise  in  the  pH  (Fig.  1-14-11)  and,  as  with  malonate,  the  formation  of 
the  EI  complex  is  exothermic  at  low  temperatures  and  endothermic  at 
high  temperatures. 

It  is  somewhat  surprising  that  tro ns-aconitate  is  a  reasonably  effective 
inhibitor  of  the  respiration  of  intact  cells,  inasmuch  as  penetration  into  the 
cells  should  be  difficult.  The  following  inhibitions  have  been  observed: 
22-36%  of  endogenous  respiration  of  various  tumor  shces  and  28%  of 
endogenous  respiration  of  liver  slices  at  unspecified  concentration  (Wein- 
house  et  al,  1951),  25%  of  Paramecium  respiration  at  10  mM  (Holland  and 
Humphrey,  1953),  and  40%  of  the  ion-linked  respiration  of  barley  roots 
at  20  mM  (Ordin  and  Jacobson,  1955).  However,  no  inhibition  of  the 
respiration  of  Australorbis  mince  at  10  mM  was  reported  (Weinbach, 
1953).  The  oxygen  uptake  resulting  from  the  addition  of  citrate  or  cis- 
aconitate  to  rat  liver  slices  is  strongly  inhibited  by  30  mM  ^raws-aconitate 
(Sherman  and  Corley,  1952).  The  most  complete  study  of  respiratory  in- 
hibition is  by  Saffran  and  Prado  (1949)  with  rat  liver  and  kidney  slices. 
In  the  latter  the  inhibition  is  27%  at  2  mM  and  73%  at  20  mM,  which  is 
quite  comparable  to  malonate.  The  inhibition  by  2  mM  ^rans-aconitate 
is  not  altered  by  adding  malate  or  fumarate,  but  is  reversed  with  5  mM 
citrate  or  cis-aconitate.  In  other  experiments  the  sensitivity  to  trans- 
aconitate  is  unexplainably  less.  The  inhibition  of  aconitase  in  liver  and 
mammary  gland  homogenates  leads  to  a  fairly  marked  depression  of  the 
conversion  of  citrate  to  CO2  and  of  acetate  to  fatty  acids  by  fmws-aconitate 
(Abraham  et  al.,  1960).  The  synthesis  of  mammary  fatty  acids  is  inhibited 
45%  by  7.1  mM  and  75%  by  21.4  mM.  Accumulation  of  citrate  accompa- 
nies the  inhibition  in  kidney,  liver,  and  tumor  slices  (Weinhouse  et  al., 


274  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

1951;  Saffran  and  Prado,  1949),  and  this  is  augmented  by  the  addition  of 
cycle  intermediates  such  as  pyruvate  or  malate.  These  results  indicate 
clearly  that  aconitase  is  being  inhibited  intracellular ly. 

Essentially  nothing  is  known  of  the  possible  effects  of  ^rans-aconitate  on 
cellular  functions.  No  depression  of  Paramecium  motility  is  seen  at  10  mM 
(Holland  and  Humphrey,  1953).  However,  the  active  transport  of  ions  by 
barley  roots  is  markedly  reduced  (Ordin  and  Jacobson,  1955).  K+  and  Br" 
uptakes  are  inhibited  32%  and  33%,  respectively,  by  10  milf  ^raws-aco- 
nitate,  and  63%  and  47%,  respectively,  by  20  mM.  These  inhibitions  are 
probably  not  specific  but  the  result  of  the  depression  of  respiration. 

^rans- Aconitate  is  known  to  occur  naturally  in  many  plant  tissues  and 
is  abundant  in  sugar  cane  juice.  It  is  formed  from  acetate-C^^  in  corn  tis- 
sues and  95%  of  the  aconitate  which  accumulates  is  in  the  trans  form,  it 
being  out  of  equilibrium  with  the  cycle  acids;  further  evidence  for  its  com- 
partmentalization  during  endogenous  formation  is  provided  by  the  fact 
that  it  is  metabolized  quite  readily  when  it  is  added  to  corn  roots  (Mac- 
Lennan  and  Beevers,  1964).  It  was  suggested  by  Rao  and  Altekar  (1961) 
that  it  may  arise  from  ci'.s-aconitate  through  the  mediation  of  an  aconitate 
isomerase,  which  they  isolated  from  soil  organisms.  Some  pseudomonads 
are  capable  of  metabolizing  fraw^^-aconitate  without  previous  exposure  to 
it  and  other  strains  can  adapt  to  utilizing  it  (Altekar  and  Rao,  1963). 


FUMARASE 

Fumarase  has  been  studied  more  intensively  than  most  enzymes  with 
regard  to  interactions  with  competitive  inhibitors,  the  effects  of  pH  on 
these  interactions,  and  the  nature  of  the  active  center.  A  generalized 
representation  of  the  bindings  of  fumarate  and  L-malate  to  the  enzyme  is 
shown  in  Fig.  1-6-2,  the  pH  effects  are  discussed  in  Chapter  1-14  (page 
691),  the  apparent  p^,'s  for  various  competitive  inhibitors  are  given  in 
Table  1-14-2,  and  the  P-K'^'s  of  the  two  catalytically  active  sites  for 
fumarase  and  its  substrate  complexes  are  given  in  Table  1-14-3. 

Emphasis  in  this  section  will  be  directed  to  a  more  accurate  delineation 
of  the  active  center  configuration  and  to  a  more  quantitative  expression 
of  the  ways  in  which  competitive  inhibitors  interact  with  the  active  center. 
Fumarase  possesses  four  important  groups:  two  cationic  groups  for  binding 
the  COO"  groups  of  the  substrate  in  the  trans  position,  and  two  ionizable 
groups  interacting  with  the  groups  on  the  a-  and  /5-carbon  atoms  and  in- 
volved in  the  addition  or  removal  of  water.  The  latter  enzyme  groups  wiU 
be  designated  as  Rl  and  R^  in  conformity  with  Wigler  and  Alberty  (1960); 
each  may  exist  in  the  protonated  form,  RlH  or  Rj^H.  The  p^^'s  of  these 
groups,  which  are  6.3  and  6.9  in  the  free  enzyme,  point  to  their  phenolic 
or  imidazole  nature;  indeed,  it  is  possible  that  these  two  groups  are  identical, 


rUMARASE 


275 


but  evidence  against  this  comes  from  the  study  of  inhibition  by  the  tar- 
trates. The  catalytically  active  form  of  fumarase  may  be  represented  by 
EH,  in  which  one  of  these  groups  is  protonated,  although  both  E  and  EHg 
are  capable  of  binding  both  substrates  and  inhibitors. 

The  values  of  K^  for  several  competitive  inhibitors  at  pH  6.35  and  23° 
(Table  2-1)  may  be  used  as  a  rough  and  provisional  means  of  evaluating  the 
relative  energies  of  ])inding,  bearing  in  mind  that  these  are  apparent  K-&, 
that  the  enzyme  exists  in  three  different  ionized  states  for  each  of  which 
the  binding  is  different  (so  that  the  K^s  are  in  a  sense  composite),  and 
that  the  various  inhibitors  alter  the  pA^^'s  of  the  enzyme  groups  in  different 
ways  in  the  EI  complexes.  However,  some  reasonable  conclusions  may 
be  drawn  from  these  AF  values. 

Table  2-1 

Inhibitor  Constants"  and  Relative  Binding  Energies 
FOR  Competitive  Inhibitors  of  Fumarase 


Inhibitor 


Apparent  K^ 
(mM) 


Relative  —AF 
(kcal/mole) 


Adipate 

Succinate 

Glutarate 

Malonate 

D -Tartrate 

Mesaconate 

Maleate 

L-a-Hydroxy-/5-sulfopropionate  * 

D-Malate 

Citrate 

<rans-Aconitate 


100 

1.35 

52 

1.73 

46 

1.80 

40 

1.90 

25 

2.16 

25 

2.16 

11 

2.64 

10 

2.70 

6.3 

2.97 

3.5 

3.32 

0.63 

4.32 

°  Values  of  Ki  determined  at  pH  6.35  and  23°. 

*  The  Ki  for  L-a-hydroxy-^-sulfopropionate  was  changed  from  16.5  mM  as  given 
in  the  table  (Massey,  1953  b)  to  correspond  to  the  value  in  the  curve  presented  (16.5 
is  probably  a  misprint  for  10.5.) 


[A)  Since  all  monocarboxylates  and  the  methyl  ester  of  fumarate  are 
without  inhibitory  activity,  it  must  be  assumed  that  at  least  two  negatively 
charged  groups  are  necessary  for  binding.  However,  it  is  evident  that  for 
the  more  tightly  bound  substances  other  attraction  forces  are  involved.  If 
we  assume  that  these  additional  forces  arise  from  hydrogen  bonding  be- 


276  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

tween  hydroxyl  groups  and  the  Rl  and  R^  enzyme  groups,  polarization 
of  double  bonds,  and  interactions  of  a  third  C00~  group  (in  the  tricar- 
boxylates),  the  tentative  values  shown  in  the  following  tabulation  may  be 
assigned  for  the  contributions  made  by  the  various  interactions  to  the 
total  binding: 

Two  C00~  groups  1.75       kcal/mole 

— OH  group  hydrogen  bonding  0.5-1.5  kcal/mole 

— C=C —  polarization  1.9         kcal/mole 

Additional  CHgCOO-  group  0.5         kcal/mole 


The  hydrogen  bonding  and  polarization  values  are  minimal  since,  in  part, 
they  were  derived  from  the  K„'s  of  the  substrates  (for  fumarate  K„^  = 
=  1.78  mM  and  —  JF  =  3.71  kcal/mole,  and  for  L-malate  Z,„  ==  4.0  mM 
and  —  AF  =  3.24  kcal/mole).  The  ^„/s  may  not  represent  dissociation 
constants  but  in  any  case  the  true  K^'s  would  be  equivalent  to  or  smaller 
than  the  K„'s,  so  that  the  binding  energies  for  the  substrates  may  be 
somewhat  higher.  The  cis  configuration  of  maleate  reduces  the  attraction, 
but  the  value  for  maleate  in  the  table  should  be  corrected  since  the  pK^ 
—  5.9  (at  23°  and  around  0.1  ionic  strength)  and  only  74%  of  the  total 
maleic  acid  would  be  in  the  form  of  maleate=:  this  increases  —  JF  to  2.82 
kcal/mole.  The  difference  in  binding  between  fumarate  and  maleate  is 
thus  at  least  0.9  kcal/mole.  It  is  also  interesting  that  the  introduction  of 
a  methyl  group  into  fumarate  to  form  mesaconate  brings  about  a  1.55 
kcal/mole  or  greater  reduction  in  the  binding  energy,  resulting  possibly 
from  a  steric  displacement  and  lowered  polarization  interaction.  The  1.9 
kcal/mole  estimated  for  electrical  polarization  of  the  double  bond  is  not 
unreasonable  and  actually  corresponds  fairly  closely  to  that  calculated, 
using  appropriate  molar  refractions  and  an  interaction  distance  of  4  A. 
A  factor  of  unknown  importance  is  the  possible  deformation  of  the  dicar- 
boxylates  to  fit  the  active  site  and  the  energies  that  would  be  involved 
with  the  different  inhibitors. 

(B)  If  these  conclusions  are  valid,  the  interaction  energy  for  fumarate  is 
approximately  half  due  to  coulombic  ion-ion  forces  and  half  due  to  the 
inductive  polarization  by  a  strong  dipole.  It  is  possible  that  one  of  the  R 
groups  on  the  enzyme  is  positively  charged  and  the  other  negatively 
charged  on  the  active  enzyme,  as  suggested  by  Massey  (1953  b).  The  ionic 
interactions  serve  to  orient  the  fumarate  at  the  active  center,  and  the 
polarization  not  only  stabilizes  the  complex  but  initiates  the  addition  of 
water. 

(C)  The  third  COO"  group  of  citrate  and  trans-aconitate  seems  to  be 
able  to  interact  with  an  adjacent  positive  group  on  the  enzyme,  but  rela- 


FUMARASE  277 

tively  weakly  (0.3-0.6  kcal/mole).  In  fact,  the  low  interaction  energies  of  the 
terminal  COO"  groups  (0.87  kcal/mole  for  each  group,  which  may  be  com- 
pared with  the  3.3-3.6  kcal/mole  binding  per  COO"  group  of  malonate  or 
succinate  on  succinate  dehydrogenase)  might  indicate  that  the  distance 
between  them  and  the  enzyme  cationic  groups  is  relatively  great  (perhaps 
12-15  A),  or  could  even  point  to  a  type  of  interaction  other  than  ion-ion 
attraction. 

(D)  L-a-Hydroxy-/?-sulfopropionate  is  bound  at  least  0.54  kcal/mole  less 
tightly  than  L-malate,  much  of  the  affinity  of  this  analog  resulting  from 
the  OH  group  interaction.  This  would  indicate  that  the  sulfonate  group 
is  not  a  very  good  substitute  for  a  COO"  group  in  this  case.  It  would  be 
interesting,  in  this  connection,  to  have  inhibition  data  on  L-/5-sulfopropio- 
nate. 

(E)  When  one  turns  to  the  effects  of  pH  on  the  binding  of  these  inhi- 
bitors, it  is  evident  that  the  situation  is  more  complex  than  assumed  from 
the  data  at  a  single  pH  (see  Fig.  1-14-11).  Several  inhibitors  exhibit  a  pro- 
gressive decline  in  binding  with  increase  in  the  pH,  but  in  the  case  of  fu- 
marate  the  pZ,„  rises  between  pH  7  and  8,  and  the  pZ,„-pH  curve  for  l- 
malate  shows  several  changes  of  slope.  The  inhibitor  D-malate  also  shows 
an  increase  in  binding  between  pH  7  and  8.  Massey  (1953  b)  suggested  that 
binding  to  different  sites  might  be  involved.  However,  since  deviant  be- 
havior is  noted  with  substances  containing  a  double  bond  or  OH  group, 
it  is  possible  that  the  affects  of  pH  on  the  polarization  and  hydrogen  bonding 
interactions  may  be  involved.  Succinate  exhibits  a  linear  decrease  in  jiK^ 
with  pH  whereas  fumarate  behaves  quite  differently.  Mesaconate  and  ma- 
leate  have  succinate-type  pH  dependences  and  it  is  possible  that  the  pola- 
rization interaction  is  sterically  prevented  in  these  substances,  as  postulated 
above.  Apparently  some  change  in  the  active  center  ionization  occurs 
between  pH  7  and  8  which  alters  these  interactions,  and  we  shall  return  to 
this  problem  later  when  more  recent  inhibition  data  have  been  presented. 
(If  the  crude  approach  to  these  problems  in  the  preceding  paragraphs 
serves  either  to  irritate  or  activate  others  to  further  theoretical  or  experi- 
mental study,  a  purpose  will  have  been  accomplished.) 

One  would  expect  the  most  potent  competitive  inhibitors  of  fumarase 
to  have  either  a  polarizable  group  (such  as  — C=C — )  or  a  group  capable 
of  forming  hydrogen  bonds  (such  as  OH).  Substitution  of  groups  at  the 
double  bond  of  fumarate  seems  to  reduce  the  binding,  and  acetylene-di- 
carboxylate  has  not  been  studied.  Thus  one  is  left  with  the  tartrates  as 
possibly  interesting  inhibitors,  and  they  were  investigated  by  Wigler  and 
Alberty  (1960)  in  an  excellent  study  designed  to  establish  the  more  intimate 
nature  of  the  catalysis.  The  variation  of  the  inhibitions  with  pH  allowed  the 
determination  of  the  p^^'s  of  the  enzyme  groups,  the  changes  in  these  pro- 
duced by  complex ing  with  the  inhibitors,  and  the  dissociation  constants 


278  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


COO 
H^i   .H 
^C^ 

1 

COO' 

1 

COO 
H.   i    OH 

1 

COO 
H^  i    .OH 

^C^ 

1 

1 

H-^^^H 

coo- 

1 
H-^  ;     OH 

coo- 

1 
HO^  i  ^H 

coo- 

1 

Q 

U^ 1 ^OH 

coo- 

Succinate 

L-Tartrate 

D -Tartrate 

wieso- Tartrate 

of  the  inhibitors  with  the  variously  protonated  forms  of  the  enzyme.  It 
was  found  that  meso-tartrate  is  the  most  potent  inhibitor  of  this  group 
and  it  was  concluded  that  this  configuration  of  OH  groups  allows  the 
formation  of  two  hydrogen  bonds  with  the  Rl  ^^^  ^d  enzyme  groups 
(Fig.  2-2).   The  singly  protonated  form  of  the  enzyme,  EH,  binds  the 

C-—C  FUMARATE 

/         \ 

Rp-H  COO  0 


© 


ooc 


Rjj-H 


L-MALATE 


MESO- TARTRATE 


Fig.  2-2.  Simplified  scheme  of  the  fumarase  active  site   described   by  Wigler  and 

Alberty  (1960).  The  cationic,  R^,  and  R^  groups  occur  on  different  levels  and  are  so 

located  they  can  interact  with  certain  isomers  of  substrates  and  inhibitors. 

meso-tartrate  most  tightly.  The  binding  energies  contributed  by  the  hy- 
drogen bonds  can  be  estimated  from  the  relative  interactions  of  succinate 
and  the  tartrates.  Weak  hydrogen  bonds  are  indicated  for  the  doubly  pro- 
tonated form  of  the  enzyme,  EHgl,  with  D-  and  L-tartrates,  while  stronger 
bonds  (—  2.8  to  —  3.3  kcal/mole)  are  formed  with  meso-tartrate  and  the 
less  protonated  enzyme  (see  tabulation). 


INHIBITION  OF  XANTHINE   OXIDASE  279 


AF    for    displacement   of  succinate 
(kcal/mole) 

EH2I  EHI  EI 

D-Tartrate  —0.5  0  +0.4 

L-Tartrate  —0.5  +0.8  +0.7 

meso-Tartrate  — 1.6  —  2.8  —  3.3 

The  fumarase  from,  liver  is  inhibited  differently  from  the  heart  enzyme, 
in    that    mesaconate    (methylfumarate)    is    inactive    whereas    citraconate 

H3C      ^COO'  H         COO"                     V*^ 

C  C                                   C=CH2 

II  II                                    I            ' 

/C^         .  ^C^                                CH3 

H         COO  H3C          H                              i,^- 

Citraconate  Crotonate  Itaconate 

(methylmaleate)  inhibits  well  (Jacobsohn,  1953).  Itaconate  inhibits  less 
potently  and  crotonate  even  less  potently  (about  the  same  as  succinate). 
The  fact  that  crotonate  inhibits  at  all  is  interesting,  in  that  it  suggests 
that  one  C00~  is  sufficient  if  a  double  bond  is  present.  czs-Aconitate  and 
^raws-aconitate  inhibit  equally  and  weakly.  DL-/5-Fluoromalate  inhibits  com- 
petitively the  conversion  of  malate  to  fumarate  by  fumarase  (Krasna, 
1961).  Assuming  that  both  optical  isomers  inhibit  equally  (which  may 
not  be  true),  K^  =  28  mM,  and  K„^  for  malate  is  3.5  niM.  The  inhibition 
of  malate  dehydrogenase  is  much  stronger,  if  ^  being  0.16  mM,  which  may 
be  compared  to  a  K,,^  of  11  mM  for  malate. 

INHIBITION  OF  XANTHINE  OXIDASE 
BY   PURINE  ANALOGS  AND   PTERIDINES 

Some  potent  and  specific  inhibitors  of  xanthine  oxidase  have  been  dis- 
covered and  have  proved  to  be  interesting  not  only  on  the  enzyme  level 
but  because  of  the  disturbances  in  purine  metabolism  produced  in  whole 
animals.  Xanthine  oxidase  catalyzes  the  oxidation  of  hypoxanthine  and 


H3C     ^coo 

^c 

II 

OOC          H 

Mesaconate 

0 

1 1 

H 

1 

H 

Xanthine 

N         N' 


Hypoxanthine  Xanthine  Uric  acid 


280 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


xanthine  to  uric  acid,  each  step  essentially  involving  the  addition  of  water 
and  the  removal  of  two  hydrogen  atoms  which  are  transferred  to  oxygen 
along  a  typical  electron  transport  sequence.  Hydroxypurines  exhibit  keto- 
enol  tautomerism,  and  it  is  only  recently  that  spectroscopic  evidence  point- 
ing to  the  predominance  of  the  keto  form  has  been  obtained  (Mason, 
1957).  The  structures  of  the  purines  and  pteridines  to  be  discussed  will 
be  written  as  far  as  possible  in  conformity  to  these  results.  However,  in 
some  cases  it  is  difficult  to  assign  the  most  important  structure,  especially 
in  multiply  substituted  compounds.  Ionization  may  also  be  a  complicating 
factor.  Most  purines  appear  to  be  predominantly  neutral  at  physiological 
pH  (piiC^  between  1  and  4;  pK^,  between  8  and  12),  but  some,  such  as  uric 
acid,  lose  a  proton  in  the  acid  pH  range  and  physiologically  exist  as  anions 
(e.g.,  pK^  for  uric  acid  is  5.4).  The  site  of  loss  of  the  proton  is  not  known 
and  it  is  quite  possible  that  the  anions  should  be  considered  as  being 
equilibrium  mixtures  of  several  structures.  The  form  in  which  they  are 
written,  hence,  does  not  imply  that  this  structure  is  the  only  one  present, 
or  even  that  it  is  necessarily  the  most  important  structure.  Such  considera- 
tions become  important  in  treating  the  forces  between  these  compounds 
and  the  enzyme.  A  final  factor  must  be  borne  in  mind:  the  most  stable 
form  in  solution  is  possibly  not  the  form  in  which  the  purine  is  bound  to 
the  enzyme  surface,  inasmuch  as  the  interaction  may  modify  the  structure 
appreciably. 

Purine  Analogs 

Dixon  and  Tliurlow  (1924)  reported  that  xanthine  oxidase  is  inhibited 
by  various  purines,  such  as  adenine  and  uric  acid,  but  no  quantitative  data 
were  given.  There  was  no  further  investigation  until  the  introduction  of 
the  azapurines.  2-Azaadenine  and  2-azahypoxanthine,  like  many  purine 


Purine 


Guanine 


8-Azaguanine 


NH. 


NH, 


NH. 


Pyrazoloisoguanine 


INHIBITION  OF  XANTHINE  OXIDASE  281 

analogs,  are  oxidized  by  the  enzyme  in  the  8-position  (Shaw  and  Woolley, 
1952).  Eqiiimolar  concentrations  of  2-azaadenine  prolong  the  formation  of 
urate  from  xanthine  2-fold  and  from  hypoxanthine  4-fold,  the  inhibition 
being  competitive.  The  kinetics  in  such  situations  may  be  complicated 
by  two  factors:  (1)  the  disappearance  of  the  inhibitor  (Shaw  and  Woolley 
found,  for  example,  that  the  azaadenine  essentially  all  disappeared  before 
much  urate  was  formed),  and  (2)  the  inhibition  produced  by  the  product 
of  the  inhibitor  oxidation.  8-Azapurine  and  all  of  its  monohydroxyl  and 
monoamino  derivatives  are  oxidized  by  xanthine  oxidase  and  the  products 
are  frequently  inhibitory  not  only  to  xanthine  oxidase  but  to  other  enzymes. 
2-Amino-8-azapurine  is  converted  to  8-azaguanine  and  hence  can  be  used 
as  a  precursor  of  this  inhibitor  (Bergmann  et  al.,  1959). 

Certain  inhibitions  of  xanthine  oxidase  by  purine  compounds  are  sum- 
marized in  Table  2.2  The  inhibitions  are  not  always  competitive  despite 
the  close  similarity  of  substrate  and  inhibitor  structures.  Some  of  the  simple 
analogs  are  bound  more  tightly  to  the  enzyme  than  are  the  normal  sub- 
strates. 6-Chloropurine  and  pyrazoloisoguanine  are  bound  particularly  well 
and  this  brings  up  questions  regarding  the  forces  involved.  Very  little  is 
known  about  these  forces.  Ionic  forces  must  be  unimportant  and  it  is  pos- 
sible that  hydrogen  bonds,  coupled  with  an  appropriate  fit  of  the  bonding 
groups,  play  a  major  role.  PjTazoloisoguanine  is  the  4-amino-6-hydroxy 
derivative  of  pyrazolopyrimidine,  and  it  is  interesting  to  note  that  the 
4-amino  derivative  is  a  very  weak  inhibitor  relatively,  as  are  the  4-methyl- 
amino  and  l-methyl-4-amino  derivatives  (Feigelson  et  al.,  1957).  It  has 
been  stated  that  there  is  some  correlation  between  the  potency  of  the  xan- 
thine oxidase  inhibition  and  the  carcinostatic  activity  of  these  and  related 
compounds. 

When  inhibitory  purine  analogs  are  administered  to  animals  it  is  often 
difficult  to  determine  the  toxic  mechanisms  because  of  the  multiple  possible 
sites  for  interference.  The  biological  effects  of  6-mercaptopurine  seem  to 
be  related  to  its  conversion  to  the  ribonucleotide,  which  inhibits  inosinic 
acid  metabolism,  rather  than  to  any  direct  enzyme  inhibition  (Silberman 
and  Wyngaarden,  1961).  On  the  other  hand,  it  has  been  postulated  that 
8-azaguanine  induces  a  guanine  deficiency  by  inhibiting  xanthine  oxidase, 
which  operates  in  one  guanine  biosjoithetic  pathway  (i.e.,  hypoxanthine 
-^  xanthine  -^  guanine)  (Feigelson  and  Davidson,  1956  a).  It  has  been 
shown  in  one  instance  that  purine  metabolism  can  be  inhibited  in  vivo. 
6-Chloropurine  given  to  rats  at  80  mg  kg  inhibits  the  formation  of  C^^Og 
from  xanthine-6-C^*  about  40%  when  administered  20  min  before  the  xan- 
thine (Duggan  et  al.,  1961).  This  is  probably  not  due  to  a  direct  action 
on  xanthine  oxidase  but  to  the  formation  of  6-chlorourate  and  the  resulting 
inhibition  of  uricase.  6-Chloropurine  also  depresses  the  conversion  of  ace- 
tate to  lipid,  of  glycine  to  protein,  and  nucleic  acid  synthesis. 


282 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


P5 


O    ^ 

6  6 


^  ^ 


o  § 


+ 


CO    '^ 


tH   c  :r  cr 


112  1^ 


■^  ^  o 


c3     c3  ^  'i) 

e    c  g  c  c    s 

O      TO  Co  CO  o      f 

C     U)  M  M  c  ^ 

§  5  5  5  §  -2 


^     C     C 
™       Oi      (D 


o3  c« 

cS  cS 

bC  bD 

C  C 


^pq        hJQQPh-1S^^ 


S  fc  — I  SO  s  S  ;::! 


-ti    -ti    +i 


!^  ^  ^ 


,ifO    -<    o 


I    I 


+  +   i 


.:-  .-    (D 


.S  ^     .  "  —  ^ 


773     O     S)  ■:3    773 


^    c3    ce 
§  05  P^ 


O  S 


m 


2   5 


(U 


:a  ^ 


111 


cS 


CK    pJH    PQ 


"-'do 


t<  t< 


+ 


f=^ 


Q 


s  i^ 


o   J4    M    M 

13  ^  ^  "        ^ 
??  fer*  ^T-f  w        P- 


O 


§ 

.g 

§ 

^ 

^ 

s 

s 

s£ 

1^  -^ 

-« 

"^ 

;2  ^ 

f^    T^ 

772 

o 

Si  rs 

"  g 

S 

O 

"   § 

u  ^     tl, 


o 


o  § 


'S  'S  >>  ^ 

c  c  X  o 

cs  ce  o  ,i2 

5  5  ^  >.  •£ 


<D 


-  -3  -^ 

0)  cS  c 

■d  3  cS 

c3  bC  X 

N  N  SI 


>>  K   3  oi  <;  <c 

'-*  &      .        I        I 

C-]     GO     00 


:s  a 


^  .2 


"5    .-:< 


INHIBITION  OF  XANTHINE  OXIDASE  283 

A  new  xanthine  oxidase  competitive  inhibitor,  4-hydroxypyTazolo(3,4-d) 
pyrimidine  (allopurinol,  Zyloprim),  is  now  being  clinically  tested  in  hyper- 
uricemia and  for  the  potentiation  of  the  antitumor  activity  of  the  6-substi- 
tuted  purines  (e.g.  6-mercaptopurine)  (Elion  et  al.,  1963;  Information  for 
Investigators  report    from    the    Burroughs  Wellcome  Company).   It  is  a 


N 

I 
H 

4-Hydroxypyrazolo(3,  4-d)pyrimidine 

very  potent  inhibitor,  with  K^  =  0.000032  mM,  being  bound  around  100 
times  more  tightly  to  the  enzyme  than  is  xanthine,  but  it  is  also  a  substrate 
for  xanthine  oxidase  and  its  oxidation  product  is  likewise  a  potent  inhib- 
itor, with  Ki  =  0.000054  mM.  Mice  and  dogs  given  100  mg/kg  intraperi- 
toneally  show  an  increased  urinary  excretion  of  xanthine  and  hypoxanthine, 
with  a  decrease  in  aUantoin,  and  in  man  a  similar  action  has  been  demon- 
strated, serum  and  urinary  urate  being  depressed.  It  is  well  tolerated  by 
man  at  200-1000  mg/day  orally  up  to  several  weeks.  It  apparently  has  no 
antitumor  activity  itself  but  is  able  to  potentiate  the  action  of  6-mercapto- 
purine by  interfering  with  its  metabolism.  Mice  given  20  mg/kg  intraperi- 
toneally  along  with  6-mercaptopurine  exhibit  a  reduced  urinary  excretion 
of  thiourate.  The  value  of  this  analog  in  neoplastic  disease  and  gout  is 
not  yet  known. 

Inhibition  of  Uricase  by  Purine  Analogs 

It  is  appropriate  at  this  time  to  refer  to  certain  studies  on  uricase  (urate 
oxidase)  before  proceeding  with  the  inhibition  of  xanthine  oxidase  by  the 
pteridines.  Uricase  catalyzes  the  oxidative  opening  of  the  pyrimidine  ring 
to  form  aUantoin.  Many  methyl  and  ethyl  derivatives  of  urate  were  tested 
by  Keilin  and  Hartree  (1936)  on  an  enzyme  from  pig  liver;  they  found  that 
none  is  a  substrate  but  that  several  inhibit  quite  potently.  They  considered 
the  mechanism  to  be  competitive  and  stated,  "The  fact  that  the  methyl 
compounds  of  uric  acid,  although  not  oxidizable  by  the  enzyme,  inhibit 
the  oxidation  of  uric  acid  shows  that  these  methyl  compounds  react  with 
the  same  active  grouping  of  the  enzyme  molecule  as  uric  acid  itself."  It 
is  very  interesting  that  the  1,3,7-derivative  is  so  much  more  potent  than 
the  1,3,9-derivative,  particularly  in  view  of  the  greater  potency  of  the 
monomethyl  compounds  compared  to  the  latter  derivative  (the  urate  was 
10  mM  in  all  cases)   (see  tabulation).  2,6,8-Trisubstituted  purines  were 


284  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


Inhibitor 

Concentration 
(mif) 

% 

Inhibition 

1,3,9-Trimethylurate 

7.7 

16 

7-Methylurate 

9 

41 

3-methylurate 

9 

49 

1-methylurate 

9 

58 

1,3,7-Trimethylurate 

7.7 

68 

studied  by  Mahler  et  at.  (1956)  (see  tabulation)  and  the  results  give  some 
information  on  the  nature  of  the  binding  to  the  active  site.  Three  binding 
sites   were   recognized:    (1)    a   cationic   group   binding   the   2-substituent, 


Substituent  in  position: 

A',  {mM) 

2 

6 

8 

CI 

CI 

CI 

0.0008 

CI 

CI 

OH 

0.0013 

OH 

NH, 

NH2 

0.0018 

OH 

OH 

H 

0.012 

OH 

OH 

OH 

0.025  (A',  for  urate) 

CI 

NHa 

OH 

0.04 

OH 

NHa 

OH 

0.15 

NH2 

NHa 

OH 

0.5 

NH2 

OH 

NH2 

No  inhibition 

NH2 

NH, 

XH, 

\o  inhibition 

(2)  a  neutral  group  binding  the  8-substituent,  and  (3)  the  copper  which 
chelates  with  the  6-  substituent  and  the  7-N  atom.  It  is  difficult  to  under- 
stand on  this  basis  the  high  affinity  of  the  trichlorourate  for  the  enzyme, 
since  one  would  predict  that  substitution  of  chlorine  in  the  2-position  would 
reduce  binding  to  the  cationic  group  and  substitution  in  the  6-position  would 
interfere  with  the  chelation.  The  increased  binding  produced  by  chlorine 
substitution  might  be  in  part  the  result  of  a  stabilization  of  conjugative 
resonance  and  an  increased  hydrogen  bonding  (and  perhaps  an  increased 
chelation  of  two  N  atoms  with  the  copper).  Also  resonance  with  structures 
in  which  one  or  more  of  the  chlorine  atoms  are  in  the  form 

C1+ 

— N-— C— 

would  produce  strong  dipoles.  The  most  potent  inhibitor  apparently  is 
6-chloro-2,8-dihydroxy purine  (usually  misnamed  6-chlorourate),  which  is 


INHIBITION  OF  XANTHINE   OXIDASE  285 

formed  from  6-chloropurine  by  the  action  of  xanthine  oxidase,  K^  being 
0.00006  mif  (Duggan  and  Titus,  1959).  It  inhibits  urate  degradation  in 
the  rat  75%  at  a  dose  of  20  mg/kg  (Duggan  et  al.,  1961).  The  over  all  AF 
for  the  binding  to  uricase  is  approximately  10  kcal/mole  and  this  would 
indicate  the  formation  of  some  type  of  stable  bond.  It  was  pointed  out  that 
this  analog  is  very  stable  and  might  be  useful  in  studying  the  metabolism 
and  disposition  of  urate.  It  should  induce  urate  accumulation  in  the  tissues 
and  the  effects  of  this  might  have  some  bearing  on  the  manifestations  of 
gout.  The  inhibition  of  uricase  by  xanthine  (2,6-dihydroxypurine)  in  the 
tabulation  above  is  interesting  because  it  illustrates  a  novel  effect  in  a  multi- 
enzyme  system.  In  the  sequence: 

xanthine 

oxidase                    uricase 
Xanthine >  urate   >  allantoin 

the  initial  substrate  inhibits  the  second  enzyme  in  the  series,  causing  ac- 
cumulation of  urate  in  increasing  amounts  as  the  xanthine  concentration 
rises  (e.g.,  in  rat  liver  homogenates).  At  high  concentrations  the  urate  con- 
centration falls  due  to  the  substrate  inhibition  of  xanthine  oxidase  (Van 
Pilsun,  1953). 

Further  studies  on  uricase  have  been  reported  by  Bergmann  et  al.  (1963  a) 
and  some  of  the  results,  including  calculations  of  the  approximate  apparent 
relative  binding  energies,  are  presented  in  the  following  tabulation.  The 
xV-methylpurines  are  relatively  weak  inhibitors  and  are  not  included  in 
the  tabulation.  It  is  clear  that  the  best  inhibitors  contain  a  2-OH  group 
(designated  as  OH  for  convenience  but  the  keto  form  is  probably  dominant), 
and  this  position  is  the  most  important  in  the  binding;  substitution  of  the 
2-0II  with  a  2-SH  group  lowers  the  inhibitory  activity  markedly.  The 
weakening  of  the  binding  by  A^-substitution  points  to  the  importance  of 
the  imino  group  for  attachment.  The  inhibitions  are  generally"  sensitive  to 
the  pH  and  for  most  analogs  increase  with  a  rise  in  pH,  although  for  6- 
SH-8-OH-purine  there  is  a  decrease,  and  with  8-OH-purine  there  is  no 
effect.  The  pH  probably  influences  the  tautomerism,  which  is  quite  impor- 
tant since  the  binding  depends  on  the  states  of  the  N  and  OH  groups.  The 
K^  for  2,8-diOH-6-SH-purine  is  0.0026  niM  and  for  2,6-diOH-8-SH-purine 
is  0.00039  mM  (Bergmann  et  al,  1963  b).  The  2-OH-6,8-diSH-purine  (which 
is  6,8-dithiourate)  is  a  much  more  potent  inhibitor  and,  although  its  K^ 
was  not  given,  it  must  be  at  least  around  0.00004  mil/. 

Pteridines 

The  inhibition  of  xanthine  oxidase  by  synthetic  folate,  reported  by 
Kalckar  and  Klenow  in  1948,  was  soon  found  to  be  due  to  some  impurity, 
and  the  simultaneous  observation,  by  Lowry  and  Bessey,  of  the  very  po- 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


Substituent  in  position: 

(1)50  (mif)        Relative   —AF  (kcal/mole) 

2 

6 

8 

OH 

CH3S 

OH 

0.00013 

6.80 

OH 

SH 

SH 

0.0004 

6.12 

OH 

H 

Aza 

0.0016 

5.25 

OH 

SH 

H 

0.0027 

4.94 

OH 

OH 

SH 

0.0050 

4.55 

OH 

H 

OH 

0.0052 

4.52 

OH 

OH 

Aza 

0.0059 

4.45 

OH 

CH3S 

H 

0.0060 

4.44 

OH 

H 

H 

0.012 

4.01 

OH 

H 

OH 

0.012 

4.01 

OH 

SH 

OH 

0.014 

3.91 

OH 

CH3 

H 

0.017 

3.78 

OH 

OH 

H 

0.018 

3.76 

H 

CH3S 

OH 

0.032 

3.40 

OH 

OH 

CH3S 

0.038 

3.30 

H 

OH 

OH 

0.066 

2.95 

H 

SH 

OH 

0.070 

2.91 

SH 

OH 

SH 

0.080 

2.84 

H 

H 

OH 

0.11 

2.65 

SH 

SH 

OH 

0.15 

2.44 

SH 

OH 

H 

0.19 

2.30 

H 

OH 

H 

0.22 

2.21 

SH 

OH 

OH 

0.25 

2.13 

SH 

H 

OH 

0.50 

1.70 

H 

SH 

OH 

0.50 

1.70 

H 

SH 

H 

0.70 

1.19 

tent  inhibition  produced  by  2-amino-4-hydroxy-6-pteridyl  aldehyde,  a 
photolytic  product  of  folate,  led  to  the  conclusion,  subsequently  verified, 
that  this  was  the  contaminant  (Kalckar  et  al.,  1948;  Lowry  et  al.,  1949  a, 
b).  For  convenience  we  shall  follow  Hofstee  (1949)  in  designating  2-amino- 
4-hydroxypteridine  as  pterin  and  the  inhibitor  thus  as  pterin-6 -aldehyde* 
Pterin  and  many  of  its  substituted  derivatives  are  oxidized  to  varying 
degrees  by  xanthine  oxidase  while  other  derivatives  are  only  inhibitory. 

*  This  substance  has  been  variably  called  2-amino-4-hydroxy-6-pteridyl  aldehyde, 
2-amino-4-hydroxy-6-formylpteridine,  6-formylpteridine,  pteridylaldehyde,  2-amino- 
4-hydroxy-6-pteridine  carboxaldehyde,  and  2-amino-4-hydroxy-6-formylpterine,  in 
most  cases  without  either  justification  or  accuracy. 


INHIBITION  OF  XANTHINE   OXIDASE 
O 


H,N 


287 


H,N 


H,N 


«^N^ 

1       "^ 

HsN-^N-^ 

^N^ 

Pterin-6- 

-aldehyde 

O 

H 

1 

H-nA 

^Ny-o 

H2N^^^N-^ 

1 

1 

H 

Xanthopterin  Isoxanthopterin  Leucopterin 

Actually  all  the  photolytic  oxidation  products  of  folate  are  inhibitory 
but  pterin-6-aldehyde,  the  primary  product: 

Folate   ->   pterin-6-aldehyde   ->   pterin-6-carboxylate   ->   pterin   ->   isoxanthopterin 

is  by  far  the  most  potent;  indeed,  it  is  one  of  the  most  potent  inhibitors 
known. 

Inhibition  of  xanthine  oxidase  by  pterin-6-aldehyde  is  observed  at  a 
concentration  of  2  X  10^*  //g/ml  or  roughly  10"^  M  (Lowry  et  at.,  1949  a). 
Competitive  inhibition  with  respect  to  both  xanthine  and  pterin  has  been 
demonstrated,  and  a  /iC,  of  6  x  10"'  mM  calculated  for  the  milk  enzyme 
(Lowry  ef  al.,  1949b).  It  was  shown  that  35%  inhibition  occurs  when  enzyme- 
FAD  =  9.3  X  10-6  jnM,  the  substrate  pterin  =  78  x  10-«  mM,  and  pte- 
rin-6-aldehyde  =  2.26  X  10-«  mM,  this  indicating  that  2.26  X  10-«  mM 
inhibitor  completely  blocks  3.3  X  10-^  mM  enzyme  (on  the  basis  of  FAD 
content).  It  may  have  been  that  all  the  FAD  was  not  catalytically  active 
or  that  more  than  one  FAD  molecule  was  associated  with  one  enzyme 
molecule.  However,  there  is  no  doubt  that  this  is  a  mutual  depletion  system 
and  that  pterin-6-aldehyde  titrates  the  enzyme.  A  few  other  reports  will 
be  mentioned  to  illustrate  the  potency.  Both  the  milk  and  rat  liver  enzymes 
are  inhibited,  40-50%  inhibition  occurring  at  5-8  X  10"^  mM  when  xan- 
thine is  0.07  mM  (Kalckar  et  al,  1950).  The  xanthine  oxidase  from  Clostri- 
dium cylindrosporum  is  potently  inhibited  by  0.0002  mM  (Bradshaw  and 
Barker,  1960).  Milk  xanthine  oxidase  is  completely  inhibited  by  pterin-6- 
aldehyde  at  0.0033  mM  when  hypoxanthine  is  3.33  mM  (Petering  and 
Schmitt,  1950),  and  the  rat  intestine  enzyme  is  inhibited  completely  by 
0.067  mM  when  hypoxanthine  is  6.6  mM  (Westerfeld  and  Richert,  1952). 
In  many  experiments  the  substrate  concentrations  have  been  unnecessarily 
high  since  maximal  rates  are  usually  obtained  at  concentrations  well  below 


288  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

0.1  niM,  and  hence  the  true  potency  of  the  inhibition  has  been  somewhat 
obscured.  Pterin-6-aldehyde  is  actually  oxidized  very  slowly  by  xanthine 
oxidase  and  reversal  of  the  inhibition  develops  gradually.  The  oxidations 
of  aldehydes  (Kalckar  et  al.,  1950)  and  sulfite  (Fridovich  and  Handler, 
1957)  by  xanthine  oxidase  are  also  inhibited  by  pterin-6-aldehyde,  but  the 
oxidation  of  NADH  is  not  affected  (Lowry  et  al.,  1949  b). 

Although  it  has  generally  been  stated  that  the  inhibition  by  pterin-6- 
aldehyde  is  competitive,  the  Ijv  —  1/(S)  plots  are  not  linear  (Bradshaw 
and  Barker,  1960),  and  others  (Hofstee,  1949)  have  presented  only  quali- 
tative evidence  for  competition.  Deviations  from  the  classic  kinetics  might 
be  expected  because  of  the  mutual  depletion  of  free  enzyme  and  inhibitor 
concentrations,  the  difficulty  in  achieving  true  equilibrium,  and  the  oxida- 
tive removal  of  the  inhibitor.  There  is  no  evidence  against  a  competitive 
mechanism,  however,  and  competition  has  been  more  clearly  demonstrated 
for  some  other  pteridines. 

The  inhibition  by  pterin-6-aldehyde  is  quite  specific  although  it  must 
be  admitted  that  not  many  enzymes  have  been  tested.  Uricase,  glucose 
oxidase,  and  3-phosphoglyceraldehyde  dehydrogenase  are  not  affected,  but 
the  quinine  oxidase  of  liver  is  inhibited  (Kalckar  et  al.  1950;  Villela,  1963). 
Mouse  liver  guanase,  using  8-azaguanine  as  a  substrate,  is  inhibited  but 
not  potently  when  the  substrate  is  11  mM  and  30-min  preincubation  is 
allowed   (see   accompanying  tabulation)    (Shapiro   et  al.,    1952).    In   fact, 


Pterin-b-aldehyde  n,    n  ■   i   i   x- 

/n  Guanase  inhibition 
(mi/) 


1  32 

6  62 

11  80 

16  88 


xanthopterin  is  a  more  potent  and  more  rapidly  acting  inhibitor  (Dietrich 
and  Shapiro,  1953  b).  Pterin-6-aldehyde  is  not  carcinostatic  itself,  but 
potentiates  the  action  of  8-azaguanine  and  a  suppression  of  8-azaguanine 
destruction  was  claimed  as  the  mechanism,  although  the  relatively  low 
inhibitory  potency  coupled  with  the  low  doses  (20  mg/kg)  necessary  makes 
it  difficult  to  accept  this  explanation.  Byers  (1952)  investigated  the  effects 
of  pterin-6-aldehyde  injections  in  rats  (200  mg/kg  intraperitoneally)  on 
the  tissue  urate  levels  and  found  no  significant  changes,  which  might  in- 
dicate a  rapid  destruction  of  the  inhibitor  in  the  animal.  Daily  injections 
of  30  //g  pterin-6-aldehyde  in  chicks  also  does  not  alter  liver  xanthine  oxi- 
dase activity  despite  the  potent  inhibition  in  vitro  (Dietrich  et  al.,  1952). 
Other  pteridines,  although  less  potent  than  pterin-6-aldehyde,  are  nev- 


INHIBITION  OF  XANTHINE   OXIDASE  289 

ertheless  effective  inhibitors  of  xanthine  oxidase.  The  following  tabulation 
shows  the  inhibitions  after  20  min  incubation  at  pH  8.5  when  the  substrate 


Inhibitor 

Concentration 
(mM) 

%  Inhibition 

Pterin-6-aldehyde 

0.031 

100 

Pteroate 

0.033 

100 

Xanthopterin 

0.033 

72 

Isoxanthopterin 

0.066 

95 

Leucopterin 

0.032 

34 

7-methylxanthopterin 

0.077 

91 

6-MethyUsoxanthopteiin 

0.077 

89 

Xanthopterin-7-carboxylate 

0.050 

62 

Isoxanthopterin-6-carboxylate 

0.075 

23 

Pterin-6-carboxylate 

0.055 

15 

2,4-Diamino-6,7-dihydroxypteridine 

0.040 

76 

concentration  is  0.063  mM  (Hofstee,  1949).  The  potency  relative  to  pterin- 
6-aldehyde  is  not  seen  here  since  0.001  mM  inhibits  82%  under  these  con- 
ditions. From  the  K„,  for  xanthine  of  0.02  mM  (Hofstee,  1955),  a  K^  for 
pterin-6-aldehyde  of  5.1  X  10"^  mM  may  be  calculated,  which  is  higher 
than  was  obtained  by  Lowry  et  al.  (1949  b).  The  inhibition  by  xanthopterin 
appears  to  be  completely  competitive  (i^,„  for  xanthine  0.0053  milf ,  and  K, 
=  0.0016  mM),  and  xanthopterin-7-carboxylate  is  of  comparable  potency 
(Krebs  and  Norris,  1949).  Bovine  serum  xanthine  oxidase  is  inhibited  94% 
by  0.052  mM  pterin-6-aldehyde  but  only  3%  by  xanthopterin  at  a  com- 
parable concentration  (Villela  et  al.,  1956).  There  is  no  doubt  that  the  al- 
dehyde group  at  the  6-position  confers  a  strong  affinity  for  the  enzyme, 
since  when  it  is  reduced  to  a  CHgOH  group,  oxidized  to  a  COO"  group, 
or  altered  to  a  CHg  group,  the  inhibitory  activity  is  markedly  reduced  (Pe- 
tering and  Schmitt,  1950).  Pterin-6-aldehyde  is  bound  to  xanthine  oxidase 
approximately  4.9  kcal/mole  more  tightly  than  xanthopterin,  and  it  would 
appear  that  pterin  itself  is  bound  somewhat  less  tightly  than  xanthopterin. 
It  is  tempting  to  relate  the  augmenting  action  of  the  aldehyde  group  to 
the  fact  that  xanthine  oxidase  oxidizes  simple  aldehydes.  There  must  be 
a  site  on  the  enzyme  capable  of  reacting  with  aldehyde  groups,  and  it  is 
possible  that  the  pterin-6-aldehyde  is  bound  in  a  configuration  such  that 
the  aldehyde  group  interacts  in  this  manner.  Support  for  this  interpretation 
comes  from  the  observation  by  Lowry  et  al.  (1949  b)  that  pterin-6-aldehyde 
is  slowly  oxidized  to  pterin-6-carboxylate  by  xanthine  oxidase. 


290 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


CHOLINE  OXIDASE 

The  inhibitions  of  liver  choline  oxidase  reported  by  Wells  (1954)  (Table 
2-3)  recall  the  interferences  exerted  by  these  and  similar  analogs  on  cho- 
linesterase  and  tissue  acetylcholine  receptor  groups.  It  would  appear  that 
a  certain  critical  distance  between  the  N+  group  and  the  terminal  CH2OH 


Table  2-3 
Inhibition  of  Rat  Liver  Choline  Oxidase  by  Choline  Analogs" 


Inhibitor 

Relative 

rate  of 

oxidation 

Equimolar 

% 
inhibition 

Relative 
AF  of 

Series 

Ri 

R2 

R3 

binding 
(kcal/mole) 

Ethanolamine 

H 

H 

H 

0 

23 

0 

H 

Me 

Me 

0 

53 

-  0.81 

R 

H 

Et 

Et 

0 

25 

-  0.06 

R-N^^CHoCHgOH 
R^ 

Me 
Me 

Me 

Et 

Et 
Et 

79 
24 

- 

- 

Et 

Et 

Et 

0 

11 

+  0.55 

3- Aminopropanol- 1 

H 

H 

H 

0 

25 

-  0.06 

R 

Me 

Me 

Me 

5 

15 

+  0.33 

R— N^!—  CH2CH2CH2OH 
R 

H 

Me 

Et 
Et 

Et 
Et 

0 
8 

18 

7 

+  0.19 
+  0.88 

Et 

Et 

Et 

0 

14 

+  0.38 

1- Aminopropanol -2 

Me 

Me 

Me 

66 

- 

- 

R                         CH3 
R— N^— CH.— CH- OH 
R 

H 

Me 
Et 

Et 
Et 
Et 

Et 
Et 
Et 

0 

14 

4 

13 

6 

16 

+  0.43 
+  0.96 
+  0.29 

2-Amino-2-methylpropanol-l 

R              CH, 

R— N^— C-CH2OH 

/            1 
R              CH3 

H 

Me 
Et 

H 
Me 
Et 

H 
Me 

Et 

4 

105 

8 

64 
65 

-  1.09 

-  1.12 

2  -  Amino  -  2  -  methy  Ipropanediol 

R              CH, 

R— N^-C-CHjOH 

R               CH.OH 

H 
Me 

Et 

H 

Me 
Et 

H 

Me 

Et 

3 

41 

5 

41 
50 

-  0.51 

-  0.74 

"Experiments  done  with  rat  liver  homogenates.    Choline  and  all  analogs  at  37.5  mM.    The 
rates  of  oxidation  are  given  relative  to  that  for  choline  (100).    (From  Wells,  1954.) 


INHIBITION    OF    NITROGEN    FIXATION    BY    OTHER    GASES  291 

group  is  necessary  for  substrate  activity.  The  most  reactive  substrates 
are  dimethyl  or  trimethyl  compounds,  indicating  that  an  exact  fit  of  the 
cationic  head  is  necessary  to  place  the  hydroxyl  group  in  position  for 
oxidation.  Substitution  of  groups  on  the  C-1  atom  reduces  the  binding 
whereas  substitution  on  the  C-2  atom  increases  the  affinity  even  though 
the  groups  are  fairly  bulky.  The  simplest  interpretation  is  that  the  cationic 
head  anchors  the  molecules  in  position  so  that  the  CH2OH  group  can  react 
with  an  enzyme  group  on  the  opposite  side  of  a  hole  or  slit  in  the  protein. 
When  the  analogs  are  too  long  they  do  not  readily  fit  into  this  region, 
whereas  groups  protruding  from  C-2  interact  by  van  der  Waals'  forces 
with  the  walls  of  the  cavity.  A  three-point  attachment  of  the  cationic  head 
is  suggested  by  the  reduction  in  affinity  brought  about  by  altering  only 
one  of  the  R  groups,  this  perhaps  tilting  the  molecules  so  that  the  hydro- 
carbon chain  is  not  in  the  normal  direction.  The  differences  in  binding 
energies  between  these  analogs  are  rather  small  and  this  might  indicate 
that  dispersion  forces  are  mainly  involved,  but  it  may  also  be  that  changes 
in  the  electrostatic  interactions  (resulting  from  the  different  volumes  of 
the  R  groups,  for  example)  are  offset  by  opposite  changes  in  the  dispersion 
energy.  Since  choline  must  be  oxidized  to  betaine  before  it  can  serve  as 
a  methyl  donor,  it  is  interesting  that  Wells  demonstrated  the  inhibition 
of  methionine  synthesis  in  liver  homogenates  by  2-amino-2-methyl-l- 
propanol  and  its  triethyl  derivative.  Niemer  and  Kohler  (1957)  studied 
analogs  of  choline  in  which  one  of  the  methyl  groups  is  substituted  by  var- 
ious radicals  (e.g.,  — CHoCH.^OH,  — CHaCH^Br,  — CH2CH=CH2,  — CH2= 
=CH2,  and  —CH2COO-)  and  found  10-20%  inhibition  of  liver  choline 
oxidase  at  concentrations  approximately  equimolar  with  choline  (11.5  mM). 
None  of  these  analogs  is  a  potent  inhibitor,  confirming  the  importance  of 
fit  at  the  cationic  head. 


INHIBITION    OF   NITROGEN    FIXATION    BY   OTHER  GASES 

Some  simple  instances  of  competitive  interference  between  gases  in 
nitrogen  fixation  and  hydrogen  evolution  have  been  observed,  and  are 
reminiscent  of  the  suppression  of  hemoglobin  oxygenation  by  carbon  mon- 
oxide, nitric  oxide,  and  other  gases.  The  primary  product  of  nitrogen 
fixation  in  microorganisms  is  probably  ammonia,  and  the  enzyme  system 
responsible  for  this  is  generally  termed  nitrogenase.  A  few  examples  of 
inhibition  are  given  in  Table  2-4.  Most  of  these  have  been  shown  to  be 
strictly  competitive.  CO  and  NO  are  the  most  potent  inhibitors  while  H2 
and  NgO  are  relatively  weak.  O2  is  a  special  case  in  that  as  pOg  is  increased 
from  zero  the  nitrogen  fixation  accelerates,  but  above  a  certain  value, 
depending  on  the  organism,  the  rate  falls  off  (Burris,  1956).  Ethane,  neon, 
argon,  and  helium  have  no  significant  effects  (Molnar  et  al.,  1948). 


292 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


03 


O 


Zl- 

05 

05 

3 

-ti 

cS 

Tin 

oo 

3 

C 

C5 

05 

40 

^ 

o 

03 

o 

^ 

05 

cS 

lO 

73 

lO 

"e 

'e 

05 

c 

T3 

Oi 

^ 

o 

^ 

c« 

:^ 

«u 

<» 

e 

"C 

t4 

.52 

O 

03 

Sh 

C3 

Si) 

a 

^ 

^ 

U 

'C 

_a 

..^ 

>> 

;h 

tH 

(D 

Sh 

ci3 

O 
O 

3 

m 

s 

o 

K 

2       "^ 


m 


Zji 

^ 

O 

Oi 

•^ 

73 

6; 

3 

73 

— 

Si) 

cS 

i 

c 

ai 

0) 

m 

o 

O 

N      —    a       ja 


03 


pq      PQ 


s 

q 
d 

^"^ 

c3 

-S 

4^ 

d 

Y 

eS 

o3 

X 

SO 

cc 

g 
d 

S 
o 

m 

o 
d 

II 

a 

c6 

CO 

O 

d 

'3 

a 

o 
d 

II 

+3 

O 

d 

o 

d 

II 

C 
cS 
CD 

d 

II 

O 
? 

d 

o 
d 

II 

O 

O 

d 
II 

s 

O 

3 
O 

a 

CO 

d 

-2 
cS 

<M 

O 
O 

d 
II 
O 

II 

d 

51, 

]3 

.2 

d 

d 
II 

d" 

+2 
c 

a 

+3 
_0 

03 

O 

3 

_2 

a 

3 
o 

O 

4^ 

.^ 

-»^ 

> 

'■+3 

^^ 

-l-i 

g 

4^ 

4^ 

-p 

-ts 

> 

|S 

03 

IS 

03 

IS 

c3 

_2 

IS 
IS 

IS 

cS 

3 

.2 

^ 
» 

IS 

o 

d 
II 

2 

03 
> 
'-2 

II 

IS 

IS 

3 

,2 

IB 

o 
d 

II 

.2 

'+3 
IS 

.2 
IS 

IS 

|S 

> 

IS 

_2 

;^ 

IS 

IS 

.3 

"S 

IS 

> 
'-+3 

"-I3 

IS 

3 

"S 

[c 

'42 

■"" 

'S 

'■+3 

^~ 

_c 

'■S 

^^ 

-is! 
03 

[c 

'-3 

^" 

+3 

vO 

& 

-.o 

IS 

& 

& 

^  o 

& 

a 

■^ 

^ 

o~- 

^o 

o"- 

p 

TJ 

^o 

T3 

O"" 

-^o 

TJ 

3 

o^ 

o^ 

o 

o"-- 

s 

O 

^ 

j2 

o  - 

^ 

C 

O 

o^ 

3 

3 

O 

O 

C' 

«C 

o 

O 

c 

5 

ce 

lO 

o 

03 

O 

O 

O 

cS 

o 

o 

o 

CD 

O 

1 — 1 

o 

M 

O 

^ 

o 

o 

o 

-« 

'^ 

s 

e 

-2 

"S 

o 

3 

*, 

s 

s 

e 

•.!^ 

'« 

s 

s 

r2 

CO 

-2 

"S 

73 

e 

.s 

?5, 

§ 

O 

3 

S 

s 

3 

•«^ 

ts 

^ 

e 

£ 

o 

>. 

1 

o 

^    Ul       "% 


o 


o 


o 

o 

3 

3 

3 

3 

03 

o3 

<o 

<D 

X> 

XI 

>> 

>. 

O 

O 

02 

03 

INHIBITION    OF    NITROGEN    FIXATION    BY    OTHER    GASES  293 

Inhibition  of  Azotohacter  growth  by  N2O  occurs  if  N2  is  the  sole  source 
of  nitrogen,  but  does  not  if  ammonia  (Repaske  and  Wilson,  1952)  or  nitrate 
(Mozen  et  al.,  1955)  is  present,  indicating  that  the  site  of  the  block  is  prev- 
ious to  these  substances.  Likewise,  in  Clostridium  H^  inhibits  uptake  of 
N2  but  not  of  ammonia  (Hiai  et  al.,  1957).  These  inhibiting  gases  are  oc- 
casionally utilized.  For  example,  NgO  is  assimilated  by  Azotohacter,  although 
slowly,  and  this  is  inhibited  by  N,  and  Hg  (Burris,  1956).  It  is  quite  likely 
that  the  competition  in  all  these  cases,  with  the  possible  exception  of  Og, 
is  at  the  nitrogenase  active  site  binding  Ng.  Nitrogenase  is  a  metalloflavo- 
protein  containing  molybdenum  and  it  is  reasonable  that  these  gases  are 
bound  to  the  metal,  the  catalysis  of  Ng  reduction  being  of  similar  type  to 
those  mediated  by  various  inorganic  metal  preparations  (which  are  also 
inhibited  by  other  gases). 

It  will  be  necessary  before  discussing  mechanisms  of  inhibition  in  greater 
detail  to  consider  the  enzyme  hydrogenase,  which  has  recently  been  closely 
linked  to  nitrogen  fixation,  and  its  inhibitions.  This  enzyme  catalyzes  the 
reduction  of  some  unknown  primary  acceptor  by  molecular  Ho  and  the 
reduced  acceptor  then  transfers  the  hydrogen  atoms  to  other  acceptors, 
such  as  dyes,  NAD,  or  eventually  oxygen.  It  has  been  postulated  that  it 
may  in  some  instances  participate  in  the  reduction  of  Ng.  The  inhibition 
by  O2  is  primarily  due  to  oxygenation  of  the  enzyme: 

E  +  nO,  ^  E(02)„ 

and  this  inhibition  is  reversible  upon  removal  of  the  O2  by  dialysis  (Krasna 
and  Rittenberg,  1954;  Fisher  et  al.,  1954).  Prolonged  exposures  to  Og 
lead  to  progressive  inactivation  of  bacterial  hydrogenase  (Shug  et  al., 
1956).  Although  n  has  generally  been  assumed  to  be  1,  Atkinson  (1956) 
obtained  rather  complex  data  possibly  indicating  a  value  of  2  for  the 
Hydrogenomonas  facilis  enzyme.  The  hydrogenase-catalyzed  evolution  of 
H2  in  Rhodospirillum  rubrum  (Lindstrom  et  al,  1949)  and  soybean  root 
nodules  (Hoch  et  al,  1960)  is  inhibited  by  Ng.  Although  this  might  be  at- 
tributed in  part  to  a  diversion  of  the  flow  of  hydrogen  atoms  to  the  reduction 
of  N2,  Bregoff  and  Kamen  (1952)  observed  that  1  mole  of  N2  prevents 
the  release  of  several  moles  of  Hg.  One  of  the  difficulties  in  assuming  a  direct 
competition  between  Hg  and  Ng  for  the  hydrogenase  active  site  is  the 
fact  that,  despite  the  inhibition  of  Hg  evolution  by  N2,  the  exchange  reac- 
tion whereby  HD  is  formed  from  D2  and  a  hydrogen  donor  is  actually  accel- 
erated by  N2  (Hoch  et  al.,  1960).  The  Hj  inhibition  of  nitrogen  fixation 
was  previously  claimed  to  be  competitive,  but  Parker  and  Dilworth  (1963) 
found  that  Hg  causes  a  lag  in  the  N2  uptake  at  low  pN2,  whereas  in  cells 
of  Azotohacter  vinelandii  adapted  to  Hg  the  lag  is  abolished.  Taking  the 
lag  phase  into  account,  the  inhibition  is  not  competitive;  from  the  reciprocal 


294  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

plot  presented,  although  the  point  scatter  is  marked,  the  inhibition  might 
be  uncompetitive.  The  inhibition  of  nitrogen  fixation  in  soybean  root  nod- 
ules by  O2  is  complex,  due  to  both  plant  and  bacterial  components  of  the 
respiration,  but  is  competitive  when  the  O2  is  above  80%  (Bergersen, 
1962).  NgO  is  a  somewhat  more  potent  inhibitor  of  H2  evolution  than  is 
Ng,  but  the  most  potent  is  NO,  complete  and  irreversible  inhibition  being 
produced  by  concentrations  of  1%  or  greater  (Shug  et  al.,  1956).  The  hy- 
drogenase  in  cell-free  extracts  of  Proteus  vulgaris  is  inhibited  87%  by 
0.002%  or  0.00004  m.M  NO,  and  the  inhibition  at  these  low  concentrations 
is  partially  reversible  (Krasna  and  Rittenberg,  1954).  The  NO  is  neither 
oxidized  nor  reduced  by  the  enzyme. 

Although  the  configurations,  electronic  structures,  and  physical  prop- 
erties of  these  simple  gases  must  be  important  in  determining  the  inter- 
action with  nitrogenase  and  hydrogenase,  it  is  difficult  to  establish  correla- 
tions. Some  structural  and  physical  properties  that  might  relate  to  the 
interactions  of  these  molecules  are  given  in  Table  2-5.  Comparing  the  effects 
of  H2,  N2O,  ethane,  and  the  rare  gases  on  Azotobacter  nitrogen  fixation, 
Molnar  et  al.  (1948)  concluded  there  is  no  correlation  with  the  van  der 
Waals  constants  and  doubted  if  any  mechanism  could  be  based  on  physical 
properties  alone.  Wilson  and  Roberts  (1954)  postulated  that  N2O  is  inhi- 
bitory because  the  N — N  distance  is  close  to  that  in  Ng',  since  NgO  is  linear, 
the  oxygen  might  neither  interfere  nor  be  involved  in  the  binding.  If  the 
binding  is  to  metal  groups  on  the  enzymes,  the  degree  of  interaction  would 
depend  more  on  the  types  of  bond  possible  and  hence  on  the  electronic 
structures  of  both  the  gases  and  the  metal,  as  it  is  in  the  interactions  of 
O2,  CO,  and  NO  with  hemoglobin  and  cytochrome  oxidase. 

The  inhibition  of  nitrogen  fixation  by  O2  has  been  explained  as  a  compe- 
tition between  N2  and  0,  as  terminal  acceptors  for  electrons  originating 
in  the  oxidation  of  substrates  by  various  dehydrogenases,  nitrogen  fixation 


SH, »-  XH, 


being  considered  as  a  form  of  respiration  (Parker,  1954;  Parker  and  Scutt, 
1958,  1960).  It  is  likely  that  the  interrelationships  between  N,  and  Hg 
metabolism,  and  the  inhibitions  on  these  systems,  must  be  considered  in 
the  light  of  a  hydrogen  or  electron  pool  with  all  the  possible  pathways  for 
formation  and  utilization  of  hydrogen  atoms.  The  scheme  below,  modified 
from  Gest  et  al.  (1956),  may  serve  as  a  means  of  visualizing  some  of  these 
pathways.  Some  of  the  inhibitions  observed  are  due  to  competition  for 


INHIBITION    OF    NITROGEN    FIXATION    BY    OTHER    GASES 


295 


^ 


;0    :0: 


+  ^  +;?; 


;^     ^ 


^  o 


X 
O 


o 


p-( 


°<; 


S     I   T^   o   o      8   2 
§      I    ^    X    X     "^    X 


m 


s  ^ 


§   5 


fi      S-    -^     O) 


o     ;=; 


1     ^ 


296 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


Photolysis  of  water 


SH, 


(nitrogenase) 

N2  (forming  NH3) 

available  hydrogen  atoms,  while  others  are  based  on  direct  competition 
at  the  enzyme  active  sites. 


PHENOL  OXIDASES 

These  enzymes  usually  hydroxylate  monophenols  in  the  oriho  position 
and  further  oxidize  these  to  o-quinones.  They  have  often  been  named  ac- 
cording to  the  particular  substrate  chosen:  e.g.,  catechol  oxidase  (cate- 
cholase),  cresolase,  tyrosinase,  phenolase,  o-diphenol  oxidase,  or  polyphenol 
oxidase.  Different  specificities  are  observed  with  enzymes  from  various 
sources.  The  role  such  enzymes  play  in  tyrosine  metabolism  will  be  dis- 
cussed in  the  following  section. 

The  competitive  inhibition  of  potato  catechol  oxidase  by  resorcinol  was 
first  observed  by  Richter  (1934).  He  noticed  that  the  enzymes  from  various 
sources  exhibited  quite  different  susceptibilities  to  resorcinol,  those  from 
elder  (Sambucus)  and  lilac  {Syringa)  being  more  sensitive  than  the  potato 
enzyme,  and  those  from  mushroom  (Polyporus)  and  mealworm  {Tenebrio) 
less  sensitive.  The  respiration  of  apple  skin  is  inhibited  about  35%  by 
50  mM  resorcinol  and  thus  Hackney  (1948)  studied  an  extracted  catechol 
OH  OH  OH 


"OH  H3C        ^-^        OH 

Resorcinol  Orcinol 


COOH 


C^ 


CH-CH— COOH 


Nicotinic  acid 


Cinnamic  acid 


PHENOL   OXIDASES  297 

oxidase.  Although  the  inhibition  of  this  enzyme  is  reduced  by  increasing 
catechol  concentration,  her  data  do  not  correspond  to  pure  competitive 
inhibition  and  the  derived  K^  seems  to  be  in  error,  as  pointed  out  by  War- 
ner (1951).  A  potato  enzyme  oxidizing  p-cresol  is  inhibited  by  resorcinol, 
phloroglucinol,  and  orcinol  (see  tabulation),  the  inhibition  being  completely 

Inhibitor  (10  mM)         %  Inhibition  (p-cresol  =  10  mif) 


Resorcinol  78 

Phloroglucinol  43 

Orcinol  20 


reversible  by  dialysis  and  apparently  competitive  (Schneider  and  Schmidt, 
1959).  4-Chlororesorcinol  is  a  much  more  potent  inhibitor  of  the  potato 
enzyme,  oxidizing  either  catechol  or  p-cresol,  and  it  has  been  claimed  that 
the  initial  inhibition  is  competitive,  although  progressive  inactivation  oc- 
curs (Heymann  et  al.,  1954).  A  K^  of  0.024  tqM  was  calculated.  However, 
the  double  reciprocal  plots  seem  to  me  to  be  of  the  perfectly  noncompetitive 
type.  Bonner  and  Wildman  (1946)  postulated  that  the  bulk  of  spinach  leaf 
respiration  passes  through  a  polyphenol  oxidase.  p-Nitrophenol  is  a  rather 
potent  inhibitor  of  this  respiration,  1  voM  inhibiting  94%,  and  of  the  po- 
lyphenol oxidase,  whether  it  is  oxidizing  catechol  or  p-cresol.  On  the  other 
hand,  o-nitrophenol  is  only  a  weak  inhibitor  of  both.  It  was  felt  that  p-nitro- 
phenol  might  well  be  an  analog  of  naturally  occurring  substrates,  and  o- 
nitrophenol  an  analog  of  o-phenols  against  which  the  enzyme  is  inactive. 
It  is  interesting,  finally,  to  note  that  dihydroxymaleate  is  an  inhibitor  of 
catechol  oxidase  (Florkin  and  Duchateau-Bosson,  1939),  and  it  was  sug- 
gested that  the 

OH  OH 

— C  =  C— 

grouping  complexes  with  the  enzyme  in  a  manner  similar  to  the  analogous 
catechol  grouping.  Most  of  this  work  on  inhibiting  phenolic  compounds 
is  unsatisfactory  from  the  quantitative  standpoint  and  clear-cut  proofs 
of  competitive  inhibition  are  lacking.  Nevertheless,  effective  analog  in- 
hibition for  this  class  of  enzymes  has  heen  demonstrated. 

We  shall  now  turn  to  a  more  potent,  more  interesting,  and  more 
thoroughly  studied  type  of  inhibitor,  namely,  the  benzoates.  The  inhibition 
of  mushroom  catechol  oxidase  by  benzoate  itself  was  attributed  to  a  com- 
petition with  the  substrate  for  the  active  center,  although  no  direct  evidence 
for  this  was  adduced  (Ludwig  and  Nelson,  1939;  Gregg  and  Nelson,  1940), 
but  good  competitive  kinetics  for  the  inhibition  by  m-hydroxybenzoate 


298 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


were  observed  later  (Warner,  1951),  with  K^  2.5  mM  for  the  potato  en- 
zyme and  0.6  mM  for  the  mushroom  enzyme.  An  excellent  investigation 
by  Kuttner  and  Wagreich  (1953)  of  the  inhibition  of  a  catechol  oxidase 
from  Psalliota  campestris  provides  data  from  which  some  ideas  of  the  me- 
chanism may  be  obtained.  Some  of  the  inhibition  data  are  presented  in 
Table  2-6,  and  have  been  used  to  calculate  the  apparent  relative  binding 
energies.  However,  these  inhibitors  are  mostly  weak  acids  and  the  degree 

Table  2-6 

Inhibition  of  a  Phenol  Oxidase  from  Psalliota  campestris  with  Catechol  ( 1 .82  mM) 
AS  Substrate  (at  pH  5.2  and  25°) 


Inhibitor 

Concentration 

(mM) 

%  Inhibition 

Relative 
—  zli^  of  binding" 

(kcal/mole) 

Benzoate 

0.012 

50 

6.96 

p-Chlorobenzoate 

0.023 

50 

6.56 

p-Methylbenzoate 

0.04 

50 

6.23 

o-Chlorobenzoate 

0.21 

50 

5.21 

p-Methoxybenzoate 

0.26 

51 

5.10 

Nicotinate 

0.34 

48 

4.88 

o-Hydroxybenzoate 

0.38 

50 

4.84 

p-Nitrophenol 

0.48 

50 

4.70 

p-Hydroxybenzoate 

0.49 

50 

4.69 

irans-Cinnamate 

0.81 

62 

4.43 

Phenylacetate 

0.77 

50 

4.40 

o-Methylbenzoate 

1.0 

50 

4.25 

o-Nitrophenol 

1.1 

50 

4.19 

Benzoate  methyl  ester 

0.65 

30 

3.94 

Hydroquinone 

2.2 

50 

3.76 

p-Nitrobenzoate 

0.72 

19 

3.55 

Orcinol 

3.2 

50 

3.53 

Resorcinol 

4.5 

50 

3.32 

o-Nitrobenzoate 

0.72 

8 

2.95 

o-Methoxybenzoate 

8.0 

49 

2.94 

"  The  relative  energies  of  binding  to  the  enzyme  were  calculated  assuming  competi- 
tive inhibition  and  without  taking  into  account  the  state  of  ionization.  (Data  from 
Kuttner  and  Wagreich,   1953.) 


PHENOL   OXIDASES 


299 


of  ionization  at  the  experimental  pH  of  5.2  varies.  It  was  found  that  the 
inhibition  by  benzoate  decreases  with  rise  in  the  pH  (see  tabulation)  and 


pH 


%  Inhibition  by  benzoate  (0.0123  mil/) 


%  Un-ionized 


5.2 
5.8 
6.4 
7.0 


56 

28 


9.0 
2.5 
0.6 

0.2 


similar  results  were  obtained  with  some  substituted  benzoates.  This  was 
interpreted  to  mean  that  the  un-ionized  form  of  the  inhibitor  is  the  active 
one,  for  example  that  benzoic  acid  is  the  inhibitor  and  not  the  benzoate 
ion.  The  relative  binding  energies  have  been  recalculated  (Table  2-7)  on 
this  basis  and  a  somewhat  different  order  of  potency  is  obtained.  This  il- 

Table  2-7 

Inhibition  of  Mushroom  Phenol  Oxidase  Corrected  for  the  State  of  Ionization 

OF  THE  Inhibitors  " 


Corrected  relative 

Inhibitor 

P^a 

(HA)/(A,) 

—  AF  of  binding 
(kcal/mole) 

Benzoate 

4.203 

0.0914 

8.44 

o-Chlorobenzoate 

2.943 

0.0055 

8.39 

p-Chlorobenzoate 

3.978 

0.0565 

8.32 

o-Hydroxybenzoate 

3.001 

0.0063 

7.95 

p-Methylbenzoate 

4.371 

0.129 

7.47 

o-Nitrobenzoate 

2.173 

0.00094 

7.24 

p-Methoxybenzoate 

4.470 

0.157 

6.24 

o-Methylbenzoate 

3.909 

0.0487 

6.10 

p-Nitrobenzoate 

3.425 

0.0165 

6.07 

p-Hydroxybenzoate 

4.559 

0.186 

5.71 

Nicotinate 

4.854 

0.311 

5.59 

o-Methoxybenzoate 

4.092 

0.0724 

4.56 

°  Relative  binding  energy  corrected  on  the  basis  that  the  un-ionized  forms  of  the 
inhibitors  are  active.  The  ionization  constants  were  obtained  from  Dippy  (1939); 
inhibition  data  are  given  in  Table  2-6  (Kuttner  and  Wagreich,  1953.) 


300 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


lustrates  the  importance  of  taking  ionization  into  account  in  the  comparison 
of  inhibitors,  as  discussed  earlier  in  the  chapter.  It  must  be  emphasized 
that  the  absolute  values  of  the  binding  energy  are  meaningless;  it  is  only 
the  differences  between  the   —  JF  values  that  are  significant. 

Before  discussing  the  implications  of  these  results  we  will  examine  the 
data  reported  by  Krueger  (1955)  on  the  inhibition  of  a  mushroom  enzyme 
oxidizing  p-cresol  (Table  2-8).  Confirming  the  work  of  Kuttner  and  Wag- 


Table  2-8 
Inhibition  of  p-Cresol  Oxidation  by  Mushroom  Tyrosinase  ' 


Inhibitor 

Concentration 
(mif) 

%  Inhibition 

Relative 
—  AF  oi  binding 

pH  5.3 

pH7.0 

(kcal/mole) 

Benzoate 

4* 

90 

27 

5.20 

Oxalate 

4* 

83 

— 

4.78 

Cyclohexanecarboxylate 

16 

84 

— 

3.57 

Phenylacetate 

4 

54 

0 

3.51 

Fluoride 

40* 

72 

14 

2.99 

Butyrate 

40 

73 

10 

2.60 

Bromide 

40* 

56 

5 

2.56 

Iodide 

40* 

54 

8 

2.51 

Benzamide 

4* 

8 

— 

2.33 

Lactate 

20 

40 

0 

2.16 

Terephthalate 

4 

11 

— 

2.12 

Phthalate 

4 

10 

— 

2.05 

Chloride 

40 

45 

5 

1.85 

Formate 

4 

6 

— 

1.71 

20 

24 

— 

40* 

10 

0 

Acetate 

20 

21 

— 

1.68 

40 

42 

6 

Trimethylacetate 

40 

30 

4 

1.46 

Chloroacetate 

20 

16 

0 

1.39 

Benzenesulfonate 

40 

20 

0 

1.13 

Trichloroacetate 

20 

8 

— 

0.91 

"  Concentration  of  p-cresol  was  4.63  mM  except  where  indicated  by  asterisks,  in 
which  cases  it  was  9.26  n\M.  The  relative  binding  energies  were  calculated  for  pH  5.3 
and  were  not  corrected  for  ionization.  (Data  from  Krueger,  1955.) 


PHENOL   OXIDASES  301 

reich,  a  marked  decrease  in  the  inhibition  with  an  elevation  of  the  pH 
from  5.3  to  7  is  observed,  but  Krueger  interpreted  this  as  indicating  an 
ionizing  group  on  the  enzyme  with  a  pK^  around  6  and  possibly  an  imida- 
zole group.  One  reason  for  assuming  the  ionizing  group  to  be  on  the  enzyme 
is  the  pH  effect  on  inhibitions  by  the  inorganic  anions;  however,  the  in- 
hibition by  these  ions  may  be  through  a  different  mechanism  than  the 
benzoates,  and  indeed  Krueger  found  noncompetitive  kinetics  for  chloride. 
The  substrates  for  the  enzyme  are,  of  course,  un-ionized  and  this  might 
favor  the  concept  of  the  acid  form  of  the  inhibitors  being  active  and  the 
important  ionizing  group  on  the  inhibitors.  It  is  impossible  at  the  present 
time  to  decide  which  is  the  correct  interpretation  and  hence  the  signifi- 
cance of  the  pH  effects.  Ionizing  groups  on  both  enzyme  and  inhibitors 
might  also  be  considered.  It  should  be  added  that  the  following  were  found 
to  be  without  effect:  sulfate  (40  mM),  nitrate  (40  mM),  pyrophosphate  (40 
mM),  pyruvate  (20  mM),  succinate  (20  mM),  maleate  (20  mM),  fumarate 
(20  mM),  and  ethyl  benzoate  (4  mM). 

The  lack  of  information  on  the  exact  catalytic  mechanism  involved  in 
these  enzymes,  and  particularly  our  ignorance  of  the  state  and  role  of 
copper,  make  it  difficult  to  understand  the  binding  of  the  inhibitors.  There 
are  two  copper  ions  at  the  active  center  but  we  do  not  know  if  they  com- 
plex with  oxygen,  or  the  substrates,  or  both,  or  whether  one  copper  com- 
plexes with  oxygen  and  the  other  with  the  substrates.  Some  of  the  ionic 
inhibitions  might  be  due  to  the  formation  of  complexes  with  the  copper; 
such  complexes  might  be  more  difficult  to  form  at  higher  pH's  because  of 
competition  with  hydroxyl  ions.  The  strong  inhibition  by  oxalate  might 
be  due  to  chelation  of  the  copper,  but  it  is  odd  that  pyrophosphate  does 
not  inhibit  since  it  also  chelates  copper  weU.  It  is  also  surprising  that  sul- 
fate and  nitrate  do  not  inhibit  at  all,  unless  the  ionic  size  is  as  critical  as 
Krueger  believes. 

It  would  appear  that  inhibitory  activity  is  related  to  the  presence  of 
a  carboxyl  group  (excepting  the  inorganic  ions).  Benzoate  ester  and  benza- 
mide  are  bound  much  less  tightly  than  benzoate  (around  3  kcal/mole 
difference).  The  weak  action  of  benzenesuLfonate  might  be  explained  in 
three  ways;  (1)  if  the  un-ionized  form  of  the  acid  is  necessary  for  activity, 
there  would  be  less  in  the  case  of  benzenesulfonic  acid  than  with  benzoic 
acid,  (2)  the  sulfonate  group  might  be  too  bulky,  according  to  Krueger 
(although  it  is  certainly  not  much  larger  than  the  carboxyl  group),  and 
(3)  the  sulfonate  group  does  not  have  the  ability  to  form  bonds  with  the 
copper  or  hydrogen  bonds  with  an  enzyme  group.  Separation  of  the  car- 
boxyl from  the  benzene  ring  reduces  the  binding,  as  in  phenylacetate  or 
cinnamate.  The  second  requirement  for  potent  inhibitory  activity  is  a 
benzene  ring.  This  may  be  seen  by  comparing  acetate  and  phenylacetate, 
benzoate  and  nicotinate,   and  benzoate  and  cyclohexanecarboxylate;   in 


302  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

the  last  example  the  increased  binding  of  benzoate  may  be  due  to  its  great- 
er polarizability.  In  general  the  substitution  of  groups  on  the  benzene 
ring  reduces  the  affinity  for  the  enzyme.  This  may  be  due  to  steric  inter- 
ference with  the  approach  of  the  ring  or  to  inductive  effects  on  the  carboxyl 
group's  interaction  with  the  enzyme.  It  is  rather  odd  that  an  oriho  chlorine 
does  not  disturb  binding  much  while  an  ortho  methoxy  group  reduces  the 
binding  some  4  kcal/mole. 

The  forces  binding  the  substrates  and  inhibitors  to  these  enzymes  are 
thus  vague  at  the  present  time.  It  is  possible  that  hydrogen  bonds  between 
the  OH  or  COOH  groups  and  the  enzyme  are  important,  and  it  is  equally 
possible  that  bonds  to  the  copper  ions  are  involved.  One  might  conceive 
of  the  inhibitor's  COOH  group  reacting  with  either  the  two  copper  ions, 
or  with  a  copper  ion  and  a  vicinal  — NH —  group.  Copper  ions  are  able  to 
catalyze  the  oxidation  and  hydroxylation  of  phenols  nonenzymically,  and 
it  might  be  interesting  to  study  the  inhibition  of  such  reactions  by  some 
of  the  compounds  active  in  the  enzymic  reaction. 


TYROSINE   METABOLISM 

Many  interesting  and  practically  important  examples  of  analog  inhibi- 
tion are  to  be  found  in  the  general  field  of  amino  acid  metabolism,  and  we 
shall  begin  the  discussion  of  this  subject  by  considering  the  inhibitions  of 
the  various  pathways  of  tyrosine  metabolism.  Tyrosine  may  be  hydroxyl- 
ated  to  form  dihydroxyphenylalanine  (dopa),  oxidatively  deaminated  or 
transaminated  to  form  p-hydroxyphenylpyruvate,  decarboxylated  to  form 
tyramine,  or  activated  prior  to  incorporation  into  proteins;  inhibition 
of  all  of  these  reactions  by  analogs  has  been  reported.  The  scheme  on  page 
303  indicates  the  major  pathways  of  tyrosine  metabolism.  Interference 
with  these  reactions  might  be  expected  to  bring  about  physiological  changes 
due  to  the  acceleration  or  suppression  of  active  amine  synthesis,  and  also 
to  affect  melanin  formation. 

Tyrosinase  (Phenol  Oxidase) 

These  enzymes  hydroxylate  tyrosine  in  the  oriho  position  to  form  dopa 
and  further  oxidize  dopa  to  dopa-quinone.  The  enzymes  discussed  in 
the  previous  section  generally  possess  this  activity.  However,  mammalian 
tyrosinases  are  much  more  specific  than  the  plant  enzymes  and  oxidize 
tyrosine  and  dopa  more  rapidly  than  other  phenols.  Inhibitions  of  the  en- 
zymes with  tyrosine  as  the  substrate  will  be  considered  in  this  section, 
but  the  results  with  p-cresol  or  catechol  as  substrate  are  probably  generally 
applicable  to  tyrosinase  activity.  The  high  inhibitory  potency  of  4-chloro- 
resorcinol  on  potato  tyrosinase,  as  determined  by  the  rate  of  formation  of 


TYROSINE   METABOLISM 


303 


n    ^         -2  — ■ 
2    rt    tn    D    M    G 
•5  ^    O  -rl  O    5 


oS 


2    >>  e 


a 


304  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

melanin  from  tyrosine,  was  noted  by  KuU  et  al.  (1954),  for  example  (see 
following  tabulation).  Other  substitutions  in  the  4-position  usually  reduce 


Lowest 

Inhibitor 

inhibitory  concentration 

(mil/) 

4-Chlororesorcinol 

0.000069 

Resorcinol  monobenzoate 

0.0047 

m-Aminophenol 

0.0092 

4-«.-Hexylresorcinol 

0.103 

Naphthoresorcinol 

0.125 

Orcinol 

0.141 

j)-Aminophenol 

0.183 

Resorcinol 

0.183 

Phloroglucino! 

6.17 

2-Nitroresorcinol 

65 

or  abolish  the  inhibitory  activity.  The  formation  of  melanin  involves  several 
steps  and  the  inhibitions  observed  are  not  necessarily  entirely  on  tyrosinase. 
Hydroquinone  was  found  to  be  a  weak  inhibitor  but  the  monobenzyl 
ether  of  hydroquinone  is  as  potent  as  resorcinol.  This  latter  substance, 


"°A\  /r°-'''^-A\  // 


Monobenzone  (Benoquin) 


known  also  as  monobenzone  (Benoquin),  is  an  inhibitor  of  melanin  for- 
mation in  the  skin  when  applied  topically,  can  produce  leucoderma  in 
Negroes,  and  is  used  in  various  conditions  of  melanosis.  Some  have  thought 
that  it  releases  hydroquinone  after  penetration  into  the  skin  but  this  is 
questionable  in  view  of  its  own  inhibitory  activity. 

An  active  tyrosinase  occurs  in  the  Hardin-Passey  mouse  melanoma  and 
is  probably  responsible  for  pigment  formation.  It  is  competitively  inhib- 
ited by  various  tyrosine  analogs  (Lerner  et  al.,  1951).  The  values  of  K^ 
shown  in  the  tabulation  were  calculated  on  the  basis  of  a  K„,  of  0.60  mM 


Inhibitor  J^,  (mM) 

iV-Acetyl-L-tyrosine  0.140 

iV-Formyl-L-tyrosine  0.177 

3-Amino-L-tyrosine  0.314 

3-Fluoro-L-tyrosine  1.25 


TYROSINE   METABOLISM  305 

obtained  from  the  reciprocal  plots.  3-Nitro-L-tyrosine  and  O-methyl-L- 
tyrosine  are  not  inhibitory.  The  effects  of  3-substitution  may  be  mediated 
through  inductive  effects  on  the  4-OH  group  and  its  interaction  with  the 
enzyme,  whereas  iV-substitution  must  lead  to  an  altered  position  of  binding 
to  prevent  oxidation.  L-Phenylalanine  and  phenylpjTuvate  inhibit  com- 
petitively the  tyrosinase  from  melanoma,  inhibit  the  incorporation  of  ty- 
rosine-C^*  into  melanin,  and  depress  the  respiration  of  tumor  tissue  with 
tyrosine  as  the  substrate  (Boylen  and  Quastel,  1962).  The  high  concentra- 
tions of  these  inhibitors  in  phenylketonuria  might  be  responsible  for  the 
reduced  pigment  formation  in  these  individuals. 

Tyrosine  :  a-Ketoglutarate  Transaminase 

The  effects  of  numerous  analogs  on  the  formation  of  p-hydroxyphenyl- 
pyruvate  from  tyrosine  and  a-ketoglutarate  by  a  dog  liver  transaminase 
were  reported  by  Canellakis  and  Cohen  (1956  b);  some  of  the  results  are 
given  in  Table  2-9.  Certain  of  these  analogs  are  transaminated  (e.g.,  the 
3-substituted  tyrosines)  and  the  inhibitory  activity  varies  inversely  with 
their  abilities  to  act  as  substrates.  The  rapid  transamination  of  3-fluoro- 
tyrosine  may  partly  explain  its  toxic  effects  and  inhibition  of  growth, 
since  fluorofumarate  or  fluoroacetoacetate  may  be  formed.  Comparing  the 
hydroxyl-substituted  phenylalanines,  it  is  seen  that  a  m-  or  p-hydroxyl 
is  necessary  for  tight  binding,  the  contribution  to  the  binding  energy 
being  over  2  kcal/mole.  The  if,,;  for  L-tyrosine  is  0.71  mM,  so  that  its  rel- 
ative binding  energy  is  at  least  —  4.47  kcal/mole,  and  that  of  m-hydroxy- 
DL-phenylalanine  is  at  least  —  4.75  kcal/mole  (—  5.18  kcal/mole  if  only 
the  L-isomer  is  active),  which  may  be  compared  with  the  —  2.55  kcal/ 
mole  for  L-phenylalanine.  Hydroxyl  groups  in  the  o-positions,  on  the  other 
hand,  do  not  augment  binding  very  much.  The  tighter  the  binding  between 
a  basic  hydroxyl  and  an  acidic  enzyme  group,  the  greater  the  inhibitory 
activity;  ring  substituents  modify  the  electronic  character  or  basicity  of 
the  phenolic  group.  A  carboxylate  group  is  necessary  for  strong  binding, 
as  may  be  seen  by  comparing  tjTosine  with  tyramine,  and  p-hydroxyben- 
zoate  with  p-cresol.  The  a-amino  group  may  also  be  involved  in  the  binding 
(—  AF  for  p-hydroxyphenylacetate  is  2.87  kcal/mole),  but  lacking  data  on 
p-hydroxyphenylproprionate  it  is  not  possible  to  evaluate  this  accurately. 
The  strong  inhibition  by  epinephrine  is  rather  surprising  in  the  light  of 
the  absence  of  a  carboxylate  group,  but  the  /?-hydroxyl  or  iV-methyl  group 
may  contribute  to  make  up  for  this  deficiency.  A  similar  study  on  the  rat 
liver  enzyme  has  been  reported  by  Jacoby  and  La  Du  (1964). 

p-Hydroxyphenylpyruvate   Oxidase 

The  further  oxidation  of  the  product  of  tyrosine  transamination  is  ca- 
talyzed by  an  enzyme  from  dog  liver  and  is  inhibited  markedly  by  phenyl- 


306 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


Table  2-9 
Inhibition  of  Tyrosine  :  a-KETOGLUTARATE  Transaminase  from  Dog  Liver  " 


Relative 

Inhibitor 

(I)/(S) 

%  Inhibition 

—  zJi^  of  binding 
(kcal/mole) 

3-Araino-DL-tyTosine 

1 

85 

6.00 

3,4-Dihydroxy-L-phenylalanine 

2 

70 

5.03 

Epinephrine 

2 

70 

5.03 

m-Hydroxy-DL-phenylalanine 

2 

60 

4.75 

2,5-Dihydroxy-L-phenylalanine 

2 

35 

4.13 

3-Fluoro-DL-tyrosine 

2 

20 

3.65 

3,5-Dibromo-L-tyrosine 

2 

10 

3.15 

o-Hydroxy-DL-phenylalanine 

2 

7 

2.91 

p-Hydroxybenzoate 

10 

26 

2.87 

p-Hydroxyphenylacetate 

10 

26 

2.87 

p-Aminobenzoate 

10 

21 

2.70 

L-Tryptophan 

10 

21 

2.70 

p-Nitro-DL-phenylalanine 

2 

5 

2.69 

L-Phenylalanine 

2 

4 

2.55 

o-Hydroxybenzoate  (saHcylate) 

10 

17 

2.53 

p-Nitrobenzoate 

10 

15 

2.44 

3,5-Diiodo-L-tyrosine 

2 

3 

2.36 

D-Tyrosine 

2 

3 

2.36 

j)-Cresol 

10 

13 

2.34 

Tjrramine 

10 

9 

2.09 

"  Relative  binding  energies  were  calculated  on  the  basis  of  competitive  inhibition 
with  K^  =  0.71  mM  for  L-tyrosine.  No  correction  for  ionization  was  made.  It  should 
be  noted  that  the  different  isomers  of  the  DL-compounds  may  have  different  activities, 
so  that  the  binding  energy  of  the  active  form  should  be  increased.  (Data  from  Canel- 
lakis  and  Cohen,  1956  b.) 

pyruvate,  essentially  complete  inhibition  occurring  with  0.4  mM  in  the 
presence  of  2  mM  substrate  (Zannoni  and  La  Du,  1959).  The  inhibition  is 
negligible  for  the  first  5  min  but  then  increases  to  become  complete  at 
around  20  min.  This  might  be  due  to  protection  by  the  substrate  and  pre- 
incubation experiments  would  have  been  informative.  An  equally  good 
inhibitor  is  m-hydroxyphenylpyruvate  but  phenylacetate,  2,5-dihydrox- 
yphenylpyruvate,  p-hydroxyphenyllactate,  p-hydroxybenzoate,  and  ho- 
mogentisate  are  not  inhibitory. 

Other  Pathways  of  Tyrosine  Metabolism 

Little  is  known  about  the  effects  of  analogs  on  tyrosine  decarboxyla- 
tion, although  this  might  well  be  an  important  site  to  block  if  one  wished 


TYROSINE    METABOLISM 


307 


to  reduce  the  tissue  tyramine  concentration.  Mardashev  and  Semina 
(1961)  found  that  the  tyrosine  decarboxylase  from  Streptococcus  fecalis  is 
inhibited  20%  by  8.3  mM  cysteine  and  25%  by  8.3  mM  homocysteine 
when  substrate  concentration  is  2.8  mM.  This  is  a  general  phenomenon  seen 
with  several  amino  acid  decarboxylases  and  is  presumably  due  to  the  for- 
mation of  complexes  of  the  inhibiting  amino  acids  with  pyridoxal  phosphate. 
The   tyrosine-activating   enzyme    of   pig   pancreas   catalyzes   the   first 

Amino  acid  +  ATP  +  E   ->  E-amino  acyl-AMP  +  PP 

step  in  the  incorporation  of  amino  acids  into  proteins.  This  also  would 
be  a  very  interesting  step  to  investigate  from  the  standpoint  of  analog 
inhibition,  but  our  present  information  is  meager.  Schweet  and  Allen 
(1958)  found  that  3-fluoro-lL-tyrosine  activated  50%  the  rate  of  L-ty- 
rosine  and  does  not  inhibit.  L-tyrosinamide  inhibits  weakly  but  tyramine 
inhibits  the  phosphate  exchange  reaction  80%  at  0.2  mM,  this  inhibition 
being  reduced  by  increase  in  substrate  concentration.  The  importance  of 
the  p-hydroxyl  group  is  indicated  by  the  fact  that  no  inhibition  is  seen 
with  phenylethylamine. 


Dihydroxyphenylalanine  (Dopa)  Decarboxylase 

This  enzyme  is  on  the  pathway  leading  from  tyrosine  to  the  important 
catecholamines,  epinephrine  and  norepinephrine,  and  lately  has  been  the 
subject  of  much  investigation  because  of  the  possible  clinical  applications 
of  producing  a  selective  block  at  this  step.  Inhibitors  have  indeed  been 
found  which  are  effective  in  vivo,  reduce  amine  formation,  and  lower  the 
blood  pressure.  The  sequence  of  reactions  for  the  formation  of  amines 
from  dopa  is  the  following: 


HO 


HO 


NH3 
-CH— COO' 


dopa 


decarboxylase 


HO 


CH^  CH,— NH, 


Dopa 


Dopamine 


dopamine  /3-oxidase 


OH 

1  H 

CH— CH,— NH, 


Epinephrine 


308 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


A  block  of  dopa  decarboxylase  would  thus  decrease  the  rate  of  formation 
of  these  three  physiologically  active  amines  in  the  tissues. 

Of  historical  interest  are  the  following  observations  on  the  analog  inhi- 
bition of  this  enzyme:  dopamine  (Blaschko,  1942),  epinephrine  (Schapira, 
1946),  various  aromatic  amines  (tryptamine,  tyramine,  phenylethylamine, 
etc.)  (Polonovski  et  al.,  1946),  and  the  various  hydroxy,  methoxy,  and 
dimethoxy  derivatives  of  phenylethylamine,  the  dimethoxy  analogs  being 
the  most  potent  (Gonnard,  1950). 

These  studies  were  extended  in  important  ways  by  Sourkes  (1954), 
who  discovered  the  potent  inhibiting  activity  of  certain  a-methylphenyl- 
alanines  (Table  2-10).  Especially  inhibitory  is  a-methyldopa  and  this  sub- 


Table  2-10 
Inhibition  of  Dopa  Decarboxylase  from  Pig  Kidney  Cortex  " 


Inhibitor 

Concentration 

(mM) 

%  Inhibition 

Relative 

—  /iF  oi  binding 

(kcal/mole) 

a-Methyldopa 

0.01 

22 

5.94 

0.1 

71 

0.5 

98 

a-Methyl-3-hydroxy-PA 

0.05 

45 

5.57 

0.5 

74 

5 

95 

a-Methyl-3-hydroxy-4- 

4.3 

92 

4.86 

methoxy-PA 

a-Methyl-3,4-dimethoxy-PA 

0.01 

16 

4.65 

0.1 

20 

iV-Acetyl-3,4-dimethoxy-PA 

3.6 

44 

3.32 

A'^-Methyldopa 

0.8 

11 

3.12 

1.7 

30 

Diiodotyrosine 

2.3 

15 

2.68 

3,4-Dimethoxy-PA 

5 

25 

2.59 

2,4-Dimethoxy-PA 

5 

22 

2.49 

3-Methoxy-4-hydroxy-PA 

5 

18 

2.33 

a-Methyl-3-methoxy-PA 

5 

18 

2.33 

a-Methyltyrosine 

5 

16 

2.25 

a-Methyl-PA 

2 

6 

2.13 

A'^-Methyl-3-methoxy-4- 

1.6 

0 

— 

hydroxy-PA 

"  Concentration  of  DL-dopa  4  mM,  pH  6.8,  preincubation  with  inhibitor  15  min. 
PA  —  phenylalanine,  and  dopa  =  3,4-dihydroxyphenylalanine.  Relative  binding  ener- 
gies calculated  on  the  basis  of  competitive  inhibition,  which  may  not  be  strictly  true; 
in  any  event,  these  values  give  a  better  indication  of  the  relative  inhibitory  potency 
than  the  per  cent  inhibition  at  different  concentrations.   (Data  from  Sourkes,  1954.) 


TYKOSINE    METABOLISM  309 

stance  has  been  thoroughly  studied  biochemically  and  pharmacologically 
during  recent  years.  It  is  interesting  that  these  analogs  have  very  little 
inhibitory  activity  toward  tyrosine  decarboxylase  and  that  a-methyl- 
tyrosine  does  not  inhibit  dopa  decarboxylase  strongly,  both  facts  pointing 
to  the  importance  of  the  3-hydroxyl  group  in  the  binding  to  the  enzyme. 
This  is  also  seen  by  comparing  or-methylphenylalanine  and  its  hydroxylated 
derivatives:  The  addition  of  a  4-hydroxyl  has  little  effect,  whereas  a 
3-hydroxyl  increases  the  binding  energy  over  3  kcal/mole.  A  3-methoxy 
group  seems  to  be  ineffective. 

The  inhibition  by  a-methyldopa  was  shown  to  be  pseudoirreversible 
by  varying  the  enzyme  concentration  and  using  the  graphic  procedure  of 
Ackermann  and  Potter  (1949).  At  concentrations  of  0.01-0.03  mM,  the 
inhibition  being  15-25%,  the  behavior  is  fairly  reversible,  but  at  concen- 
trations of  0.1  mM  or  above  there  is  marked  nonlinearity  of  the  curves. 
As  pointed  out  by  Sourkes,  these  data  indicate  merely  that  K^  is  low  and 
the  affinity  for  the  enzyme  is  high.  The  binding  might  be  to  the  apoenzyme, 
to  a  great  extent  through  the  phenolic  groups,  or  the  inhibition  could  be 
the  result  of  reaction  with  pyridoxal  phosphate.  The  former  mechanism 
was  favored  by  Sourkes  on  the  basis  of  the  following  evidence  against  a 
reaction  with  the  coenzyme.  (1)  The  inhibition  is  reversible  by  dialysis. 
(2)  The  rate  of  nonenzymic  reaction  of  a-methyldopa  with  pyridoxal 
phosphate  is  too  slow  at  inhibiting  concentrations  to  be  significant.  (3)  In- 
crease in  pyridoxal  phosphate  concentration  does  not  alter  the  inhibition 
significantly.  (4)  Analysis  for  pyridoxal  phosphate  at  the  end  of  inhibition 
experiments  showed  no  loss.  (5)  Tyrosine  decarboxylase  is  also  a  pyridoxal 
phosphate  enzyme  and  is  not  inhibited.  None  of  this  evidence  is  completely 
conclusive  and  it  is  possible  that  a-methyldopa  can  form  a  reversible 
complex  with  pyridoxal  phosphate  on  the  enzyme  surface,  so  that  increase 
in  coenzyme  concentration  would  not  be  effective  and  analysis  for  total 
coenzyme  would  not  detect  the  small  amount  combined.  5-Hydroxytryp- 
tophan  decarboxylase  is  also  potently  inhibited  by  a-methyldopa  (it  is 
possible  that  the  decarboxylases  for  dopa,  5-hydroxytryptophan,  trypto- 
phan, tyrosine,  and  phenylalanine  in  mammalian  tissues  represent  a  single 
enzyme)  and  S.  E.  Smith  (1960  a)  has  investigated  the  mechanism,  using 
the  mouse  brain  enzyme.  Plots  of  1/(S)  against  l/v  showed  pure  competitive 
inhibition  with  respect  to  substrate  at  higher  coenzyme  concentrations 
(above  0.01  mM),  but  at  low  coenzjTne  concentrations  the  inhibition  be- 
comes noncompetitive  with  substrate.  In  contrast  to  dopa  decarboxylase, 
increase  of  coenzyme  concentration  leads  to  a  reduction  in  the  inhibition 
(Fig.  2-3).  Smith  inclines  to  a  coenzyme  inactivation  mechanism  but  admits 
that  the  inhibition  is  incompletely  explained.  If  a-methyldopa  forms  a 
complex  with  pyridoxal  phosphate  on  the  enzyme  surface,  which  it  can  do 
because  it  is  decarboxylated  slowly,  it  might  be  considered  to  be  an  in- 


310 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


hibitor  which  enters  into"  the  catalytic  sequence  of  reactions  but  is  not 
able  to  complete  the  process  readily.  This  mechanism  is  also  suggested  by 
the  results  of  Lovenberg  et  al.  (1963)  on  kidney  aromatic  amino  acid  de- 
carboxylase, using  tryptophan  as  the  substrate.  When  a-methyldopa  is 
preincubated  with  the  enzyme  in  the  absence  of  pyridoxal-P  the  inhibition 
is  noncompetitive  and  potent,  but  when  pyridoxal-P  is  added  during  the 
preincubation  period  the  inhibition  is  competitive  with  respect  to  substrate 


0.001 

■  METHYL  -  OOPA 

Fig.  2-3.  Inhibition  of  mouse  brain  5-hydroxytryptophan 
decarboxylase  by  a-methyldopa.  The  sohd  curves  show 
the  rate  of  formation  of  serotonin  in  //g/g/hr,  and  the 
dashed  curves  show  the  fractional  inhibition.  The  curves 
X-X-X-X  show  the  results  without  addition  of  pyridoxal-P, 
and  the  curves  0-0-0-0  show  the  results  after  addition 
of  0.008  mM  pyridoxal-P.   (From  S.  E.  Smith,   1960  a.) 


and  is  less  potent.  If  no  preincubation  is  done  and  a-methyl  dopa  is  added 
with  the  substrate,  competitive  inhibition  is  observed.  The  data  suggest 
that  a-methyldopa  interacts  specifically  with  the  enzyme-pyridoxal-P  com- 
plex (the  enzyme  as  isolated  contains  tightly  bound  pyridoxal-P),  and  the 
protection  or  reversal  of  the  inhibition  by  exogenous  pyridoxal-P  may  be 
due  to  the  reactivation  of  the  enzyme-bound  coenzyme.  The  reaction  of 
the  a-methyldopa  with  pyridoxal-P  may  involve  the  cyclization  of  a 
Schiff  base  (Mackay  and  Shepherd,  1962). 

Another  comprehensive  study  of  dopa  decarboxylase  was  made  by  Hart- 
man  et  al.  (1955),  who  determined  the  inhibitory  activities  of  some  200 
compounds.  The  results,  some  of  which  are  presented  in  Table  2-11,  enable 


TYROSINE    METABOLISM 


311 


Table  2-11 
Inhibition  of  Dopa  Decarboxylase  from  Pig  Kidney  Cortex" 


Inhibitor 


Concen- 
tration 
(mM) 


% 
Inhibition 


Relative 

-  AF  of 

binding 

(kcal/mole) 


Cinnamates 


O 


CH=CH— COO' 


3-Mercapto- 

0.1 

78 

6.90 

3,4-Dihydroxy-  (ethyl  ester) 

0.3 

90 

6.78 

2, 6-Dihydroxy- 

0.2 

84 

6.70 

3, 4-Dihydroxy- 

0.4 

74 

5.90 

2-Hydroxy-3,  5-dibromo- 

0.4 

65 

5.63 

3-Hydroxy-  (methyl  ester) 

0.4 

60 

5.50 

2-Hydroxy-  (methyl  ketone) 

0.2 

31 

5.19 

3-Hydroxy-  (methyl  ketone) 

0.2 

30 

5.15 

3-Hydroxy- 

0.4 

37 

4.92 

a-Methyl-3-hydroxy- 

0.4 

31 

4.75 

3-Hydroxy-6-sulfonate- 

0.4 

31 

4.75 

a-Ethyl-3-hydroxy- 

0.4 

27 

4.64 

3 -Hydroxy-  (amide) 

0.4 

19 

4.35 

2-Hydroxy- 

2 

52 

4.30 

2,  4-Dichloro- 

2 

41 

4.03 

2-Chloro- 

2 

36 

3.91 

4-Hydroxy- 

0.4 

7 

3.65 

3-Nitro- 

2 

15 

3.19 

3-Amino- 

2 

10 

2.91 

2, 4-Dihydroxy- 

2 

9 

2.83 

2,  3-Dimethoxy- 

2 

8 

2.75 

4-Chloro- 

2 

8 

2.75 

Unsubstituted 

2 

0 

~ 

Hydrocinnamates       <(        \—  CH2CH2—  COO" 

V      / 

3-Hydroxy-2,  4,  6-triiodo- 

0.2 

90 

7.04 

a-Methyl-3-hydroxy-2,  4,  6-triiodo- 

0.4 

91 

6.68 

3,  4-Dihydroxy- 

0.44 

63 

5.52 

a-Methyldopa 

0.2 

37 

5.35 

4-Hydroxy- 

2 

9 

2.83 

3-Hydroxy- 

2 

0 

- 

312  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Table  2-11   (continued) 


Inhibitor 


Concen- 
tration 
(mA'i) 


% 
Inhibition 


Relative 

-  AF  of 

binding 

(kcal/mole) 


d-Hydroxy- 
Unsubstituted 

Phenylacetates 

3, 4-Dihydroxy- 
2, 5-Dihydroxy- 
3,  4-Dimethoxy- 
3 -Hydroxy - 
Unsubstituted 


CH,— COO" 


2 

88 

5.49 

2 

28 

3.68 

2 

18 

3.32 

2 

11 

2.97 

2 

0 

_ 

Phenylglycines 

3,  4-Dihydroxy- 
3-Hydroxy- 
4-Hydroxy- 
Unsubstituted 


=\  NH, 

CH 
\ 
COO 


2 

16 

2 

0 

2 

0 

2 

0 

3.24 


Phenylpyruvates 

3,  4-Dihydroxy- 

3-Hydroxy- 

4-Hydroxy- 

2, 5-Dihydroxy- 

Unsubstituted 


CH5— CO— COO' 


0.2 

1 

2 

2 

2 


60 
72 
60 
60 
60 


5.92 
5.26 
4.50 
4.50 
4.50 


Benzoates  (\       n — -  COO" 

2-Hydroxy-3,  5-diiodo- 
2-Hydroxy-3,  5-dibromo- 
3,  5-Dibromo- 
3,4,  5-Trihydroxy- 
3-Hydroxy-4,  6-dibromo- 
2 -Hydroxy- 5-amino- 
2-Hydroxy-4-nitro- 


0.2 

47 

5.60 

0.2 

33 

5.24 

0.2 

10 

4.33 

2 

42 

4.07 

2 

36 

3.91 

2 

28 

3.68 

2 

26 

3.62 

TYROSINE    METABOLISM 


313 


Table  2-11    (continued) 


Inhibitor 

Concen- 
tration 
(niiW) 

% 
Inhibition 

Relative 

-  AF  of 

binding 

(kcal  mole) 

2,  5-Dihydroxy- 

2 

21 

3.44 

2,  4,  6-Trihydroxy- 

2 

16 

3.24 

3,  4-Dihydroxy- 

2 

0 

- 

2. 4-Dihydroxy- 

2 

0 

- 

Miscellaneous 

5- (3.  4-Dihydroxycinnamoyl)salicylate 

0.002 

87 

9.69 

HO 


COO" 


CH=CH-CO 


OH 


°  Concentration  of  dopa  2  mM  and  pH  6.8.    (Data  from  Hartman  et  al.,   1955. ) 

one  to  speculate  further  about  the  nature  of  the  binding  to  the  active 
center.  Several  inhibitors  more  potent  than  a-methyldopa  were  found. 
The  basic  structure  for  inhibition  was  written  as: 


HO 


(HO) 


O 

I        I       II 

c=c-c— X 


where  X  is  OH,  0-alkyl,  alkyl,  or  aryl.  It  is  rather  surprising  that  the 
negatively  charged  carboxylate  group  is  not  necessary,  esters  and  amides 
being  as  potent  as  the  acids,  and  it  may  be  that  the  CO  group  is  critical. 
The  positively  charged  amino  group  is  also  not  necessary,  since  3,4-dihy- 
droxyhydrocinnamate  is  a  good  inhibitor,  and  this  would  make  it  likely 
that  the  binding  of  the  inhibitors  is  not  too  much  dependent  on  pyridoxal 
phosphate.  The  importance  of  the  3-  and  4-hydroxyls  is  again  evident  and 
all  the  potent  inhibitors  have  phenolic  groups;  apparently  only  the  sulf- 
hydryl  group  can  replace  the  hydroxyl.  Halogens  have  the  ability  to  aug- 
ment binding  when  they  are  the  only  substituents  but  particularly  when 
a  hydroxyl  group  is  also  present;  3,5-dibromobenzoate  and  2,4-dichloro- 
cinnamate  are  bound  reasonably  well  (at  least  2  kcal/mole  more  than  the 
unsubstituted  compounds).  The  only  unsubstituted  inhibitor  is  phenyl- 
pyruvate,  which  must  be  significant,  although  of  what  is  not  clear.  The 
interaction  of  the  side  chain  must  be  complex  and  involve  different  types 
of  forces.  If  one  compares  all  the  3,4-dihydroxy  derivatives,  it  is  seen  that 


314  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

inhibitory  activity  increases  with  the  length  or  bulk  of  the  side  chain. 
Also  it  may  be  noted  that  the  Hnear  cinnamate  derivatives  are  generally 
more  potent  than  the  hydrocinnamates.  One  must  conclude  that  the  most 
important  binding  groups  are  the  hydroxyls,  the  ring,  the  side-chain  car- 
bonyl,  and  any  more  terminal  groups,  which  just  about  includes  all  of  the 
molecule.  The  specificity  of  these  inhibitors  may  well  be  quite  high,  since 
3-hydroxycinnamate  and  3,4-dihydroxycinnamate  (caffeate)  were  tested 
on  tyrosine  decarboxylase,  glutamate  decarboxylase,  histidine  decarbo- 
xylase, and  succinate  dehydrogenase,  and  found  to  be  without  effect. 
Tyrosinase  is  inhibited  somewhat,  especially  by  caffeate. 

There  is  evidence  in  patients  with  phenylpyruvic  oligophrenia  of  a  dis- 
turbance in  tyrosine  metabolism  (hypopigmentation),  and  it  is  possible 
that  the  phenyl  acids  which  are  abnormally  high  might  be  inhibiting  some 
step  or  steps  in  these  pathways.  Fellman  (1956)  therefore  studied  the  ef- 
fects of  such  substances  on  the  dopa  decarboxylase  from  beef  adrenal 
medulla.  The  order  of  inhibitory  potency  is  phenylpyruvate  >  phenyllac- 
tate  >  phenylacetate  >  phenylalanine.  Phenylpyruvate  inhibits  77%  when 
equimolar  (3.3  milf )  with  L-dopa.  The  low  plasma  epinephrine  levels  found 
in  these  patients  thus  might  be  due  to  such  an  inhibition,  but  another  point 
of  attack  would  have  to  be  adduced  for  a  suppression  of  melanin  formation. 
It  is  interesting  that  this  enzyme  is  inhibited  quite  strongly  by  norepi- 
nephrine and  dopamine,  whereas  epinephrine  exerts  no  effect  (Fellman, 
1959).  The  susceptibilities  of  dopa  decarboxylases  from  various  tissues 
are  obviously  different,  since  the  results  of  Fellman  are  often  different 
from  those  of  previous  workers. 

cf-Methyldopa  and  related  analogs  can  effectively  inhibit  decarboxylases 
in  vivo,  thereby  interfering  with  amine  formation  and  modifying  tissue 
function.  Direct  evidence  for  an  in  vivo  inhibition  was  obtained  by  intra- 
muscular injection  of  «-methyldopa  into  guinea  pigs  and  demonstration 
of  a  marked  depression  of  the  decarboxylation  of  both  dopa  and  5-hydroxy- 
tryptophan  in  isolated  kidney  15-30  min  afterward  (Westermann  et  at., 
1958).  Indeed,  the  inhibition  of  5-hydroxytryptophan  decarboxylation  is 
complete  and  after  90  min  is  83%.  More  indirect  evidence  has  been  obtain- 
ed by  showing  that  these  analogs  prevent  the  pharmacological  actions  of 
dopa  and  5-hydroxytryptophan,  these  actions  being  dependent  on  decar- 
boxylation of  these  substances  to  dopamine  and  serotonin,  respectively. 
Injection  of  dopa  leads  to  a  rise  in  the  blood  pressure  which  is  probably 
primarily  due  to  dopamine  (although  some  norepinephrine  and  epinephrine 
may  also  be  formed).  This  pressor  response  can  be  blocked  by  several  de- 
carboxylase inhibitors,  including  5-(3-hydroxycinnamoyl)salicylate  (Po- 
grund  and  Clark,  1956)  and  a-methyldopa  (Dengler  and  Keichel,  1958). 
There  is  no  effect  on  the  response  to  dopamine  or  norepinephrine.  The 
increase  in  cardiac  contractility  induced  by  dopa  is  also  completely  blocked 


TYROSINE   METABOLISM  315 

by  a-methyldopa.  5-Hydroxytryptoplian  causes  bronchoconstriction  in 
guinea  pigs  and  central  excitation  in  mice  (if  brain  monoamine  oxidase  is 
blocked),  these  effects  being  due  to  serotonin,  and  pretreatment  of  the 
animals  with  a-methyldopa  prevents  these  actions  (Westermann  et  al., 
1958). 

Decarboxylase  inhibition  should  lead  to  a  decrease  in  the  tissue  concen- 
trations of  certain  amines  and  this  has  been  demonstrated.  The  degree 
of  reduction  will  depend  on  the  relative  rates  of  formation  and  metabolism 
of  the  amines,  as  well  as  on  the  magnitude  of  the  decarboxylase  inhibition 
(which  will  depend  in  part  on  the  penetration  of  the  analogs  into  the  tis- 
sues), since  we  are  dealing  with  steady-state  multienzyme  systems.  Injec- 
tion of  a-methyldopa  (200  mg/kg)  into  dogs  leads  to  a  lowering  of  serotonin 
in  the  caudate  nucleus  (1.03  to  0.43  //g/g  at  3  hr),  a  more  prolonged  lowering 
of  norepinephrine  (2.41  to  1.77  /ngjg  at  24  hr)  (Goldberg  et  al.,  1960),  and 
a  fall  of  total  catecholamines  in  the  brain  stem  (0.17  ^-  0.08  //g/g),  heart 
(0.59  -^  0.26  //g/g),  and  spleen  (0.94  -^  0.43  //g/g)  (Stone  et  al,  1962).  In 
the  mouse,  brain  serotonin  is  reduced  but  norepinephrine  is  unaffected 
(S.  E.  Smith,  1960  a).  The  urinary  amines  in  four  hypertensive  patients 
were  decreased  by  a-methyldopa  (1-4  g  per  day):  the  reductions  were 
81%  for  t>Tamine,  63%  for  serotonin,  and  55%  for  tryptamine  (Gates 
et  al,  1960). 

Altering  these  amine  levels  should  produce  physiological  disturbances. 
It  has  been  found  that  a-methyldopa  lowers  the  blood  pressure  and  causes 
sedation  in  the  dog  (Goldberg  et  al,  1960),  decreases  coordinated  activity 
and  produces  miosis  in  mice  (S.  E.  Smith,  1960  a),  and  in  a  variety  of  animals 
induces  a  syndrome  similar  to  that  produced  by  reserpine  (including  hy- 
pothermia), a  drug  releasing  amines  from  the  tissues  (S.  E.  Smith,  1960  b). 
a-Methyldopa  is  being  studied  clinically  for  the  reduction  of  hypertension 
and  a  preliminary  report  (Gates  et  al.,  1960)  indicated  its  effectiveness, 
doses  of  1-4  g/day  for  1  week  leading  to  a  fall  in  supine  blood  pressure 
from  187.0/115.4  to  173.4/108.1  and  in  standing  blood  pressure  from  177.7/ 
119.3  to  138.6/98.0,  the  controls  being  given  a  placebo  in  a  double-blind 
study. 

It  is  thus  clear  that  a-methyl-dopa  can  inhibit  certain  amino  acid  de- 
carboxylases in  vivo,  can  alter  amine  levels  in  tissues,  and  can  produce 
physiological  disturbances  that  could  reasonably  be  attributed  to  the  in- 
hibition. Recently,  however,  more  detailed  studies  of  tissue  amines  have 
indicated  that  other  mechanisms  are  possibly  operative.  Injections  of 
a-methyl-3-hydroxyphenylalanine  into  guinea  pigs  lead  to  a  reduction 
in  brain  amines  (Fig.  2-4),  the  degree  of  lowering  and  the  duration  of  the 
effect  depending  on  the  amine.  a-Methyldopa  acts  similarly  but  is  slightly 
more  potent.  Simultaneously  there  is  an  inhibition  of  amino  acid  decar- 
boxylase (Fig.  2-5).  Cardiac  norepinephrine  is  even  more  potently  reduced 


316 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


100 


75 


50 


25- 


%  OF 
NORMAL 


SEROTONIN     / 

/ 

/       / 

/dopamine 

^^ 

/ 

y"^  norepinephrine 

\y 

1                           1 

20 

time  (HOURS)      — 


40 


60 


80 


100 


Fig.  2-4.  Effects    of    a-methyl-w-tyrosine    on    brain    amine    concentrations    in    the 
guinea  pig  following  an  intraperitoneal  dose  of  400  mg/kg.  (From  Hess  e<  a?.,  1961.) 


and  remains  at  a  lower  level  for  a  longer  time  than  in  brain.  The  results 
on  serotonin  and  dopamine  levels  in  the  brain  correspond  as  expected  to 
the  time  course  of  decarboxylase  inhibition,  but  the  prolonged  depletion 
of  norepinephrine  is  difficult  to  explain  on  this  basis.  Since  dopamine  levels 
return  to  normal  long  before  norepinephrine,  there  must  be  either  an  in- 
hibition of  the  /?-hydroxylation  of  dopamine  to  norepinephrine  or  an  in- 
terference with  the  tissue  binding  of  norepinephrine.  Hess  et  al.  (1961) 
showed  that  these  analogs  inhibit  /5-hydroxyIation  only  at  relatively  high 
concentrations,  which  might  have  been  produced  soon  after  injection  but 
certainly  would  not  occur  several  hours  later,  and  thus  inclined  to  the  sec- 
ond explanation.  The  lack  of  inhibition  of  dopamine  /5-oxidase  has  been 
confirmed  by  Creveling  et  al.  (1962) 

The  time  course  for  catecholamine  depletion  in  mouse  brain  and  heart 
following  administration  of  these  analogs  is  similar  to  that  following  re- 
serpine,  except  the  return  toward  normal  in  the  brain  is  somewhat  faster. 
Porter  et  al.  (1961)  examined  different  analogs  for  ability  to  reduce  nor- 
epinephrine in  the  brain  and  heart,  and  compared  these  results  with  their 
effectiveness  in  inhibiting  decarboxylation  of  5-hydroxytryptophan  in 
kidney  (Table  2-12).  Some  lack  of  correlation  between  the  two  activities  is 


TYROSINE    METABOLISM 


317 


Table  2-12 

Effective  Doses  of  Analogs  in  Inhibiting  Kidney  Decarboxylase 
AND  Lowering  Brain  and  Heart  Norepinephrine  in  Mice  " 


Analog 


ED^o  (mg/kg) 


Inhibition 

of  renal 

decarboxylase 


Depletion 
of  norepinephrine 


Brain 


Heart 


L-a-Methyl-3,4-dopa                                                      2.63  32               21 

L-a-Methyl-2,3-dopa                                                         1.35  >100  >100 

L-a-Methyl-3-hydroxy-PA                                             11.7  12                 1 

L-a-Methyldopamine  No  inhibition  >100                 5 

L-a-Methyl-3-hydroxyphenylethylamine  No  inhibition  33                  0.7 


"  Analogs  injected  intraperitoneally.  Kidney  decarboxylase  activity  determined 
45  min  after  injection.  Doses  required  to  half  deplete  tissues  of  norepineplirine  in 
last  two  columns.  Since  it  has  been  shown  that  only  the  L-isomers  are  active,  results 
are  given  on  this  basis  for  more  convenient  comparison.  (Data  from  Porter  et  al.,  1961.) 


100 


0  5 

TIME    (HOURS) 

Fig.  2-5.  Inhibition  of  amino  acid  decarboxylase  in  guinea  pig  tissues  by  a-methyl- 
TO-tyrosine  injected  intraperitoneally  at  400  mg/kg.  (From  Hess  et  al.,  1961.) 


318  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

evident  and  led  to  the  conclusion  that  an  action  other  than  decarboxylase 
inhibition  is  involved,  this  probably  being  an  interference  with  the  binding 
of  amines  in  the  tissues.  Since  the  a-methylamino  acid  analogs  can  be 
slowly  decarboxylated  to  the  corresponding  a-methylamines  in  the  body, 
and  since  these  amines  have  the  ability  to  deplete  norepinephrine,  it  was 
suggested  that  at  least  part  of  the  tissue  amine  lowering  is  due  to  displace- 
ment by  the  a-methyl  analogs  of  the  amines.  This  mechanism  has  been 
subscribed  to  by  several  recent  workers.  Maitra  and  Staehelin  (1963)  ad- 
ministered a-methyldopa  to  rats  and  guinea  pigs  and  found  the  cardiac 
catecholamine  levels  to  be  insignificantly  altered.  They  detected  an  increase 
in  the  a-methylnorepinephrine  level,  however,  and  a  corresponding  de- 
crease in  norepinephrine,  indicating  the  displacement  of  the  normal  catechol- 
amine with  its  analog.  MuschoU  and  Maitra  (1963)  further  demonstrated 
that  a-methylnorepinephrine  stored  in  the  sympathetic  nerve  endings  is 
released  by  nerve  stimulation  and  is  active  on  various  adrenergic  receptors. 
Pletscher  et  al.  (1964)  after  injecting  a-methyldopa  into  rats,  found  marked 
reduction  of  brain  serotonin  several  hours  later  and  felt  that  inhibition  of 
the  decarboxylase  could  not  explain  the  results.  They  inclined  to  the  view 
that  a-methyldopa  must  also  release  or  displace  stored  amines,  and  might 
also  interfere  with  the  uptake  of  amino  acids  by  the  brain.  However,  S.  E. 
Smith  (1963)  had  shown  that  a-methyldopa  is  only  a  very  weak  inhibitor 
of  5-hydroxy tryptophan  uptake  in  brain  slices  (50%  inhibition  at  6.7  vaM), 
although  it  inhibits  the  decarboxylase  potently  (50%  inhibition  at  0.00056 
mM).  It  may  be  noted  that  other  analogs  may  inhibit  uptake  more  than 
decarboxylation.  Day  and  Rand  (1964)  showed  that  a-methyldopa  can 
restore  the  activity  in  animals  whose  catecholamine  levels  have  been  de- 
pleted by  treatment  with  reserpine,  presumably  by  the  formation  of  a- 
methylnorepinephrine,  which  is  generally  only  1/9-1/2  as  pharmacologically 
potent  as  norepinephrine;  this  does  not  provide  direct  evidence  for  the 
mechanism  of  inhibition  by  a-methyldopa,  but  clearly  shows  that  it  forms 
an  active  amine  analog. 

The  principles  involved  in  the  interpretation  of  these  results  are  impor- 
tant in  the  general  field  of  analog  inhibition  and  the  disturbances  produced 
in  tissue  function,  and,  furthermore,  the  foregoing  experiments  might  be 
carelessly  construed  as  invalidating  the  decarboxylase  inhibition  mechanism; 
thus  some  critical  comments  may  not  be  out  of  place. 

(1)  It  is  unfortunate  that  Porter  etal.  (1961)  did  not  determine  decarbox- 
ylase inhibition  in  brain  and  heart  for  comparison  with  amine  depletion 
in  these  tissues,  since  the  inhibition  in  kidney  may  be  quite  different.  In 
the  first  place,  the  penetration  of  the  analogs  into  the  three  tissues  may 
vary.  Indeed,  Hess  et  al.  (1961)  found  that  a-methyldopa  concentrations 
in  brain,  heart,  and  kidney  are  in  the  ratio  1:1.66:3.28  at  1  hr  and  1:1.8:9.8 
at  5  hr  after  injection.  The  concentrations  of  the  other  analogs  in  the  tis- 


TYROSINE    METABOLISM  319 

sues  are  not  known,  but  it  was  demonstrated  that  a-methyldopa  does  not 
penetrate  into  brain.  In  the  second  place,  the  decarboxylases  from  the  var- 
ious tissues  may  have  different  susceptibilities  to  the  analogs,  as  we  have 
noted  above.  Perhaps  cardiac  decarboxylase  is  resistant  to  a-methyldopa 
while  the  renal  enzyme  is  more  sensitive. 

(2)  Monoamine  oxidase  inhibitors  were  administered  during  norepineph- 
rine depletion,  and  norepinephrine  levels  immediately  rose  (Hess  et  al., 
1961).  It  was  concluded  that  "biosynthesis  of  norepinephrine  can  still 
occur  in  animals  treated  with  or-methylamino  acids."  All  these  results 
mean  is  that  inhibition  of  the  decarboxylase  was  not  complete.  In  any 
sequence 

AillB^C 

an  inhibition  of  (1)  will  lower  B  and  an  inhibition  of  (2)  will  raise  B  (see 
Chapter  1-7). 

(3)  It  is  stated  that  the  decarboxylase  step  is  the  most  rapid  in  the 
over-all  sequence  and  therefore  cannot  be  rate-limiting  (Hess  et  al.,  1961). 
The  conclusion  was  that  the  decarboxylase  must  be  inhibited  very  strongly 
for  any  effect  to  be  observed.  First,  it  is  very  difficult  to  establish  that 

very  very 

Tyrosine  — >  dopa  — >  dopamine  — >  norepinephrine 

sloiv  fast  slow 

these  are  the  relative  rates  of  the  reactions  in  vivo,  where  the  concentra- 
tions and  states  of  the  enzymes  are  quite  different  than  when  isolated  from 
the  cells.  Second,  the  rate  of  formation  of  norepinephrine  in  a  steady  state 
is  controlled  by  the  first  reaction  (or  a  previous  reaction)  since  these 
reactions  are  virtually  irreversible.  An  inhibition  of  dopa  decarboxylase 
will  not  alter  the  rate  of  norepinephrine  formation  as  long  as  a  steady 
state  is  maintained;  the  concentration  of  dopamine  will  also  be  unchanged 
in  the  steady  state.  However,  it  has  been  demonstrated  that  dopamine 
concentration  falls,  indicating  that  a  departure  from  a  steady  state  has 
occurred.  One  must  also  consider  the  other  possible  metabolic  pathways 
for  dopamine  (e.g.,  oxidation  and  0-methylation),  since  this  is  a  divergent 
sequence.  A  certain  depression  of  the  decarboxylation  need  not  be  reflected 
in  the  same  depression  of  norepinephrine  formation,  even  under  nonsteady- 
state  conditions;  the  latter  can  be  either  greater  or  less  than  the  inhibition 
of  dopamine  formation.  In  the  case  of  serotonin  formation,  the  decar- 
boxylation is  the  last  step,  and  whether  it  will  be  inhibited  or  not  will 
depend  on  the  degree  to  which  5-hydroxytryptophan  concentration  can 
rise  to  overcome  the  block.  In  any  event,  it  has  been  shown  that  the 
in  vivo  inhibition  of  decarboxylase  by  these  analogs  can  be  very  high 
and  sometimes  complete. 


320  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

(4)  The  fact  that  certain  a-methyl  analogs  of  the  catecholamines  can 
deplete  tissues  of  the  amines,  although  they  do  not  inhibit  the  decarboxy- 
lase, is  not  evidence  against  a  decarboxylase  inhibition  mechanism  for  the 
a-methylamino  acids,  but  indicates  another  mechanism,  which  may  play 
a  role  in  the  prolonged  lowering  of  norepinephrine  levels  without  neces- 
sarily being  involved  in  the  initial  rapid  fall  in  tissue  amines.  The  relatively 
rapid  return  of  serotonin  and  dopamine  levels  to  normal  (Fig.  2-4)  suggests 
that  there  is  no  generalized  disturbance  in  tissue  amine  binding,  but  that 
the  effect  is  specifically  on  norepinephrine.  The  most  satisfactory  position 
at  the  present  time  might  be  the  following:  the  initial  marked  fall  in  tissue 
amines  brought  about  by  the  a-methyl  analogs  is  primarily  due  to  an  in- 
hibition of  decarboxylation  (perhaps  supplemented  at  peak  concentrations 
by  inhibition  of  other  steps,  such  as  /?-hydroxylation),  and  further  distur- 
bances in  amine  binding  are  progressively  produced  by  the  a-methylamines 
formed  from  the  inhibitors,  so  that  even  when  the  decarboxylase  is  normally 
active  again  the  tissues  cannot  concentrate  certain  of  the  normal  catechol- 
amines. 

Dopamine  p-Hydroxylase 

This  enzyme  catalyzes  the  synthesis  of  norepinephrine  from  dopamine 
and,  as  we  have  seen,  its  inhibition  by  analogs  could  be  both  theoretically 
and  practically  important.  Hess  et  al.  (1961)  found  that  a-methyl-3-hy- 
droxyphenylalanine  does  not  inhibit  at  2  tclM.  but  inhibits  50%  at  4  toM. 
The  concentration  of  a-methyldopa  1  hr  after  injection  is  given  as  376 
//g/g  in  the  heart  and  this  could  mean  a  concentration  around  750  //g/ml 
of  intracellular  fluid  (assuming  extracellular  fluid  has  a  low  concentration 
at  this  time).  This  is  approximately  equivalent  to  4  mM  so  that  appreciable 
inhibition  might  occur.  Inhibition  data  indicate  that  3-methyl-3-hydroxy- 
phenylalanine  concentrations  in  the  tissues  are  roughly  the  same  as  a- 
methyldopa  concentrations.  Until  more  is  known  about  the  nature  of  the 
inhibition  of  this  enzyme,  it  might  be  safe  to  conclude  that  it  plays  some 
role  in  the  effects  of  the  or-methyl  analogs  of  phenylalanine. 

Various  amines  can  inhibit  this  enzyme  (Goldstein  and  Contrera,  1961). 
When  dopamine  concentration  is  0.26  niM,  the  following  inhibitions  are 
observed:  tyramine  at  2.9  raM  (75%),  /?-phenylethylamine  at  3.3  vaM 
(45%),  amphetamine  at  5.9  milf  (35%),  and  3-methoxydopamine  at  4.8 
mM  (15%).  None  of  these  inhibitors  appears  to  be  potent  enough  to  be 
practically  important  in  reducing  norepinephrine  synthesis  and,  further- 
more, these  amines  are  so  pharmacologically  active  that  their  use  is  limited. 
Benzyloxy amine,  and  particularly  the  p-hydroxyl  derivative,  inhibit  this 
enzyme  rather  potently,  0.01  mM  of  the  latter  blocking  almost  completely 
after  90  min  (van  der  Schoot  et  al.,  1963),  this  being  attributed  to  the  iso- 
steric  relation  between  phenethylamines  and  benzyloxyamines. 


i 


TRYPTOPHAN  METABOLISM  321 

TRYPTOPHAN    METABOLISM 

Tryptophan  is  involved  in  several  important  metabolic  pathways,  form- 
ing active  substances  as  well  as  being  incorporated  into  proteins,  so  that 
many  attempts  to  block  these  pathways  specifically  with  analogs  have 
been  made.  Growth  inhibition  and  physiological  disturbances  are  readily 
produced  by  many  of  these  analogs.  (See  scheme  on  page  322). 

Synthesis   of  Tryptophan 

L-Tryptophan  is  a  potent  feedback  inhibitor  of  the  conversion  of  5- 
phosphoshikimate  to  anthranilate,  an  early  reaction  in  tryptophan  biosyn- 
thesis, and  5-methyltryptophan  also  inhibits,  although  not  so  strongly, 
a  phenomenon  (i.e.,  inhibition  of  a  biosynthetic  step  by  an  analog)  termed 
false  feedback  inhibition  by  Moyed  (1960).  It  is  likely  that  this  mechanism 
explains  the  bacteriostatic  activity  of  this  analog.  The  condensation  of 
anthranilate  and  5-phosphoribosyl-l-pyrophosphate  is  not  inhibited  by 
5-methyltryptophan,  but  6-fiuorotryptophan  is  inhibitory.  A  later  reac- 


COO 


NH, 


Tryptophan  Anthranilate 

tion  in  this  sequence,  the  conversion  of  anthranilic  deoxyribonucleotide  to 
indoleglycerol-3-phosphate,  is  inhibited  by  a  variety  of  anthranilate  de- 
rivatives, especially  the  3-  and  4-methyl  analogs  (Gibson  and  Yanofsky, 
1960).  The  final  reaction,  the  condensation  of  indole  and  serine  to  form 
tryptophan,  catalyzed  by  tryptophan  synthetase,  is  a  major  site  of  the 
attack  by  4-methyltryptophan,  which  is  a  bacterial  growth  inhibitor 
(Trudinger  and  Cohen,  1956).  The  5-  and  6-methyl  indoles  are  fairly  potent 
competitive  inhibitors,  with  K^  values  near  0.1  roM  (Hall  et  al.,  1962). 
They  are  also  antibacterial.  The  growth  depression  of  E.  coli  is  counteracted 
by  tryptophan  (Fig.  2-6).  At  least  two  sites  for  the  inhibition  have  been 
demonstrated.  Tryptophan  synthetase  is  inhibited  competitively,  but  there 
is  also  a  block  of  the  much  earlier  formation  of  anthranilate.  There  are  no 
effects  on  the  immediate  metabolism  of  anthranilate  or  on  tryptophanase, 
which  indeed  readily  splits  the  analog  to  4-methylindole.  The  bacteriostatic 
action  is  probably  due  mainly  to  suppression  of  tryptophan  synthesis  rather 
than  to  a  disturbance  of  tryptophan  utilization.  Thus  several  steps  in  the 
biosynthesis  are  susceptible  to  analogs  and  it  is  quite  possible  that  other 


322 


2.  ANALOGS  OF  ENZYME  EEACTION  COMPONENTS 


™      rt 


O)       T3 


+  a 


x: 
a 
o 


S    S    S 


^     eg     CO 


o 

x: 
a, 


TRYPTOPHAN  METABOLISM 


323 


unstudied  reactions  are  likewise  inhibited.  A  more  indirect  mechanism  is 
the  inhibition  of  the  synthesis  of  tryptophan  synthetase  in  growing  cells, 
whereby  tryptophan  formation  is  further  reduced. 


100 


Fio.  2-6.  The  inhibition  of  E.  coli  growth  by  4-methyltryptophan 
at  0.01    mM   and   its   antagonism   by   increasing   tryptophan   con- 
centrations.  (From   Trudinger   and   Cohen,   1956.) 

Tryptophanase 

This  bacterial  and  fungal  enzyme  may  be  involved  in  the  fermentation 
of  tryptophan  and  is  responsible  for  the  putrefactive  reaction  in  the  in- 
testines. It  is  potently  and  competitively  inhibited  by  the  product  indole 
and  less  potently  by  some  analogs  (as  shown  in  the  accompanying  tabula- 


Inhibitor 


Relative  —  AF  of  binding 
(kcal/mole) 


Indole 

/?-3-Indolylpropionate 

3-Indolylacetate 

/?-3-Indolylethylamine 

DL-/3-l-Methyl-3-indolylalanine 


5.40 
4.05 
3.76 
2.55 
1.44 


324  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

tion)  (Gooder  and  Happold,  1954).  The  importance  of  the  indole  N  in 
binding  is  indicated  by  the  weak  inhibition  with  the  last  analog  (tryptophan 
with  the  indole  N  methylated)  and  the  failure  of  indene  (the  hydrocarbon 


/ 
.CHoCOO"  /\  /CH,— CH 


NH^ 


IJ  COO' 

N'  "^        N 

I  I 

H  CH3 

3-Indolylacetate  0-1 -Methyl- 3 -indolylalanine 


.CHoCHoNH, 


N 
I 
H 

0-3-Indolylethylamine 

analog  of  indole)  to  inhibit,  while  the  importance  of  the  carboxylate  group 
is  reflected  in  the  weak  inhibition  by  the  indolylethylamine.  The  potency 
of  indole  may  be  attributed  to  the  fact  that  it  may  not  have  to  be  oriented 
in  a  manner  necessary  for  reaction  of  the  side  chain. 

Tryptophan    Pyrrolase   (Tryptophan    Peroxidase) 

This  enzyme  initiates  one  of  the  most  important  catabolic  pathways  of 
tryptophan  and  is  readily  inhibited  by  certain  analogs.  Hayaishi  (1955  b) 
found  the  Pseudomonas  enzyme  to  be  sensitive  to  the  hydroxytryptophans, 
and  calculated  the  values  of  K;  shown  in  the  following  tabulation.  Since 


Substance 


Ki  or  Kg  Relative  —  Zli^  of  binding 

(milf)  (kcal/mole) 


5-  Hydroxy  tryptophan  0 .  002  8 .  06 

7-Hydroxytryptophan  0.12  5.55 

L-Tryptophan  0.4  4.81 


5-hydroxytryptophan  is  normally  formed  from  tryptophan  on  the  pathway 
to  serotonin,  its  potent  inhibition  of  the  pyrrolase  suggests  that  it  may 
play  a  role  in  regulating  tryptophan  metabolism.  The  enzyme  from  rat 
liver  is  also  inhibited  by  5-hydroxytryptophan  and  even  more  potently 


TRYPTOPHAN  METABOLISM  325 

by  serotonin  (Frieden  et  al.,  1961).  The  K/s  in  the  following  tabulation 
indicate  3-indolylacrylate  to  be  the  most  effective  inhibitor.  Other  inhibi- 
tions observed  when  (S)  =  (I)  =  3  mM  are:  indole  69%,  tryptazan  50%, 


Inhibitor 

K, 

Relative  —  Ji"  of  binding 

(mM) 

(kcal/mole) 

3-Indolylacrylate 

0.012 

6.99 

Serotonin 

0.067 

5.93 

5-HydroxytrjT)tophan 

0.094 

5.71 

3-Indolylbutyrate 

0.16 

5.39 

Tryptamine 

0.20 

5.25 

3-Indolylpropionate 

0.29 

5.02 

3-Indolylacetate 

0.91 

4.32 

^-Methyltryptophan 

1.1 

4.20 

D-Tryptophan 

1.6 

3.97 

6-Fluorotr>i)tophan 

2.0 

3.83 

5-Fluorotryptophan 

2.2 

3.77 

5-methyltryptophan  38%,  and  6-methyltryptophan  33%.  The  analogs  with 
altered  side  chains  are  competitive  while  the  others  are  mainly  noncompeti- 
tive. The  roughly  equivalent  binding  of  tryptamine  and  3-indolylpropionate, 
and  of  serotonin  and  5-hydroxytryptophan,  might  indicate  that  the  binding 
is  primarily  with  the  indole  ring,  the  side  chains  contributing  little,  and  this 
is  substantiated  by  the  fact  that  indole  is  bound  approximately  as  well 
as  3-indolylpropionate.  The  stronger  binding  of  3-indolylacrylate  compared 
to  3-indolylpropionate  (about  2  kcal/mole)  is  thus  difficult  to  account  for 
unless  there  is  a  modification  of  the  interaction  of  the  indole  N.  Tryptophan 
pyrrolase  is  an  inducible  enzyme  in  the  rat  but  none  of  these  analogs  is 
active,  although  Sourkes  and  Townsend  (1955)  found  a-methyltryptophan 
to  induce  after  subcutaneous  injection. 

Tryptophan    Hydroxylase   (Phenylalanine   Hydroxylase) 

Excessive  feeding  of  phenylalanine  leads  to  low  blood  serotonin,  a  low 
excretion  of  5-hydroxyindoleacetate,  and  a  decrease  in  brain  serotonin, 
and  this  has  usually  been  attributed  to  an  inhibition  of  5-hydroxytrypto- 
phan decarboxylase.  However,  no  accumulation  of  5-hydroxytryptophan 
has  been  demonstrated  and  it  is  possible  that  the  site  of  the  inhibition  might 
be  earlier,  perhaps  on  the  hydroxy lation  of  tryptophan  (Freedland  et  al., 
1961).  Hydroxylating  preparations  from  rat  liver  are  indeed  quite  potently 
inhibited  by  L-phenylalanine,  and  also  by  phenylpyruvate  and  phenyUac- 
tate  (K,„  for  L-tryptophan  is  29  mM,  and  K^  for  L-phenylalanine  is  0.22  mM). 


326  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

It  is  likely  that  the  same  enzyme  is  responsible  for  the  hydroxylation  of 
both  phenylalanine  and  tryptophan,  since  the  K^  for  phenylalanine  is  close 
to  the  iii,,,  when  it  is  the  substrate;  the  affinity  for  tryptophan  is,  however, 
much  less.  These  results  may  help  to  explain  some  of  the  changes  observed 
in  phenylpyruvic  oligophrenia  (see  pages  329  and  429). 

Tryptophan-Activating  Enzyme 

An  activating  enzyme  from  pancreas  is  specific  for  tryptophan  with 
respect  to  other  normal  amino  acids  but  can  activate  certain  analogs  of 
tryptophan  (Sharon  and  Lipmann,  1957).  The  analogs  tested  fall  into  three 
categories: 

Group  I  are  activated  (tryptazan,  azatryptophan,  5-fluorotryptophan, 
and  6-fluorotryptophan). 

Group  11  are  inhibitory  (tryptamine,  D-tryptophan,  /5-methyltryptophan, 
5-hydroxytryptophan,  5-methyltryptophan,  and  6-methyltryptophan). 

Group  III  are  inactive  (indole,  indoleacetate,  6-methyltryptazan,  and 
iV-acety  Itry  ptophan ) . 


CH,— CH  ^-^^  CH^— CH 

^coo'  f       H       H  coo' 

N  N  N 

I  I 

H  H 


r 


Tryptazan  Azatryptophan 

Reciprocal  plots  were  said  to  indicate  competitive  inhibition  but  actually 
do  not  show  pure  competition,  and  it  might  better  be  designated  as  partially 
competitive  inhibition.  For  analogs  to  be  activated  they  must  match  the 
size  of  tryptophan,  and  the  introduction  of  bulkier  groups  prevents  reaction 
with  ATP  but  allows  binding  and  inhibition.  Azatryptophan  and  tryptazan 
are  both  incorporated  into  proteins.  In  E.  coli,  azatryptophan  permits  syn- 
thesis of  proteins  and  nucleic  acids  but  many  of  the  enzymes  are  not  in 
the  normal  forms  (Pardee  and  Prestidge,  1958).  The  synthesis  of  adaptive 
enzymes,  such  as  /?-galactosidase,  is  inhibited  very  rapidly,  and  phage  for- 
mation is  blocked  more  readily  than  bacterial  growth.  The  induced  synthe- 
sis of  maltase  in  yeast  is  also  very  potently  suppressed  by  tryptazan,  in- 
corporation of  all  amino  acids  being  simultaneously  blocked  (Halvorson 
etal.,  1955).  5-Methyltryptophan  and  6-methyltryptazan  inhibit  moderate- 
ly,  while   6-methyltryptophan  is  inactive. 


GLUTAMATE  METABOLISM 


327 


GLUTAMATE   METABOLISM 

Glutamate  occupies  a  central  position  in  many  important  metabolic 
pathways  and  serves  to  link  amino  acid  metabolism  with  the  tricarboxylic 
acid  cycle.  Its  relationship  to  the  biochemically  active  glutamine  and  the 
physiologically  active  /-aminobutyrate  (GABA)  makes  possible  specific 
inhibitions  of  glutamate  reactions  of  great  interest.  The  reactions  catalyzed 
by  enzymes  studied  with  respect  to  analog  inhibition  are  shown  in  the 
following  scheme. 


glutamylhydroxamate 


D  -glutamate 


(4) 
o-ketoglutarate 


(7) 
(3)  (8) 


glutamine 


(6) 


(1,2) 


L-glutamate      — 

(5) 

r 
>'  -aminobuty  rate 


(10) 


proline 


A'-pyrroline-5-carboxylate 

a 

(10) 
glutamic -7  -semialdehyde 


(1)  L-glutamate  dehydrogenase 

(2)  glutamate  transaminases 

(3)  glutamate  racemase 

(4)  D-glutamate  oxidase 

(5)  glutamate  decarboxylase 


(6)  glutamine  synthetase 

(7)  glutaminase 

(8)  formylglycinamidine  phosphoriboside  synthetase 

(9)  7  -glutamyl  transferase 

(10)  A'-pyrroline-5-carboxylate  dehydrogenase 


Glutamate  Decarboxylase 

Several  analogs  of  glutamate  inhibit  its  utilization  by  Lactobacillus  ara- 
binosus,  and  thus  a  study  of  decarboxylase  from  E.  coli  was  undertaken 
by  Roberts  (1953).  The  two  most  potent  inhibitors  are  a-oximinoglutarate 
and  a-methylglutamate,  but  the  former  is  probably  active  by  virtue  of 
its  hydrolysis  to  hydroxylamine  which  inactivates  pyridoxal  phosphate. 
The  latter  analog  inhibits  competitively  when  added  with  the  substrate, 
but  if  it  is  preincubated  with  the  enzyme  the  inhibition  becomes  progres- 
sively more  noncompetitive  and  difficulty  reversible.  The  rates  of  combina- 
tion with  the  enzyme  and  dissociation  from  the  enzyme  are  very  slow. 
The  methyl  group  interferes  with  the  normal  binding  of  the  molecule  so 
that  decarboxylation  does  not  occur,  but  by  some  unknown  mechanism 
brings  about  a  type  of  binding  that  is  very  strong,  a  behavior  seen  with 
some  other  a-methylamino  acids.  The  decarboxylation  of  glutamate  in 
rat  brain  homogenates  is  also  inhibited  competitively  by  aspartate  {K„^ 


328  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

=  21  mM,  and  K^  =  23  milf )  (Wingo  and  Awapara,  1950),  which  is  rather 
surprising  because  of  the  shorter  intercarboxylate  distance. 

Ghitamate  decarboxylase  from  the  squash  Curcurbita  moscliata  is  inhib- 
ited competitively  by  a  variety  of  organic  acids  (see  accompanying  tabu- 
lation), and  the  results  are  of  some  interest  with  regard  to  structure  and 

Inhibitor  (13.6  mM)  %  Inhibition 


Monocarboxylates 

Formate  17 

Acetate  42 

Propionate  10 

7i-Butyrate  18 

Isobutyrate  8 

n-Valerate  25 

Iso  valerate  15 

n-Caproate  24 

Isocaproate  20 

Dicarboxylates  (saturated) 

Oxalate  0 

Malonate  0 

Succinate  0 

Glutarate  24 

Adipate  49 

Pimelate  58 

Suberate  37 

Dicarboxylates  (unsaturated) 

Fumarate  0 

Maleate  14 

Citraconate  0 

Mesaconate  7 

Itaconate  11 

Tricarboxylates 

cis-Aconitate  29 

irans-Aconitate  18 


intercarboxylate  distance  (Ohno  and  Okunuki,  1962).  Glutamate  concen- 
tration was  27.2  mM  in  all  cases.  Many  amino  acids  examined  are  weakly 
inhibitory  or  without  effect.  The  maximal  inhibition  by  pimelate  in  its 
series  probably  indicates  that  a  cambering  of  the  molecule  is  necessary  for 
binding  of  the  two  carboxylate  groups,  or  possibly  that  the  cationic  groups 
of  the  enzyme  are  farther  apart  than  in  glutamate.  The  relatively  high 


Relative 

—  Zli^of  bindin 

g  (kcal/mole)" 

Hanson  (1958) 

Tashian   (1961) 

3.15 

4.16 

— 

4.26 

2.62 

2.62 

2.53 

3.73 

1.75 

3.83 

0.60 

— 

0.53 

— 

GLUTAMATE  METABOLISM  329 

inhibition  by  acetate  is  surprising;  if  the  monocarboxylates  interact  with 
the  cationic  groups,  one  might  expect  propionate  or  butyrate  to  be  more 
inhibitory. 

The  problem  of  the  abnormal  brain  development  in  phenylpyruvic  oligo- 
phrenia prompted  an  investigation  of  the  effects  of  the  phenyl  acids  on 
brain  glutamate  decarboxylase  by  Hanson  (1958);  the  results  are  presented 
in  the  tabulation  below  in  which  they  are  compared  with  those  of  Tashian 


Inhibitor 


p-Hydroxyphenylacetate 

o-Hydroxyphenylacetate 

Phenylpyruvate 

Phenylacetate 

p-Hydroxyphenylpyruvate 

Phenylalanine 

Phenyllactate 

°  Relative  —  AF's  of  binding  adjusted  so  that  value  for  phenylpyruvate  is  the  same 
in  each  series. 

(1961),  who  also  used  rat  brain.  There  are  some  rather  striking  differences, 
part  of  which  may  be  due  to  the  procedures  used  since  these  analogs, 
although  stated  to  be  competitive,  present  deviations  from  classic  kinetic 
formulations  (and  for  this  reason  the  binding  energies  calculated  from  ap- 
parent K/s  are  probably  not  very  reliable).  If  these  analogs,  which  are 
present  in  high  concentrations  in  the  blood  enter  the  brain  readily,  it  is 
possible  that  they  depress  the  formation  of  y-aminobutyrate  (GABA) 
which  may  be  essential  for  normal  neurological  development.  In  branched- 
chain  ketonuria  (maple  sugar  urine  disease)  various  keto  and  hydroxy 
fatty  acids  accumulate  in  the  body,  which  Tashian  (1961)  showed  also 
inhibit  glutamate  decarboxylase  (relative  —  JF's  of  binding  for  a-hydroxy- 
isovalerate,  a-ketoisovalerate,  and  the  corresponding  isocaproates  from 
3.58  to  4.50  kcal/mole  on  the  scale  above).  The  enzyme  from  E.  coli,  on 
the  other  hand,  is  relatively  insensitive  to  any  of  these  analogs. 

L-Glutamate  Dehydrogenase 

The  oxidative  deamination  of  L-glutamate  in  beef  liver  homogenates  is 
catalyzed  by  a  NAD-linked  dehydrogenase,  the  inhibition  of  which  at 
pH  8.4  was  thoroughly  studied  by  Caughey  et  al.  (1957).  The  accompanying 
tabulation  shows  the  K^  values  for  competitive  analogs,  from  which  the 


330  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


Inhibitor 

(mM) 

Relative  —  Zli*'  of  binding 
(kcal/mole) 

5-Bromofuroate 

0.059 

6.00 

5-Chlorofuroate 

0.063 

5.96 

5-Nitrofuroate 

0.17 

5.35 

m-Iodobenzoate 

0.46 

4.74 

»n-Bromobenzoate 

0.54 

4.65 

Isophthalate 

0.66 

4.62 

Glutarate 

0.58 

4.60 

a-Ketoglutarate 

0.73 

4.45 

w -Chlorobenzoate 

1.02 

4.25 

D-Glutamate 

2.0 

3.83 

m-Nitrobenzoate 

3.4 

3.51 

Trimesate 

4.0 

3.40 

Fumarate 

6.8 

3.08 

Succinate 

11 

2.78 

Adipate 

16 

2.55 

relative  binding  energies  have  been  calculated.  Inhibitors  were  classed  as 
competitive  if  the  interaction  constant  a  is  greater  than  10;  trimesate  and 
D-glutamate  are  only  partially  competitive,  with  a  =  1.7.  It  was  postulated 

CCXD' 
^O.     ^COO" 


COO 

Furcate  Isophthalate 


COO 


COO' 


OOC       \^       COO 


Trimesate  Indole-3-carboxylate 

that  the  enzyme  contains  two  cationic  groups  at  a  separation  optimal  for 
interaction  with  the  carboxylate  groups  of  glutamate,  glutarate,  and  iso- 
phthalate. The  effects  of  some  other  inhibitors  for  which  the  ^/s  were 
not  calculated  are  shown  in  the  following  tabulation;  most  of  these  are 


GLUT  AM  ATE  METABOLISM  331 

relatively  weak  (fumarate  is  included  for  comparison  with  values  in  the 
preceding  table).  L-Glutamate  was  2  mM  in  each  case.  The  following  are 
not  inhibitory  at  10  mJ-l:  L-glutamine,  L-diethylglutamate,  /?-methylglu- 
tarate,  citrate,  o-  and  p-hydroxybenzoate. 


Inhibitor 

Concentration 
(mM) 

%  Inhibition 

Benzoate 

10 

12 

Furoate 

10 

26 

Phthalate 

10 

17 

Terephthalate 

6.7 

16 

Isophthalate 

2 

50 

<ran.5-Aconitate 

8 

20 

Indole-2-carboxylate 

7.5 

21 

Indole-3-carboxylate 

8.3 

38 

m-Hydroxybenzoate 

10 

27 

Pyridine-2,6-dicarboxylate 

10 

27 

Fumarate 

10 

29 

In  the  various  substituted  benzenes  the  7neta  compounds  are  invariably 
the  most  potent  inhibitors,  presumably  because  the  intergroup  distances 
are  close  to  the  enzyme  intercationic  separation.  The  dipoles  of  the  halogen 
and  nitro  compounds  may  interact  with  the  cationic  group  since  they  may 
be  represented  as: 

However,  there  is  no  correlation  of  inhibitory  activity  with  dipole  moment. 
One  might  think  that  the  dipole-cation  interaction  would  be  weaker  than 
the  carboxylate-cation  interaction,  but  the  hydration  of  the  inhibitors  must 
also  be  considered.  Less  water  needs  to  be  displaced  when  the  dipoles  ap- 
proach the  enzyme  cationic  group.  It  was  pointed  out  that  all  good  in- 
hibitors are  reasonably  planar  and  the  presence  of  bulky  groups  protruding 
lowers  the  affinity.  The  particularly  good  binding  of  the  furoates  may  be 
related  to  some  interaction  of  the  ring  0  with  the  enzyme.  The  reverse 
reaction  from  or-ketoglutarate  to  glutamate  is  inhibited  less  than  the  for- 
ward reaction  by  glutarate,  isophthalate,  and  5-bromofuroate,  and  the 
inhibitions  are  noncompetitive. 

The  L-glutamate  dehydrogenase  from  cockroach  muscle  mitochondria 


332  2.  ANALOGS  OF  p:;nzyme  reaction  components 

is  similarly  inhibited  (see' accompanying  tabulation),  and  here  glutarate 
also  appears  to  present  a  relatively  good  fit  to  the  active  site  (Mills  and 
Cochran,  1963).  Glutamate  was  15  mM  in  all  cases.  The  reverse  reaction 
catalyzed   by   the   glutamate  dehydrogenases   (both   NAD-   and   NADP- 


Inhibitor  (3  mM) 

%  Inhibition 

Succinate 

15 

Fumarate 

15 

Malate 

15 

Glutarate 

65 

Adipate 

30 

D-Glutamate 

55 

L- Aspartate 

20 

D-Aspartate 

10 

linked)  from  Fusarium  oxysporum  is  also  inhibited  by  glutarate,  the  K,^ 
for  a-ketoglutarate  being  2.1  mM,  and  the  K^  for  glutarate  1.52  mM 
(Sanwal,  1961). 

Glutaminase 

The  deamidation  of  glutamine  is  inhibited  by  the  product  glutamate 
and  this  is  not  a  reversal  of  the  equilibrium  but  a  competition  for  the 
active  center,  as  first  pointed  out  by  Krebs  (1935).  L-Glutamate  and  d- 
glutamate  inhibit  guinea  pig  kidney  glutaminase  equally  (98%  at  25  mM 
when  glutamine  is  8.7  mM).  Inhibition  by  glutamate  has  been  confirmed 
for  the  enzyme  from  guinea  pig  kidney  (van  Baerle  et  al.,  1957),  pig  kidney 
(Klingman  and  Handler,  1958),  dog  kidney  (Sayre  and  Roberts,  1958), 
and  rat  brain  (Blumson,  1957).  The  inhibition  has  generally  been  found 
to  be  noncompetitive  with  respect  to  glutamine,  but  oddly  is  competitive 
with  phosphate  on  the  phosphate-activated  glutaminase  from  dog  kidney. 
The  ammonium  ion  is,  however,  strictly  competitive  with  glutamine  on 
both  the  pig  and  dog  kidney  enzymes.  Another  type  of  glutaminase  (called 
glutaminase  II),  which  is  transaminating  in  the  presence  of  pyruvate  and 
is  obtained  from  guinea  pig  kidney,  is  not  inhibited  by  even  100  mM 
glutamate  (Goldstein  et  al.,  1957).  Sayre  and  Roberts  (1958)  pictured  the 
active  center  as  containing  two  cationic  groups,  one  binding  the  phosphate 
and  one  the  glutamine  carboxylate  group;  the  negatively  charged  phosphate 
also  interacts  with  the  positive  or-amino  group  of  glutamine.  Since  the  active 
enzyme  is  the  phosphate  complex,  it  is  easy  to  see  why  phosphate  would 
antagonize  inhibitions  produced  by  certain  substances  (e.g.,  dyes  such  as 
bromosulfalein  or  bromcresol  green  which  complex  with  both  enzyme 
cationic  sites),  but  it  is  difficult  to  understand  why  glutamate  inhibits 


GLUTAMATE  METABOLISM  333 

competitively  with  respect  to  phosphate.  It  was  stated  that  for  a  substance 
to  compete  with  glutamine  it  should  have  affinity  for  the  enzyme-phosphate 
complex,  and  it  would  seem  that  glutamate  may  fall  into  this  category. 

Few  other  analogs  have  been  tested  on  this  enzyme.  Krebs  (1935)  ob- 
served a  mild  inhibition  by  DL-/?-hydroxyglutamate  (22%  at  80  mM  when 
glutamine  40  mM),  and  Girerd  et  al.  (1958)  reported  inhibition  by  ethyl 
D-glutamate  and  DL-/5-methylglutamate.  Because  of  the  postulated  role 
of  glutaminase  in  renal  function,  these  two  latter  analogs,  along  with 
L-glutamate  and  bromosulfalein,  were  tested  in  vivo.  These  inhibitors  re- 
duce diuresis  in  rats  around  50%  at  50  mg/kg  subcutaneously,  whereas  a 
group  of  five  less  potent  glutaminase  inhibitors  actually  increase  diuresis. 

Formylglycinamidine  Phosphoriboside  Synthetase 

Glutamine  participates  in  purine  biosynthesis  by  contributing  its  amide 
N.  Azaserine  and  6-diazo-5-oxo-L-norleucine  (DON)  are  potent  inhibitors 
of  inosinate  biosynthesis  and  lead  to  the  accumulation  of  formylglycina- 
mide  phosphoribotide  (FGAR).  These  substances  may  be  considered  as 
analogs  of  glutamine  and  have  been  shown  to  inhibit  formylglycinamide 
ribonucleotide  amidotransferase  competitively  with  respect  to  glutamine 


N:=N=CH- 

-CO— O-CH2- 
Azaserine 

-CH 

COO' 

N=N=CH- 

NH3 
/ 
-CO— CH2CH2— CH 

^COO 
DON 

(Levenberg  et  al.,  1957).  The  K„,  for  glutamine  is  0.615  mM  and  the  K^s 
for  azaserine  and  DON  are  0.034  mM  and  0.0011  mM,  respectively.  Once 
azaserine  binds  to  the  enzyme,  however,  an  irreversible  reaction  occurs, 
due  perhaps  to  an  alkylation  of  the  enzyme.  French  et  al.  (1963  a)  pointed 
out  that  50%  inhibition  can  be  obtained  with  a  (S)/(I)  ratio  of  2100  with 
DON.  Phosphoribosyl-PP  amidotransferase,  another  enzyme  catalyzing 
the  transfer  of  the  amide  nitrogen  of  glutamine,  is  also  inhibited  compe- 
titively by  DON,  with  a  K^  of  0.019  mM,  and  much  more  weakly  by  aza- 
serine (Hartman,  1963  b).  A  slow  covalent  binding  of  DON  to  the  enzyme 
occurs  following  the  initial  reversible  attachment,  and  this  is  accelerated 
by  the  presence  of  phosphoribosyl-PP  and  Mg++  on  the  enzyme,  indicating 
that  the  active  site  for  the  reaction  of  glutamine,  or  the  binding  of  DON, 
is  partly  dependent  on  the  other  substrate  and  the  cofactor.  Blocking  of 
an  SH  group  prevents  the  attachment  of  DON  to  the  enzyme,  suggesting 
that  this  SH  group  is  catalytically  functional  in  the  nitrogen  transfer  and 
the  irreversible  binding  of  DON,  as  French  et  al.  (1963  b)  concluded  from 
their  work  with  azaserine  on  the  formylglycinamide  ribonucleotide  amido- 
transferase. 


334  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Glutamate  Transaminases 

This  enzyme  apparently  possesses  two  cationic  groups  properly  separated 
to  interact  with  glutamate  and  the  other  dicarboxylates,  because  it  is  inhib- 
ited best  by  glutarate  of  all  the  saturated  dicarboxylates,  as  shown  in  the 
accompanying  tabulation  (aspartate  =1.7  mM  and  a-ketoglutarate  =  6.7 
mM)  (Jenkins  et  al.,  1959).  This  is  one  instance  in  which  the  «-methyl 
analog  has  no  affinity  for  the  enzyme.  Very  similar  results  were  obtained 


Inhibitor  (40  mM) 

Malonate 

Succinate 

Glutarate 

Adipate 

Piraelate 

Suberate 

Maleate 

a-Methylaspartate 

a-Methylglutaniate 


by  Velick  and  Vavra  (1962),  glutarate  inhibiting  the  most  potently  of  the 
saturated  dicarboxylates;  from  their  values  for  K,  one  can  calculate  that 
glutarate  is  bound  0.85  kcal/mole  more  tightly  than  succinate,  and  0.32 
kcal/mole  more  tightly  than  adipate.  Of  the  three  phthalates,  o-phthalate 
is  the  most  inhibitory  {K^  =  5  mM),  m-phthalate  intermediary  (K^  =  8 
mM),  and  j^-phthalate  the  least  active  {K,  =  10  mM).  The  results  of  the 
studies  above  were  obtained  on  pig  heart  transaminase,  and  from  the  lim- 
ited data  reported  by  Goldstone  and  Adams  (1962)  it  would  appear  that 
the  enzyme  from  rat  liver  is  different  in  that  glutarate  is  about  10  times 
more  potent  than  maleate  as  an  inhibitor.  The  alanine:a-ketoglutarate 
transaminase  from  rat  liver  is  inhibited  moderately  and  competitively  by 
certain  amino  acids,  such  as  leucine  and  valine,  and  also  by  maleate,  but 
no  data  were  given  for  comparison  with  the  aspartate:  a-ketoglutarate  trans- 
aminase (Segal  et  al.,  1962).  Fluorooxalacetate  inhibits  the  latter  enzyme 
from  heart  competitively  with  respect  to  oxalacetate  (and  to  a-ketoglutarate 
and  aspartate  when  the  reverse  reaction  is  run),  K^,,  being  3.5mMforof- 
ketoglutarate  and  0.5  mM  for  aspartate,  and  /f,  being  0.1  mM  (Kun  et 
al.,  1960).  It  is  also  slowly  transaminated  to  fluoroaspartate  which  likewise 
inhibits  the  enzyme. 


Inhibition 

Relative 

-  AF  (kcal/mole) 

0 

<0.17 

31 

1.49 

72 

2.56 

62 

2.29 

0 

<0.17 

0 

<0.17 

78 

2.76 

35 

1.60 

0 

<0.17 

ARGINASE  335 

Other    Enzymes   in   Glutamate   Metabolism 

Analog  inhibition  has  been  reported  for  several  other  enzymes  involved 
in  the  metabolism  of  glutamate  but  quantitative  studies  from  which  struc- 
ture-action relationships  may  be  derived  have  not  been  made.  Some  of 
these  results  are  summarized  in  Table  2-13.  An  interesting  inhibitor  is  the 
convulsant  isolated  from  agenized  flour,  methionine  sulfoximine,  which  is 
inhibitory  to  the  incorporation  of  methionine  into  proteins  in  bacteria 
and  brain.  Since  these  actions  are  to  a  great  extent  antagonized  by  methio- 
nine, this  substance  has  generally  been  considered  as  a  methionine  antago- 
nist, but  glutamine  also  is  antagonistic.  Sellinger  and  Weiler  (1963)  have 
shown  that  methionine  sulfoximine  inhibits  brain  glutamine  synthetase 
competitively  with  respect  to  glutamate,  some  inhibition  being  seen  at 
0.01  mM  and  around  50%  inhibition  at  1  mM  with  low  glutamate  concen- 
trations {K^  =  0.05-0.064  mM).  The  relation  between  this  inhibition  and 
the  convulsant  action  is  not  clear  but  it  was  postulated  that  methionine 
sulfoximine  interferes  in  some  vague  manner  with  glutamine  synthesis 
in  an  intracellular  compartment  in  the  brain. 

ARGINASE 

The  hydrolysis  of  arginine: 

\                                            /            (+H2O)      +  /  \ 

C— NH— CHoCH^CH,— CH        ^ ^-^-   H3N— CHoCHoCHz— CH  +  CO 

+     <^  \         -  \  -  / 

HgN  COO  COO        H2N 

Arginine  Ornithine  Urea 

is  catalyzed  by  the  Mn++-activated  enzyme  arginase  and  is  a  step  in  the 
urea  cycle.  Arginase  is  inhibited  by  the  product  ornithine,  as  first  shown 
by  Gross  (1921)  and  confirmed  by  Bach  and  Williamson  (1942),  who  found 
that  the  inhibition  is  much  more  marked  in  liver  extracts  than  in  slices 
(50%  inhibition  given  by  5.3  mM  ornithine  in  extracts  and  by  15.9  mM 
in  slices  when  argmine  is  3.56  mM).  Edlbacher  and  Zeller  (1936)  noted  that 
several  amino  acids  inhibit,  but  ornithine  is  the  most  potent  with  lysine 
running  a  close  second. 

These  inhibitions  were  compared  and  subjected  to  quantitative  treat- 
ment by  Hunter  and  Downs  (1945)  in  the  publication  wherein  the  first 
use  of  the  single-curve  plot  (type  F)  was  made  (see  Chapter  1-5).  Orni- 
thine and  lysine  inhibit  competitively  {K ,  =  4.1  mM  and  4.8  mM,  respec- 
tively) but  other  amino  acids  are  usually  only  partially  competitive  (since 
the  DL-forms  of  the  inhibitors  used  and  only  the  L-isomers  are  active,  it  is 
likely  that  these  constants  should  be  halved).   Calculations  of  relative 


336 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


tf 


Oh 


c^  <N 


O 


o  ~ 


a>        Oi        ir- 


-tj         I— ( 


Xi 

J2 

^ 

o 

O 

e 

05 

05 

, . 

■o 

-73 

"^S 

O 

c 

C 

N 

C5 

c3 

cS 

-t^ 

tH 

tM 

'> 

^^ 

eS 

cS 

o 

^ 

M 

M 

_c 

v 

C 

fi 

c 

<D 

01 

.A 

tH 

>, 

>, 

c3 

cS 

< 

< 

05 

> 

-ki 


rO 


U;  o      c 


c    II 


"H       '^        -fc^        -Li 


^  rz  rz  .—  ."S 

"    ■"  :2  :2  :2  :2 

e    "^  IS  IS  IS  2 

S   —  e  c  c  c 


p.   5     c. 


o    ;5 


o 


&  c 


4J      O 


o    2 


^      c    ^  ^  '^  ^ 

S  S  S  2 

rit   CO    CO  CO  cc  CO 

^     O     CO  CO  CO  CO 


-  o°  2 

■S  o  .S  s? 

•S  t^  ^  Sj 

(N  ,2  '=^. 

|-H  iM  ,_,    lO 


S  *« 

O    (M 


.S  -a 

ft  -^ 

S  ic 
o  ^ 


s 

s 

c8 

g 

13 

S 

o 

J 

2 

o 

c 

o 

1 

2 

T3 

>> 

3 
O 

C 

-2 

g 
3 

5 

o3 

S 

3 

5 

3 

5 

j 

"ao 

>, 

w 

3 

5 

-< 

> 

>> 

S 

1 

Ph 

< 

_3 

u 

0 

«iX 

=Ci 

Q 

'O 

'^ 

j 

?^ 

u 

<ai. 

0 

Q 

13     O 


o 


eS 

-1-3 

"2 

o 

JS 

'x 

-^ 

(a 

o 

3 

S 

-2 

>> 

ou 

N 

c8 

aj 

3 

s 

_s 

■k^ 

c 

3 

c« 

-»j 

O 

3 

c 

5 

eS 


s  ^1 


O 


ARGINASE  337 

interaction  energies  from  the  constants  given  are  shown  in  the  following 
tabulation.  Two  types  of  inhibition  may  be  possible:  (1)  competitive  inhi- 
bition by  diamino  compounds  that  are  probably  oriented  as  the  substrate, 
and  (2)  partially  competitive  or  noncompetitive  inhibition  by  monoamines 


^  .  .,  .  Relative   —  Ji^  of  binding 

Inhibitor  ,,      ,,      ,  , 

(kcal/mole) 


L-Ornithine  3.82 

L-Lysine  3 .  72 

L-Norvaline  2 .  53 

L-Isoleucine  2 .  48 

L- Valine  1.98 

L-Cysteine  1 .  94 

L-Leucine  1.81 

L-a-AminobutjTate  1.40 

L-Phenylalanine  1 .  09 

L-Xorleucine  0 .  87 

L-Proline  0.78 

L- Aspartate  0 .  67 

L- Alanine  0.61 

L-Citrulline  0.47 

L-Serine  0.25 

L-Tryptophan  0.21 

L-Histidine  -0.82 

Glycine  —0.99 


that  may  be  oriented  otherwise.  Inhibitory  activity  increases  in  the  straight- 
chain  series  from  glycine  to  norvaline;  each  additional  methylene  group 
contributes  1.17  kcal/mole  to  the  binding  energy,  a  value  similar  to  that 
found  in  other  series,  and  undoubtedly  due  to  dispersion  forces.  It  is  odd 
that  there  is  a  sudden  and  marked  drop  in  the  affinity  on  adding  another 
methylene  group  to  form  norleucine.  As  pointed  out  by  Hunter  and  Downs, 
it  is  difficult  to  establish  structural  correlation  in  the  series  of  substituted 
alanines  (shown  in  the  accompanying  table);  why,  for  example,  is  cysteine 
bound  so  much  more  tightly  than  serine?  Substitution  in  the  a-amino  group 
(carbamyl  or  formyl)  always  reduces  the  activity;  hence  this  probably  con- 
stitutes one  binding  group.  The  experiments  were  done  at  pH  8.4  and 
therefore  all  carboxyl  groups  were  essentially  completely  ionized,  but  there 
would  be  some  variation  between  the  inhibitors  with  respect  to  the  fraction 
of  the  amino  groups  protonated  {pK^'s  for  these  inhibitors  run  from  8.2 
to  10.6). 


338  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


_,,,.,        ,  Relative  —  AF  oi  bmdim 

bubstituent  group 

(kcal/mole) 


—Imidazole  —0.82 

—Indole  0.21 

-OH  0.25 

-H  0.61 

-COO-  0.67 

-Phenyl  1.09 

-CH3  1.40 

-SH  1.94 


The  relative  sensitivities  to  these  amino  acids  probably  vary  with  the 
source  of  the  arginase.  For  example,  the  mouse  liver  enzyme  is  inhibited 
somewhat  more  strongly  by  L-lysine  than  L-ornithine,  and  even  D-lysine 
is  a  weak  inhibitor  (inhibitions  at  1  mM  are  47%  for  L-lysine,  42%  for  orni- 
thine, and  8%  for  D-lysine)  (Nadai,  1958).  Johnstone  (1958)  stated  that 
there  is  evidence  in  intact  ascites  cells  that  ornithine  can  interfere  with  the 
transport  of  arginine  into  the  cells,  as  well  as  inhibit  arginase  intracellularly, 
so  that  this  additional  site  of  inhibition  must  be  borne  in  mind. 

An  attempt  was  made  by  Sen  (1959)  to  determine  the  effects  of  L-lysine  in 
vivo  in  order  to  evaluate  its  possible  use  in  uremia.  Bilaterally  nephrec- 
tomized  dogs  show  an  increase  of  blood  urea  at  a  rate  of  15-16  mg/day  and 
die  in  80-84  hr.  When  L-lysine  is  injected  daily  and  intravenously  at  a 
dose  of  1  g,  the  blood  urea  rise  only  3-4  mg/day  and  the  animals  survive 
for  274-278  hr.  Thus  it  would  seem  possible  to  reduce  urea  formation  in 
vivo  with  this  competitive  inhibitor,  but  the  clinical  benefit  of  this  remains 
to  be  tested. 

L-AMINO   ACID   OXIDASES 

These  enzymes  from  snake  venoms  and  mammalian  tissues  oxidize  the 
L-isomers  of  the  monoamino-monocarboxylate  amino  acids,  the  substrate 
used  often  being  L-leucine.  The  snake  venom  enzyme  was  shown  by  Zeller 
and  Maritz  (1944)  to  be  inhibited  by  various  aromatic  sulfonates.  Some  of 
the  results  are  shown  in  Table  2-14,  from  which  it  may  be  seen  that 
p-nitrodiphenyl  sulfonate  is  the  most  potent  inhibitor  studied.  The  inhi- 
bitions seem  to  be  competitive  and  a  complex  between  the  sulfonate  group 
and  an  enzyme  amino  group  was  postulated.  Benzoate  is  a  rather  weak 
inhibitor  of  rat  kidney  L-amino  acid  oxidase,  10  mM  inhibiting  28% 
(Blanchard  et  al.,  1944).  However,  the  ammonium  ion  is  a  surprisingly  good 
inhibitor,  12  mM  producing  69%  depression  of  activity. 


L-AMINO   ACID   OXIDASES 


339 


Table  2-14 
Inhibition  of  Snake  Venom  l-Amino  Acid  Oxidase  by  Sulfonates" 


Inhibitor 


Sulfonate 
(mM) 


L-Leucine 

(mA/) 


Inhibition 


HO 


N=N- 


SO, 


2.9 


0.01 


//_\Vn=n^/   Vso; 


2.9 


0.01 


76 


2.9 


2.9 


2.9 


0.01 


0.01 


0.01 


66 


55 


50 


0,N 


0.33 


1.67 


31 


0,N 


//  \V_N„_//   V-so; 


0.07 


0.01 


21 


a  From  Zeller  and  Maritz  (1944). 


340  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

The  L-amino  acid  oxidase  from  the  hepatopancreas  of  Cardium  tubercula- 
tum is  more  specific  than  the  venom  or  kidney  oxidases,  since  many  L-amino 
acids  are  not  oxidized  but  are  inhibitory  (Roche  et  al.,  1959).  The  inhibitions 
by  16.7  mM  L-leucine  are  shown  in  the  following  tabulation.  At  pH  7.6 
the  inhibitions  are  competitive,  but  not  at  pH  9.2.  Apparently  this  enzyme 
can  complex  with  both  l-  and  D-isomers  although  in  no  case  is  a  direct 
comparison  possible. 


%  Inhibition  at: 

Amino  acid 

pH7.6 

pH9.2 

DL- Alanine 

89 

71 

D-Alanine 

39 

— 

L-Serine 

54 

73 

Glycine 

.51 

56 

L-Glutamate 

26 

— 

L- Proline 

24 

74 

L- Valine 

24 

30 

L-Threonine 

20 

14 

L- Aspartate 

13 

12 

D-Histidine 

12 

— 

D- Leucine 

5 

— 

D-AMINO  ACID  OXIDASE 

D-Amino  acid  oxidase  is  also  inhibited  by  certain  amino  acids,  but  no 
thorough  studies  have  been  reported.  The  enzyme  from  pig  kidney  oxidizing 
D-leucine  is  inhibited  by  DL-leucinamide,  DL-leucylglycine,  glycyl-DL- 
leucine,  and  DL-leucylglycylglycine,  but  not  glycylglycine  (Heimann- 
Hollaender  and  Lichtenstein,  1954).  It  is  interesting  that  the  oxidation 
of  D-phenylalanine  is  inhibited  by  DL-iV-ethylphenylalanine  and  the  oxida- 
tion of  D-leucine  by  DL-A^-ethylleucine,  since  A'-substituted  amino  acids  are 
usually  not  inhibitory  for  any  enzymes  acting  on  amino  acids.  D-Lysine  is 
a  good  inhibitor  of  this  enzyme  oxidizing  D-alanine  (Ky„  =  3.3  mM,  and 
K^  =  5  mM),  but  L-lysine  is  completely  inactive  (Murachi  and  Tashiro, 
1958).  The  D-amino  acid  oxidase  from  pig  kidney  with  glycine  as  a  substrate 
is  competitively  inhibited  by  L-leucine  {K,  =  1  mM)  (Neims  and  Hellerman, 
1962).  Pyruvate  not  only  competes  with  D-alanine  {K^  =  43  mM)  but  accel- 
erates the  photodecomposition  of  FAD  (Yagi  and  Natsume,  1964).  However, 
the  most  interesting  and  best  studied  inhibition  of  D-amino  acid  oxidase  is 
that  of  benzoate  and  its  derivatives,  and  the  opportunity  will  be  taken  in 
this  section  of  discussing  not  only  the  actions  of  the  benzoates  on  this  en- 
zyme but  also  on  other  enzymes  and  metabolism  in  general. 


D-AMINO   ACID   OXIDASE 


341 


Benzoates   and    Related    Compounds   on    D-Amino   Acid    Oxidase 

The  oxidation  of  D-alanine  in  slices  and  homogenates  of  rat  liver  and 
kidney  was  shown  to  be  markedly  inhibited  by  1  mM  benzoate  by  Klein  and 
Kamin  (1941).  A  preparation  of  D-amino  acid  oxidase  from  pig  kidney  was 
thus  obtained  and  found  to  be  inhibited  79%  by  0.1  mM,  this  being  rever- 
sible upon  dialysis.  Several  substituted  benzoates  are  also  inhibitory  but 
all  are  less  potent  than  benzoate;  benzamide  is  inactive.  The  inhibitions  of 
a  lamb  kidney  D-amino  acid  oxidase  by  benzoate  and  j9-amino-benzoate 
were  shown  by  Hellerman  et  al.  (1946)  to  be  competitive  with  substrate. 
The  rate  of  spontaneous  inactivation  of  the  apoenzyme  is  reduced  by  either 
substrate  or  FAD,  and  benzoate  was  shown  by  Burton  (1951  a)  to  have  a 
comparable  action,  indicating  combination  with  the  active  center. 

Before  discussing  the  more  detailed  mechanism  of  the  inhibition  we 
shall  turn  to  three  studies  providing  information  on  the  structural  require- 
ments for  inhibition.  Bartlett  (1948)  compared  many  substituted  benzoates 
(Table  2-15)  and  found  only  four  to  be  more  potent  inhibitors  than  benzoate, 

Table  2-15 
Inhibition  of  Pig  Kidney  d-Amino  Acid  Oxidase  by  Substituted  Benzoates  " 


Substituent 

Relative 

—  Ji^  of  binding 

(kcal/mole) 

ortho 

meta 

para 

F 

4.26 

6.25 

CI 

3.27 

6.81 

5.39 

Br 

2.11 

6.40 

5.05 

I 

2.84 

5.39 

3.90 

OH 

4.55 

5.20 

3.27 

NHa 

4.26 

4.26 

2.84 

NO, 

2.52 

5.12 

4.90 

CH3 

2.11 

6.25 

5.12 

OCH3 

2.11 

3.90 

3.78 

COOH 

1.85 

2.11 

2.84 

None 

— 

5 .  57 

— 

"  The  substrate  is  DL-alanine  at  30  mM  and  the  pH  7.6.  Binding  energies  calculated 
from  concentrations  for  50%  inhibition.   (Data  from  Bartlett,   1948.) 


pure  competitive  inhibition  with  respect  to  substrate  being  observed. 
J.  R.  Klein  (1953,  1957)  demonstrated  inhibition,  often  potent,  by  various 
aromatic  carboxylates  (Table  2-16),  and  Frisell  et  al.  (1956)  provided  further 


342  2.  ANALOCxS  OF  ENZYME  REACTION  COMPONENTS 

Table  2-16 
Inhibition   of  Pig   Kidney   d-Amino  Acid   Oxidase  by  Aromatic   Acids  " 


Inhibitor 


Ki  Relative  —  AF  oi  binding 

(mM)  (kcal/mole) 


2-Chloromethyl-5-hydroxy-l,4-pyrone 

Kojate 

TO-Toluate 

Pjrrrole-2-carboxylate 

Benzoate 

Furan-2  -acrylate 

p-Toluate 

Nicotinate 

Cinnamate 

Furan-2-carboxylate 

l,2-Pyrone-5-carboxylate  (coumalate) 

Indole-3-acetate 

Hydrocinnamate 

Phenylacetate 

l,4-Pyrone-2,6-dicarboxylate  (chelidonate) 

o-Toluate 


0.004 

7.68 

0.021 

6.65 

0.022 

6.62 

0.026 

6.52 

0.048 

6.14 

0.052 

6.08 

0.08 

5.82 

0.11 

5.63 

0.56 

4.62 

0.60 

4.58 

0.60 

4.58 

2.3 

3.74 

5.5 

3.21 

6.4 

3.12 

15 

2.59 

29 

2.  IS 

"The  substrate  is  DL-alanine  and  the  pH  8.0-8.3.  (Data  from  J.  R.  Klein,  1953, 1957.) 

data  on  aliphatic  and  heterocyclic  carboxylates  (Table  2-17).  Some  of  the 
conclusions  regarding  relations  between  structure  and  inhibition  derived 
from  these  investigations  will  be  summarized. 

(1)  A  negatively  charged  anionic  group  is  necessary  for  activity.  This  is 
seen  from  the  lack  of  inhibition  by  benzamide,  nicotinamide,  and  cinnama- 
mide.  It  is  likely  that  the  COO"  group  of  the  inhibitors  reacts  with  the  en- 
zyme cationic  site  normally  reacting  with  the  amino  acid  C00~  group.  A 
SO3""  group  can  replace  the  COO"  but  is  less  effective. 

(2)  Klein  has  emphasized  the  importance  of  a  positive  charge  at  a  distance 
from  the  COO"  group  approximating  the  separation  in  amino  acids.  Most 
of  the  potent  inhibitors  can  be  written  in  structures  possessing  such  a 
positive  charge  by  virtue  of  resonance  effects.  Benzoate,  for  example, 
resonates  between  the  following  structures: 

-  -J- 

/       \         /  /       \ 


v>\     ^% 


0 

/   \ 

0 

/   \ 

0 

/ 

//  ^ 

V / 

/  \ 

/ 

_ 

)=c  - 

+(   ) 

=  c 

\ 

\   / 

'^   \ 

\   / 

\ 

0 

\   / 

0 

\   / 

0 

D-AMINO    ACID   OXIDASE 


343 


Table  2-17 
Inhibition  of  Lamb  Kidney  d-Amino  Acid  Oxidase" 


Inhibitor  (3  mM) 


Structure 


%  Inhibition 


Crotonate 

Butyrate 

Dimethylacrylate 


Isovalerate 


Fumarate  and  maleate 


CHj—  CH=CH— COO' 

CH3—  CH2—  CHj—  COO' 

H,C 
^  \ 

C=CH— COO 

H3C 

H3C 

CH— CH,— COO" 
H3C 

'OOC  -  CH=  CH—  COO' 


99 
0 


70 


Cinnamate 


r     y—  CH=CH— COO' 


100 


Hydrocinnamate 


CH,— CHo^COO 


55* 


Cinnamamide 


(^      y — CH  =  CH-CONH2 


Phenylacetate 


CHo—  COO 


15 


/>-Toluenesulfonate 


H,C- 


SO, 


34 


Sulfanilate 


-h,nV/   V 


so. 


344 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 
Table  2-17  (continued) 


Inhibitor  (3  mM) 


Nicotinate 


Structure 


N- 
//      W 


COO 


Inhibition 


65 


Nicotinamide 


N — V 


CONH, 


2-Furoate 


O^     ^COO 


W 


100 


Indol  e  -  2  -  ca  rbox  yl  ate 


H 

N^       COO' 


100 


Benzoate 

/»-Methoxybenzoate 
o  -  Methoxybenzoate 
/J-Carboxybenzoate 
o - Carboxybenzoate 
Hz-Carboxyben^oate 

Cyclohexanecarboxylate 


€y 


COO 


(      V-coo' 


100 

100 

67 

74 

3 

20 

53* 


Hydroquinone 

Tribromophenol 

/>-Aminophenol 

Triiodophenol 

Catechol 

3, 5-Dihydroxyphenol 

2, 3-Dihydroxyphenol 

Resorcinol 


73 

40 

30 

19 

19 

5 

0 

0 


"  The  substrate  is  o-alanine  at  6.25  mM  and  the  pH  8.3.  The  inhibitions  marked 
with  an  asterisk  are  at  least  partially  due  to  contamination  with  the  unsaturated 
compounds.     (Data  from  Frisell  (7  (v/.,    1956.) 


D-AMINO    ACID    OXIDASE 


345 


and  some  of  the  other  inhibitors  may  be  written  as: 


,/0 


Nicotinate 


^O 


2-Furancarboxylate 


HOH,C 


W    A-""-""=V 


HCH^=CH-CH=C:f    - 
O 


O 


Cinnamate 


Crotonate 


Indole  -  2  -  carboxylate 


Such  structures  would  be  less  possible  or  impossible  for  hydrocinnamate, 
l,4-pyrone-2,6-dicarboxylate,  cyclohexanecarboxylate,  phenylacetate,  and 
some  of  the  other  weaker  inhibitors.  The  orientation  relative  to  the  sub- 
strate, according  to  Klein,  would  be  represented  as: 


+ 
/C-CC 


Substrate 


Inhibitor 


where  X  is  carbon,  oxygen,  or  nitrogen. 

(3)  Frisell  and  his  co-workers,  on  the  other  hand,  emphasized  the  impor- 
tance of  a  double  bond  near  the  carboxylate  group.  Saturation  of  crotonate, 
cinnamate,  dimethylacrylate,  and  benzoate  definitely  reduces  the  inhibition. 
They  postulated  that  this  double  bond  might  correspond  in  position  on  the 
enzyme  to  the  double  bond  of  the  iminoketonic  form  of  the  dehydrogenated 
product,  and  thus  according  to  their  theory  the  structural  correspondence 
would  be: 


Product 


— C> 


^c-c: 


,o 


Benzoate 


-'^--c-cr" 


Aliphatic  carboxylate 


(4)  These  two  theories  are  not  incompatible,  since  we  have  seen  that 
the  presence  of  a  positive  charge  generally  depends  on  resonance,  which  in 
turn  requires  double  bonds  and  either  conjugation  or  hyperconjugation. 


346  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

(5)  The  different  potencies  of  the  substituted  benzoates  would  be  explained 
on  the  basis  of  the  effects,  such  groups  would  have  on  resonance  and  the 
magnitude  of  the  positive  charge,  but  in  addition  other  factors  must  play 
a  role,  for  example  the  dipole  moments  of  the  ring  X  bonds  and  the  possible 
interactions  of  the  substituent  groups  themselves.  One  might  expect  ortho 
groups  to  decrease  the  binding  and  it  is  true  that  all  are  less  inhibitory  than 
benzoate.  Bartlett  pointed  out  that  the  inhibition  increases  markedly  with 
the  electronegativity  of  the  halogens.  It  is  very  interesting  to  attempt  to 
interpret  the  results  in  Table  2-15  but  without  more  knowledge  about  the 
nature  of  the  binding  any  hypotheses  must  be  vague  for  the  time  being. 

(6)  It  is  unlikely  that  the  degree  of  ionization  of  the  carboxyl  group  is 
important  here  because  all  of  this  work  was  done  between  pH  7.6  and  8.3 
where  a  negligible  fraction  is  undissociated.  The  p/iT^'s  of  all  the  substituted 
benzoates  tested  run  from  2.85  to  4.65.  However,  the  series  of  phenols 
studied  by  Frisell  et  al.  (1956)  may  have  to  be  considered  in  terms  of  the 
ionization  of  the  OH  groups  when  relating  structure  to  activity. 

(7)  The  K-^  for  some  selected  inhibitors  were  determined  by  Frisell 
et  al.  (1956)  and  from  these  the  relative  binding  energies  may  be  estimated 
(see  accompanying  tabulation).  The  1.45  kcal/mole  difference  between  cin- 


Source  of  enzyme  Inhibitor 


A",         Relative  —  AF  oi  binding 
{mM)  (kcal/mole) 


Lamb  kidney 

Cinnamate 

0.076 

5.85 

Crotonate 

0.038 

6.28 

Pig  kidney 

Cinnamate 

0.22 

5.20 

Benzoate 

0.021 

6.65 

Indole-2-carb 

axylate 

0.0034 

7.77 

namate  and  benzoate  (it  is  1.52  kcal/mole  from  Klein's  data)  may  be 
attributed  to  a  more  satisfactory  location  of  the  positive  charge  in  benzo- 
ate. The  lesser  affinity  of  the  enzyme  for  nicotinate  compared  with  benzoate 
(0.51  kcal/mole)  could  be  due  to  the  reduction  in  resonance  brought  about 
by  the  asymmetry  of  the  former.  If  the  1.94  kcal/mole  difference  between 
pyrrole-2-carboxylate  and  furan-2-carboxylate  is  related  to  the  different 
amounts  of  positive  charge  on  the  nitrogen  and  oxygen  atoms,  this  would 
not  be  surprising  since  pyrrole  should  resonate  more  effectively.  Finally, 
if  the  positive  charge  theory  is  valid,  the  3.02  kcal/mole  difference  between 
benzoate  and  phenylacetate  might  indicate  the  amount  of  binding  contri- 
bution from  the  positive  charge.  The  positive  charge  may,  of  course,  be 
stabilized  somewhat  by  an  enzyme  anionic  site  with  which  it  interacts. 


D-AMINO   ACID   OXIDASE  347 

Although  the  concept  that  these  inhibitors  bind  to  the  enzyme  and  com- 
pete with  the  substrate  is  very  reasonable,  some  recent  evidence  may 
indicate  that  the  situation  is  a  little  more  complex.  The  values  of  K^  for 
benzoate  should  be  the  same  for  every  substrate,  according  to  the  classic 
treatment,  but  Klein  (1956,  1960)  has  found  that  they  are  not,  although 
all  the  1/v  — 1/(S)  plots  are  definitely  competitive.  The  A;/s  may  vary  as 
much  as  almost  3-fold  (see  tabulation  below).  One  possibility  is  that  in 


Substrate 

Relative   F,„ 

Km 

[mM) 

{mM) 

Alanine 

1.00 

6.3 

0.059 

Proline 

1.66 

5.8 

0.070 

Phenylalanine 

1.39 

14.0 

0.092 

Valine 

0.60 

4.6 

0.096 

Isoleucine 

0.79 

4.1 

0.104 

Methionine 

1.32 

5.3 

0.163 

the  aqueous  extracts  of  pig  kidney  there  are  different  oxidases  for  each 
substrate  but  Klein  prefers  to  assume  that  the  inhibitors  may  react  with 
the  ES  complex  to  release  the  substrate: 

ES  +  I  ^  EI  4-  S 

The  equilibrium  constant  K  =  (ES)  (I)/(EI)  (S)  depends  on  the  substrate 
used  and  in  the  usual  rate  equation  for  competitive  inhibition,  {l)KJK^ 
would  be  replaced  by  {l)IK.  Since  K  =  KJK„  the  determined  values  of  K 
should  be  inversely  proportional  to  K,.  They  are  not  inversely  proportional 
to  K„„  but  is  K„,  =  K,  in  this  case?  It  may  be  noted  that  this  mechanism  is 
not  quite  the  same  as  uncoupling  inhibition,  since  there  an  ESI  complex  is 
formed  and  the  substrate  is  not  forced  off.  I  must  admit  that  I  cannot 
easily  visualize  how  a  competitive  inhibitor  can  actively  displace  an  enzyme- 
bound  substrate  molecule. 

Although  Hellerman  et  al.  (1946)  reported  the  inhibition  of  D-amino  acid 
oxidase  by  benzoate  to  be  independent  of  FAD  concentration,  there  is 
more  recent  evidence  that  certain  aromatic  carboxylates  and  phenols  not 
only  can  compete  with  substrate,  but  can  also  either  compete  with  FAD 
or  complex  directly  with  FAD  (Yagi  et  al.,  1957,  1959,  1960).  The  constants 
for  each  of  these  reactions  are  given  for  a  few  of  these  inhibitors  in  the  fol- 
lowing tabulation.  Only  the  carboxylates  compete  with  substrate,  while 
the  phenols  act  by  the  other  two  mechanisms;  the  substances  with  both 
COO-  and  OH  groups  react  in  all  three  ways,  although  competition  with 
the  substrate  is  the  most  important.  Yagi  and  his  group  have  recently 


348  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


K^  (mi/) 


Inhibitor 


Competition 

Competition 

Complex 

with  substrate 

with  FAD 

with  FAD 

0.0145 

0.35 

1.05 

0.65 

6.35 

15 

75 

0.305 

7.25 

— 

— 

0.85 

— 

— 

9.15 

— 

— 

56.5 

32 

— 

28 

110 

— 

5.0 

0.5 

— 

3.9 

0.02 

— 

2.3 

0.04 

Benzoate 

Salicylate 

^-Aminosalicylate 

p-Arainobenzoate 

p-Nitrobenzoate 

Aniline 

Phenol 

m-Aminophenol 

p-Nitrophenol 

2,4-Dinitrophenol 

2,6-Dinitrophenol 

2,4,6-Trinitrophenol  —  0.64  0.012 


crystallized  the  complex  of  apoenzyme,  FAD,  and  benzoate,  and  found 
equimolar  amounts  of  each  component  present. 

It  would  be  interesting  to  have  more  data  on  the  effects  of  these  inhibitors 
on  L-amino  acid  oxidases.  Benzoate  inhibits  the  L-amino  acid  oxidases 
from  rat  kidney  (Blanchard  et  al.,  1944)  and  snake  venom  (Zeller  and  Maritz, 
1945),  but  does  not  inhibit  the  enzyme  from  Neurospora  even  at  10  niM 
(Burton,  1951  b). 

Benzoate   on   Other   Enzymes   and    Metabolism 

The  endogenous  ammonia  formation  and  respiration  of  rat  kidney  slices 
are  inhibited  52%  and  66%,  respectively,  by  25  mM  benzoate  (Herner, 
1944).  One  might  assume  that  the  former  might  be  attributed  to  inhibition 
of  amino  acid  oxidases,  but  this  is  unlikely  because  benzamide,  phenyl- 
acetate,  and  /?-phenylpropionate  are  even  more  potent  inhibitors  than 
benzoate.  The  deamination  of  various  l-  and  D-amino  acids  in  kidney  is, 
however,  strongly  inhibited  by  benzoate,  at  least  in  part  competitively. 

The  inhibition  of  respiration  by  benzoate  was  first  observed  by  Grif- 
fith (1937)  in  slices  and  minces  of  various  tissues  in  the  presence  of  glucose, 
but  no  analysis  of  the  site  of  action  has  been  published,  although  in  con- 
nection with  recent  studies  on  the  salicylates  some  effects  on  mitochondria 
have  been  investigated.  Benzoate  inhibits  the  oxidations  of  citrate,  a- 
ketoglutarate,  and  succinate  in  rat  kidney  homogenates  but  is  invariably 
less  active  than  salicylate  (E.  H.  Kaplan  et  al.,  1954);  and  neither  the  Og 
uptake  nor  the  P  :  0  ratio  is  affected  markedly  in  rat  liver  mitochondria, 
although  salicylate  uncouples  strongly  (Brody,  1956).  Endogenous  phosphor- 


D-AMINO    ACID    OXIDASE  349 

ylation  in  liver  mitocliondria  is  inhibited  55%  by  10  mM  benzoate  (Wein- 
bacli,  1961).  One  can  conclude  from  this  limited  material  that  benzoate  is 
certainly  a  weak  inhibitor  of  cycle  oxidations  and  phosphorylations.  Bosund 
(1959,  1960  a,  b)  has  investigated  the  effects  of  benzoate  on  the  metabolism 
of  glucose  and  pyruvate  in  Proteus  vulgaris  in  attempting  to  elucidate 
the  mechanisms  for  the  bacteriostatic  activity.  There  is  no  interference 
with  glucose  metabolism  to  the  acetate  level,  and  acetate  was  found  to 
accumulate.  The  respiratory  quotient  is  increased  from  1.24  to  1.82  by 
benzoate  during  the  oxidation  of  pyruvate  and  the  Oa/pyruvate  ratio  is 
decreased.  The  oxidation  of  pyruvate  in  yeast  is  quite  strongly  inhibited 
by  benzoate  (50%  at  0.4  mM),  especially  at  low  pH's  where  penetration  is 
better,  but  acetate  oxidation  is  less  sensitive.  It  is  quite  possible  that  these 
inhibitions  play  a  role  in  the  suppression  of  growth,  which  for  yeast  requires 
5  mM  benzoate  at  pH  5.1  and  60  mM  at  pH  6.  It  is  clear  that  much  more 
work  must  be  done  before  the  mechanisms  of  respiratory  inhibition  are 
understood. 

Benzoate  can  also  interfere  in  lipid  metabolism,  as  demonstrated  many 
years  ago  by  Jowett  and  Quastel  (1935  a,  b),  but  the  mechanism  is  still 
unknown.  In  liver  slices  it  was  claimed  that  benzoate  at  around  0.5-2 
mM  inhibits  specifically  the  oxidation  of  fatty  acids,  and  the  oxidation  of 
crotonate  63%  at  1  mM.  There  is  progressively  less  effect  on  the  higher 
fatty  acids,  little  inhibition  of  decanoate  being  observed.  It  is  possible 
that  the  benzoate  ring  simulates  the  aliphatic  chains  of  butyrate  or  croton- 
ate enabling  it  to  compete  with  these  substrates  for  some  enzyme;  it  would 
be  interesting  to  know  if  benzoate  can  participate  in  any  of  these  reactions 
(e.  g.,  if  benzoyl-CoA  is  formed)  and  deplete  the  systems  of  some  cofactor. 
Benzoate  is  a  weak  inhibitor  of  tyrosinase  (Ludwig  and  Nelson,  1939), 
chymotrypsin  (Foster  and  Niemann,  1955  b),  p-aminobenzoate  acetylation 
(Koivusalo  and  Luukkainen,  1959),  and  NADPH  dehydrogenase  (Kasa- 
maki  et  al.,  1963);  it  does  not  effect  shikimate  dehydrogenase  (Balinsky  and 
Davies,  1961  b)  or  D-glutamate  oxidase  (Mizushima  and  Izaki,  1958)  at 
1  mM,  or  a-ketoisocaproate  decarboxylase  at  4  mM  (Sasaki,  1962). 

Kojic  Acid 

The  potent  inhibition  of  D-amino  acid  oxidase  by  kojic  acid  is  interesting 
in  light  of  the  central  nervous  system  effects  observed  in  dogs,  rabbits,  and 
rats,  namely,  ataxia,  excitement,  and  convulsions  (Friedemann,  1934). 
Kojic  acid  was  first  isolated  by  Saito  in  1907  from  Aspergillus  oryzae  and 
has  since  been  found  in  many  species  of  Aspergillus.  It  is  a  weak  antibiotic, 
inhibiting  growth  of  most  bacteria  at  2-15  mM,  but  is  particularly  active 
against  Leptospira,  complete  growth  inhibition  being  observed  at  0.007  m.N 
(Morton  et  al.,  1945).  Toxic  effects  are  produced  in  dogs  by  150  mg/kg  and 
in  mice  by  250  mg/kg  when  injected  parenterally;  the  LD50  for  mice  is 


350  2.  ANALOGS  OF  ENZYME  EEACTION  COMPONENTS 

1.5-2.0  g/kg.  Leucocytic  activity  and  phagocytosis  are  not  affected  by  18 
raM  kojic  acid.  A  biochemical  study  was  undertaken  by  Klein  and  Olsen 
(1947),  who  found  that  the  respiration  of  muscle  and  heart  mince  is  resistant 
to  kojic  acid,  whereas  10  mM  suppresses  the  respiration  of  liver  40%, 
kidney  20%,  and  brain  15%.  The  convulsant  dose  corresponds  to  a  tissue 
concentration  around  4-50  mM.  The  oxidation  of  both  l-  and  D-amino 
acids  in  liver  homogenates  is  quite  strongly  inhibited  in  a  competitive 
fashion:  for  example,  50%  inhibition  is  given  by  0.04  niM  for  L-methionine 
and  by  0.12  mM  for  l-  and  D-phenylalanine.  Xanthine  oxidation  is  also 
inhibited  (50%  at  0.7  mM).  It  was  suggested  that  kojic  acid  may  be  an 
inhibitor  of  flavin  enzymes,  and  it  is  possible  that  some  direct  complexing 
with  FAD  may  occur.  Nevertheless,  FAD  does  not  influence  the  inhibition 
of  D-amino  acid  oxidase.  There  is  no  inhibition  of  the  oxidation  of  succinate, 
tyramine,  L-proline,  choline,  or  urate  at  5  mM  kojic  acid.  Although  the 
metabolic  effects  are  interesting,  it  is  impossible  to  correlate  any  of  these 
inhibitions  with  either  the  central  effects  in  animals  or  the  bacteriostatic 
activity. 

ANALOG    INHIBITION    OF  THE   METABOLISM 
OF  VARIOUS   AMINO   ACIDS 

Many  reports  indicate  the  influence  of  analogs  on  various  enzymes  con- 
cerned with  amino  acid  metabolism,  but  in  most  cases  insufficient  work 
has  been  done  to  draw  clear  conclusions  about  the  mechanism  of  the  bind- 
ing to  the  enzymes.  It  will  suffice  to  present  some  of  the  results  in  Table 
2-18.  Probably  many  of  these  inhibitions  are  competitive  but  they  have 
been  so  indicated  only  when  graphical  analysis  has  shown  this  to  be  true. 
A  few  of  these  inhibitions  may  be  significant  in  feed-back  control  or  in  the 
general  regulation  of  amino  acid  metabolism. 

Generally  speaking,  there  are  certain  types  of  amino  acid  analog  that 
have  proved  to  be  effective  inhibitors.  Mcllwain  (1941)  pointed  out  that 
aminosulfonate  analogs  of  amino  acids  are  frequently  bacteriostatic  and 
that  this  inhibition  is  reduced  by  adding  the  normal  amino  acids.  If  staphy- 
lococci are  trained  to  be  independent  of  exogenous  amino  acids,  the  a- 
aminosulfonates  no  longer  inhibit.  However,  the  exact  sites  of  action 
of  these  analogs  have  not  been  determined.  Umbreit  (1955  b)  has  discussed 
the  general  inhibitory  properties  of  the  a-methylamino  acids  and  pointed 
out  that  the  inhibitions  are  often  competitive  only  for  a  short  interval 
if  both  substrate  and  inhibitor  are  present,  whereas  they  are  noncompetitive 
if  the  inhibitor  is  added  first.  This  is  a  matter  of  terminology;  the  inhibitions 
are  probably  competitive  but  appear  to  be  noncompetitive  because  of  the 
very  high  affinities  of  some  analogs  for  the  enzymes  (an  inhibition  can  be 
competitive  even  though  irreversible  but  the  substrate  must  be  given  an 


ANALOG  INHIBITION  OF  THE  METABOLISM  OF  VARIOUS  AMINO  ACIDS    351 

opportunity  to  compete).  A  third  group  of  interesting  analogs  is  the  halogen- 
substituted  amino  acids,  which  are  often  potent  inhibitors  of  protein  syn- 
thesis and  growth.  The  fluoroamino  acids  are  particularly  active.  The 
m-,  0-,  and  p-fluorophenylalanines  all  inhibit  the  formation  of  the  adaptive 
maltase  in  yeast,  the  last  being  the  most  potent  (Halvorson  and  Spiegelman, 
1952).  On  the  other  hand  the  p-chloro-  and  js-bromophenylalanines  do  not 
inhibit.  The  incorporation  of  phenylalanine  and  other  amino  acids  into 
ascites  cell  proteins  is  competitively  inhibited  by  o-fluorophenylalanine; 
this  is  in  part  due  to  a  depression  of  transport  into  the  cells  and  in  part  due 
to  block  of  some  unknown  steps  in  the  incorporation  (since  labeled  phenyl- 
alanine accumulates  in  cells)  (Rabinovitz  et  al.,  1954).  There  is  also  an  inhi- 
bition of  protein  synthesis  in  rat  liver  in  vitro  by  the  fluorophenylalanines, 
this  leading  to  a  net  breakdown  of  tissue  proteins  since  the  constant  bal- 
ance of  synthesis  and  degradation  is  disturbed  (Steinberg  and  Vaughan, 
1956).  a-Amino-/?-chlorobutjTate  is  an  analog  of  valine  and  inhibits  valine 
incorporation  into  rabbit  reticulocyte  proteins,  including  hemoglobin;  it 
was  suggested  that  the  analog  enters  a  precursor  protein  which  is  unable 
to  assume  the  proper  configuration  of  hemoglobin  and  thus  there  is  accu- 
mulation of  protein  intermediates  (Rabinovitz  and  McGrath,  1959).  As 
pointed  out  previously,  some  of  these  analogs  are  incorporated  into  cell 
proteins.  p-Fluorophenylalanine-C^*  is  incorporated  into  the  proteins  of 
muscle,  blood,  and  liver  when  fed  to  rabbits,  this  being  a  replacement  of 
phenylalanine  (Westhead  and  Boyer,  1961).  The  replacement  of  phenyl- 
alanine in  aldolase  is  25%  and  in  3-phosphoglyceraldehyde  dehydrogenase 
16%,  and  in  each  case  the  enzyme  activities  are  normal.  Despite  this  ap- 
preciable incorporation,  the  rabbits  suffer  no  obvious  biochemical  or  physi- 
ological disturbances,  so  that  mammals  may  well  differ  from  microorgan- 
isms in  the  response  to  this  analog. 

Feedback  inhibition  in  the  biosynthetic  pathways  of  amino  acids  is  an 
important  aspect  of  regulation  but  we  can  touch  only  briefly  on  this  prob- 
lem. An  interesting  example  of  this  has  been  studied  in  connection  with 
the  synthesis  of  histidine,  since  the  enzyme  inhibited  is  the  first  in  the 
pathway  and  catalyzes  a  reaction  not  involving  substrates  structurally 
similar  to  histidine  (Martin,  1963).  This  enzyme  is  phosphoribosyl-ATP 
pyrophosphorylase  and  the  reaction  catalyzed  is: 

5'-P-ribosyI-PP  +  ATP  :^  N-l-(5'-P-ribosyl)-ATP  +  PP 

Histidine  is  a  surprisingly  potent  and  specific  inhibitor  with  K,  =  0.1  mM. 
Related  compounds  inhibit  weakly  or  not  at  all;  2-methylhistidine,  for 
example,  exhibits  weak  inhibition  with  K^  =  2.4  mM.  The  inhibition  varies 
with  the  pH  but  maximal  inhibition  is  exerted  at  physiological  pH.  HgClg 
at  0.03  mM  does  not  inhibit  the  enzyme  but  blocks  almost  completely  the 
inhibition  by  histidine.  This  coupled  with  the  fact  that  the  inhibition  by 


352 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


P5 


CC   73 


A   Cm 
>>     O 

H 


S 

N 

c 


GO     05 

-H      (T^      l> 


o  o  o 


3 

s 

O 

s 

o 

CO 

'o 

o 

tJ 

c 

c 

<D 

>, 

S 

■g 

C 

<0 

<? 

<5 

"1 

a 

a 

hJ 

j 

ij 

h!] 

^ 

-^  .s 


3  03  t- 

o  o  ce 

to  CO  Kj 

I— I  t— I  K* 


M 


^ 

>. 

X 

O 

(U 

x> 

C 

kl 

<D 

4) 

T! 

l-J 

«    Xh 

^  " 


O    O    O    O    <M 


r^f»f^t^i--.i--.t>t-i>i>    OO 


P   O 


c 

"a 

c 
■3 

o 


c 

^      cS 

o  jh 

S     ft 


S    -Si  <s 


'2 

eS 

s 

O 

c 
S 


T3     Is 


ANALOG  INHIBITION  OF  THE  METABOLISM  OF  VARIOUS  AMINO  ACIDS   353 


e8    05 


CO 

CO 

'0   ^ 

eS 

o3 

_   TS 

-  "O 

C     b. 

c    b 

o    <o 

o    « 

CO  j: 

aj    ^ 

e    ft 

fi     ft 

:s  2 

^    1 

73 


>> 

^ 


C^  -  Pc;  --^  f^  -  j^  ^  (^  ^ 


©O-HCOOMMCOOO  0I>0  O  (NTt''*'*-*OpO0  O-hgOCO 

o   CO         CO     CO   CO         CO 

ll||l|||00  OOO  O  (M(M(M(N(MC^'M  rt^_^ 


c 

ft  S                           3 

a  ^ftx3ft><:2                    ® 

-^  c            •-ft'oSft'^e            -g.s-S 

>>  ft        oi=ip>>^cP>.g'c  ^^-sSib        c        c 

»o  H        H  Pm    ^-  ^    ^  -^     >^-  «    >:  -^  '«  '«    i  :S    c  -S  •=        5 

a  Q        QQoa       Wuoc        K  ^  ^=t  O  ^  Pi  ^  O       O 


•1=     »  "o. 


ft  -ks 


■S     2  .S  -ta  -»i  §"    S 

S3  S  e|  c3  ^    ^ 

«  O  (23  05  ^  ^ 


»    2 


ffi 


354 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


tf 


fc< 


Si 


ft 


T3 

C 

^— ' 

cS 

c 

u 
O 

o 

c 

o 

c 

> 

p 

w 

-H    lO    o    o 

IC    (M    lO    00 

rt    Csl    (M 


o  o  o  o  o  o  o 


C    -i:^     S     «     <»    .S     r* 

S    E?  .5  -C  -S  -3    ^ 


-—     >^    o  •  0) 


9    5«    o 


^    O    ^.CC    ^    >    ^ 
Q     Q   O     Q     Q     O     O 


Cq 


■<* 

^ 

CO 

73    ^ 

C5 

e8    »C 

'e 

C    ^ 

V 

•-    Si 

!>. 

II 

« 

^ 

|6 

1 

>o 

<>J 

<:d 

C<l 

O 

o 

-* 

Tt* 

-* 

-* 

■* 

% 

C^ 

d 

(M 

f>\ 

C<1 

d 

ec  CO  >c         CO  lo 

'o^fod—iMCOfM 


O 


O 


O 


o 


_g 

"2 

4i 

^ 

"? 

'c 

3 

"2 

Q 

^ 

<o 

T. 

eS 

^ 

1 

3 

">> 

<I> 

A 

">> 

.2 

ft 
O 

45 

cS    j2 

IS 

c 

o 

12 

<  -S 

H 

c 

3 

H 

1      o 

(N 

^ 

S 

*? 

0  o 

<^ 

<^ 

9, 

cii. 

.a    ro 

)M 

C        !« 

"2     M 

JS  '>. 

^   2 

a 

ce    y. 

's  -^ 

^^     o 

--S     ee 

>.  2 

b  > 

f>J 

?  '« 

9.    '43 

N 

PM  ^ 

CL| 

ANALOG  INHIBITION  OF  THE  METABOLISM  OF  VARIOUS  AMINO  ACIDS    355 


« 

-Q 

(^ 

(^ 

"   in 

05 

05 

s  ^ 

^ 

^ 

CS      -H 

T3 

S 

-O      « 

tH          , 

C 

C       fl 

c3 
a: 

1 

GQ 

03      o 

CO        OS 

s 

T3 

S  2 

cS 

"o 

c6     o 

< 

o 

^o 

O  OS 

O  •-H  Oi 

O  O    Oi    Tf*    rt         t^    CD 


O    rt    (M    O 


lOt^CO         <lDt^(M>OTtf(N         0>OC<I         100505    10  I        I        I  0»000 

■>*M-*i         f0»O<M-*CO<M         »OCOM         lO  >0  III  OOI><M 


(M«0  C<ICO  lOlOlO         -^ 


CO    CO    CO    o 

oooooo        I     I     I       oddd 


O  O   Q       O  O 


2  9 


s  =1  i    i     ■&! 

'-^  >>         rt     O     O 


3 

ft 
o 

01 

-0 

'>^ 

d 

t^, 

s 

o 

>. 

c^ 

<i 

H 

H 

ra     cc         J-        o<    >,    a 


ft 

ft 

's 

^ 

« 

s 

!«! 

-2 

X 

a> 

jj 

0) 

00 

C 

HP 

c3 

e3 

1 

O 

Ph 

'o 

SI 

cS 

1— 1 

1 

3 

o 

"3 

cS 
ft 

CO 

< 

J 

Q 

J 

1— 1 

►J 

ij 

fe 

w 

OJ 

p 

73 

»c 

a 

Oi 

cS 

c 

''■— ' 

4J 

fl 

3 

Cj 

Tl 

Ih 

c 

o 

o 

;^ 

o  o  o  o 

OOO 

OOO 

o  ^ 

eS 

C 

(1> 

O 

rrt 

ft 

-t:^ 

o 

^ 

crt 

ft 

ft 

o 

cc 

c 

ai 

>■. 

w 

X 

ri*. 

n 

t: 

Q 

>. 

fO 

^  pq  pq  pq  a,   *^  CM 

§s  >>^       ^^-^       P^^-Sft  ft^l 


356 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


05 


^ 


00 

so 

o 

o 

S<l 

o 

o 

GO 

j2 
C 
1— 1 

i 

_ 

^^ 


ft  ^  .ti 


w 


o 

03 

m 

Oi 

o 

O 

-C 

^ 

O 

O 

rt    F^    ^        o    CO 


^ 


03 


O 


o 


05 

o 

en      b 

o 

:?; 

CO 

:^ 

1^ 

1-3    H 

w 

o 

o 

O 

J     Jj 

J] 

Q 

<! 

Q 

Q 

D5 


EsJ 


o   o   o   o   o   o 

o   o   o   o   o   o 


o    o   o   o   o   o 


o  o 


■fi 

c2 

0) 

s 

to 

r> 

>■. 

o 


K       O)  ,iS 


^  -  ^    G 


4^       .X       .-= 


O       .S 


m  H  cc  Q  a  H 


»  «?  -2 


c  i  s  'p 

'3  to  -C     — 

S  O  & 

5  -*^  to 

-2  ti-  -o 


0) 

s 

>, 

-l^ 

>^ 

o 

J- 

o 

c 

li 


o 


M 


2    o 


ANALOG  INHIBITION  OF  THE  METABOLISM  OF  VARIOUS  AMINO  ACIDS    357 


c3 


C     C5 

s  ^ 

o  X- 


rt  -^ 


fe 


I  I 


-§  =3  "S  o  •= 

c3  r-  C3  ^  Gj 

u  C  -If  c  ^ 

^  ci  ^  :::  »3 

^  ^  &  W  c.' 

c  c  ■<  j  J 


05 


S  "^ 


Ph 


Ti 

C3 

C 

7^, 

IS 

hn 

tH 

<u 

S 

^ 

fl 

c 

.^ 

0) 

Z 

C 

o 

O 

O 

2 

o 

s 

O 

2 

LI 

C-1 

'M 

m 
-M 

a  o  o 


Q  o  o  o 


o 


•-  .S    4^ 


Z^        •■-'        -*^  r-       '^  ^7^ 


a;    .=    .=    :^ 


J    C    J    K 


-s  ^  -< 


&q 


c 

c 

d 

rt 

c 

o 

t^ 

»    S 

S 

S    c 

» 

T3 
>> 


H    « 


fe 


1-1    »c 


~    o 
P-     i 


® 

fi") 

c 

71 

rrt 

o 

0; 

0) 

^ 

c 

H 

>> 

.2      ^ 


a     c 


j3 


O     H 


358  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

histidine  is  noncompetitive  indicates  that  the  histidine  site  is  different 
from  the  catalytically  active  site.  ATP  and  5'-P-ribosyl-PP  protect  the 
enzyme  against  inactivation  by  trypsin,  and  histidine  eliminates  this  pro- 
tection, so  it  is  possible  that  histidine  modifies  the  enzyme  configuration 
by  combining  with  the  feedback  site. 

Aminooxyacetate 

An  interesting  analog  studied  recently  in  connection  with  the  physiolog- 
ically important  y-aminobutyrate  (GABA)  is  aminooxyacetate  (+H3N — 
0 — CH2COO").  This  substances  inhibits  the  GABA:  a-ketoglutarate  trans- 
aminase of  E.  coli  very  potently  (40%  at  0.0033  mM  and  100%  at  0.33 
mM  when  both  substrates  are  27  mM),  so  it  was  tested  on  a  similar  enzyme 
from  brain  and  found  to  inhibit  as  strongly  (Wallach,  1960).  Aminooxyace- 
tate in  doses  of  6-50  mg/kg  elevates  the  brain  GABA  in  several  species  as 
much  as  4-  to  5-fold,  peak  levels  being  reached  about  6  hr  after  administra- 
tion and  high  levels  remaining  up  to  24  hr  (Wallach,  1961  a;  Schumann  et  al., 
1962).  Reinvestigation  of  the  transaminase  inhibition  led  to  a  K^  of  0.0075 
mM  (Z,„  is  9.66  mM  for  a-ketoglutarate  and  27.6  mM  for  GABA);  the 
inhibition  is  competitive  with  respect  to  both  substrates.  The  alanine: 
«-ketoglutarate  transaminase  from  rat  liver  and  heart  is  also  potently 
inhibited,  around  60-80%  by  0.0001  mM  aminooxyacetate  at  pH  6.8 
(Hopper  and  Segal,  1964).  Indeed,  this  transaminase  seems  to  be  more 
sensitive  than  either  the  GABA:a-ketoglutarate  or  aspartate:or-ketoglutar- 
ate  transaminase. 

Inasmuch  as  GABA  has  been  implicated  in  certain  convulsive  disorders 
(e.  g.,  GABA  formation  in  epileptic  brains  is  apparently  depressed),  amino- 
oxyacetate was  administered  to  animals  made  convulsive  with  thiosemi- 
carbazide  and  methionine  sulfoximine  (DaVanzo  et  al.,  1961).  Anticonvul- 
sant activity  was  observed  but  there  is  some  doubt  if  this  is  correlated  with 
the  brain  GABA  levels  since  the  time  relations  are  not  correct.  The  effects 
of  aminooxyacetate  on  central  nervous  system  function  are  complex. 
There  is  first  a  progressive  depression  and  muscular  relaxation  with  loss  of 
certain  reflexes,  but  at  high  doses  tonic-clonic  convulsions  occur  (DaVanzo 
et  al.,  1964  a).  Pyridoxal-P  antagonizes  these  convulsions  and  it  was  postu- 
lated that  oxime  formation  occurs  between  aminooxyacetate  and  pyri- 
doxal-P.  Pyridoxal-P  does  not  reverse  the  transaminase  inhibition  in  the 
brain  (Wallach,  1961  b).  W^allach  also  suggested  that  a  depletion  of  suc- 
cinic semialdehyde,  which  arises  from  GABA  by  transamination,  might 
also  play  a  role  in  the  convulsant  action.  Since  a  pyridoxal  deficiency  is  pro- 
duced in  rats  by  the  administration  of  aminooxyacetate,  DaVanzo  et  al. 
(1964  b)  postulated  another  possible  mechanism  of  action,  namely,  the 
inhibition  of  pyridoxal  kinase,  inasmuch  as  McCormick  and  Snell  (1961) 
had  shown  this  enzyme  to  be  rather  potently  inhibited  by  the  condensation 


ANALOG  INHIBITION  OF  THE  METABOLISM  OF  VARIOUS  AMINO  ACIDS    359 

product  of  aminooxyacetate  and  pyridoxal.  It  may  be  that  aminooxyacetate 
should  be  classed  as  a  carbonyl  reagent  which  reacts  with  pyridoxal-P, 
rather  than  strictly  as  an  amino  acid  analog,  but  it  is  sometimes  difficult 
to  distinguish  these  actions.  Nevertheless,  the  effects  of  aminooxyacetate 
on  GABA  levels  in  the  tissues  will  undoubtedly  make  it  a  useful  compound 
for  the  study  of  the  physiological  role  of  GABA. 

Cycloserine 

D-Cycloserine  (orientomycin,  Oxamycin)  is  a  tuberculostatic  antibiotic 
isolated  from  Streptomyces  which  can  be  considered  as  a  cyclic  form  of 
aminooxyalanine  or  0-aminoserine.  It  is  written  in  the  zwitterion  state  be- 

I        I 
H  nh; 

Cycloserine 

cause  there  is  some  evidence  that  this  is  the  inhibitory  form  (Neuhaus  and 
Lynch,  1964).  Its  actions  are  similar  to  aminooxyacetate  in  many  respects 
but  differ  occasionally  in  interesting  ways.  D-Cycloserine  inhibits  the  growth 
of  many  bacteria  and  this  is  often  well  antagonized  by  D-alanine,  and 
competitively  inhibits  the  incorporation  of  D-alanine  into  a  uridine  nucleo- 
tide necessary  for  the  synthesis  of  cell  wall  material.  The  growth  of  myco- 
bacteria is  50%  suppressed,  for  example,  by  D-cycloserine  at  0.03-0.045  mM 
and  this  is  reversed  by  D-alanine  but  not  by  L-alanine  (Zygmunt,  1963). 
L-Cycloserine  is  also  inhibitory  to  certain  bacteria  and  this  is  antagonized 
by  L-alanine.  It  is  interesting  that  *S.  aureus  can  develop  a  50-fold  resistance 
to  D-cycloserine,  no  cross-resistance  with  other  antibiotics  being  observed 
(Howe  et  al.,  1964).  In  animals  it  produces  sedation,  lethargy,  muscular 
relaxation,  ataxia,  an  accentuated  startle  response,  and,  above  all,  epi- 
leptic convulsions  (Dann  and  Carter,  1964;  Holtz  and  Palm,  1964).  In  these 
respects  it  at  least  superficially  acts  like  aminooxyacetate  and,  furthermore, 
these  effects  are  antagonized  by  pyridoxal. 

D-Cycloserine  inhibits  certain  enzymes  dependent  on  pyridoxal-P,  such 
as  the  transaminases  and  glutamate  decarboxylase,  and  it  has  been  postu- 
lated that  it  simply  reacts  with  pyridoxal-P  to  form  a  substituted  oxime. 
However,  this  is  not  the  usual  reaction  in  which  a  Schiff  base  is  produced, 
but  involves  an  opening  of  the  cycloserine  ring.  The  inhibition  of  GABA: 
or-ketoglutarate  transaminase  is  initially  competitive  with  respect  to  GABA 
{K^  is  0.25  TCiM  for  the  enzyme  from  cat  brain  and  around  0.6  mM  for  the 
enzyme  from  E.  coli),  but  a  secondary  progressive  irreversible  inhibition 
also  occurs  (Dann  and  Carter,  1964).  This  may  be  related  to  the  hypothesis 
of  Khomutov  et  al.  (1961)  that  the  decyclicized  cycloserine  forms  an  oxime 


360  2.  ANALOGS  OF  ENZYME  KE ACTION  COMPONENTS 

bond  with  enzyme-bound  pyridoxal-P,  and  also  an  acyl  bond  with  a  cat- 
ionic  group  at  the  active  site,  these  two  groups  thus  being  connected  by  a 
bridge  would  prevent  access  to  the  site.  The  L-asparagine:a-ketoglutarate 

I 
HC = NH— 0— CHa— CH— CO 

I  I 

Pyr  X 


Apoenzyme 

transaminase  is  inhibited  by  L-cycloserine  with  respect  to  L-asparagine 
{K^  =  0.0001  mM)  but  D-cycloserine  is  a  much  weaker  inhibitor  {K^  = 
1  mM)  (Braunstein  et  al.,  1962).  L-Cycloserine  likewise  competitively 
inhibits  the  L-alanine:  a-ketoglutarate  transaminase  {K^  =  0.008  mM). 
This  taken  with  previous  work  indicates  that  the  cycloserines  inhibit 
specifically  those  enzyme  attacking  substrates  of  the  same  optical  isomerism. 
Another  possible  site  of  action  of  D-cycloserine  in  bacteria  would  be  the 
D-alanyl-D-alanine  synthetase,  a  block  of  which  would  prevent  the  incor- 
poration of  D-alanine  into  cell  wall  material.  Alanylalanine  is  often  able  to 
reverse  the  cycloserine  inhibition  of  bacterial  growth,  sometimes  more  ef- 
fectively than  alanine.  One  example  of  this  is  the  inhibition  of  the  prolif- 
eration of  agents  of  the  psittacosis  group  in  chick  embryo  yolk  sac  (Moulder 
et  al.,  1963).  Chick  embryos  infected  with  the  mouse  pneumonitis  organism, 
for  example,  are  well  protected  by  D-cycloserine  at  0.004-0.008  mM,  and,  of 
all  the  possible  reversors  tested,  only  alanylalanine  is  effective.  The  d- 
alanyl-D-alanine  synthetase  of  Streptococcus  fecalis  is  inhibited  competi- 
tively with  respect  to  D-alanine  {K^  =  0.022  mM),  and  Neuhaus  and 
Lynck  (1964)  felt  that  this  enzyme  may  well  be  the  major  site  of  inhibition 
in  certain  bacteria.  It  is  unfortunate  that  cycloserine  and  aminooxyacetate 
have  not  been  accurately  compared  in  any  study. 


DIAMINE   OXIDASE   (HISTAMINASE) 

Enzymes  in  this  group  exhibit  different  degrees  of  substrate  specificity 
depending  on  the  source,  but  most  oxidatively  deaminate  diamines  of  the 
type  +H3N — (CH2),i — NH3+  (with  maximal  rates  when  n  is  around  5)  and 
histamine,  in  all  cases  primary  amines  being  attacked.  The  usual  substrate 
in  most  studies  has  been  cadaverine  {n  =  5). 

The  diamines  have  been  written  as  cations  because  the  p^^'s  are  usually 
between  8.5  and  10.5.  The  amidines  and  guanidines  exist  almost  entirely 
as  cations  at  physiological  pH  since  the  p^^'s  for  these  groups  are  around 
13  to  14.  The  structures  have  been  written  in  a  rather  unconventional  way 


DIAMINE  OXIDASE  (hISTAMINASE) 


361 


H3N— (CH2)— NHa"" 
Cadaverine 


Putrescine 


H3N— (CH2  )3— ^fH— (C  H2  )4— NH — (CH2  )3— NH^ 


*2M 

Spermine 


CH3— (CH3)^C^  + 
Monoamidines 


HjN^        ,        ,  /NH2 

HzN'^  NH2 

Diamidines 


/=\  /NH3 

Dibenzamidines 


H,N— C 

^NH— CH3 

Methylguanidine 


H,N-C 


'NH^ 


-NH— NH2 
Aminoguanidine 


H^N^  /NH2 

+    C— NH— (CH,);,— NH— C     + 

Diguanidines 


+  /NH2 

H3N— (CHs).— NH-C     + 

Agmatine 


H^N^  /NH2 

+    C— S  — (CHs)^— S-C-    + 
HgN^  NH2 

Diisothioureas 


H2N  /NH2 

+    C-NH  — (CH2)4— NH-C     + 
HgN^  NH2 

Arcaine 


to  indicate  the  equivalence  of  the  C — N  bonds  and  the  states  of  the  amino 
groups,  since  there  is  resonance  between  forms  such  as  the  following  for 
the  guanidinium  ion: 


+       /NH2 


H,N— C 


+ 
NH, 


^NH, 


H,N-C, 


,NH, 


NH, 

+    ^ 


Alkyl  substitution  does  not  reduce  the  resonance  appreciably. 

Aliphatic  monoamines,  such  as  amylamine,  are  not  substrates  nor  are 


362 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


they  readily  bound  to  the  enzyme  since  they  inhibit  weakly  (Zeller,  1940). 
The  short  aliphatic  diamines  are  very  poor  substrates  but  inhibit  quite 
well;  thus  both  ethylenediamine  and  trimethylenediamine  inhibit  the 
oxidation  of  cadaverine  by  pig  kidney  diamine  oxidase  (Zeller,  1938). 
If  the  amino  groups  of  these  inhibitors  react  with  the  same  anionic  enzyme 
sites  as  does  cadaverine,  these  anionic  groups  must  be  fairly  close  so  that 
cadaverine  would  have  to  assume  a  very  bowed  configuration.  On  the  other 

NH3+ 
\  ^ 
CH, 
\ 

CH, 

/     ' 

NH+ 


H. 

C 

HgC 

CH2 

/ 

\ 

HgC 

CH2 

+  \ 

/+ 

H3N 

NH3'^ 

- 

- 

H,C- 
^H,N 


-CH, 

NH, 


(a) 


(b) 


(c) 


hand,  the  anionic  groups  may  be  separated  by  a  distance  corresponding 
to  the  amino  groups  in  cadaverine  and  the  inhibitors  react  with  only  one 
of  the  anionic  sites  ,the  other  amino  group  interacting  with  some  anionic 
site  outside  the  active  center  (as  in  (c)). 

Certain  guanidine  derivatives  are  more  potent  inhibitors.  Guanidine 
itself  is  rather  weak,  inhibiting  cadaverine  (2  mM)  oxidation  42%  at  10 
mM,  but  methylguanidine  inhibits  63%  at  1  mM  (Zeller,  1938).  Although 
Zeller  noted  that  in  the  pig  kidney  preparation  methylguanidine  did  not 
inhibit  histamine  oxidation  very  well,  Waton  (1956)  found  marked  inhibi- 
tion of  histaminase  activity  in  cat  kidney,  0.01  mM  inhibiting  42%  and 
0.1  mM  75%.  The  oxidations  of  putrescine  and  agmatine  are  both  well 
inhibited  by  methylguanidine  (Zeller,  1940).  An  even  more  potent  inhibitor, 
however,  is  aminoguanidine.  Apparently  the  diamine  oxidases  differ  in 
sensitivity  to  aminoguanidine;  50%  inhibition  is  given  by  0.00005  mM  for 
the  enzyme  from  pig  kidney  (Schuler,  1952),  by  0.001  mM  for  the  enzyme 
from  cat  kidney  (Waton,  1956),  by  0.01  mM  for  the  enzyme  from  rabbit 
liver  (Kobayashi,  1957),  and  the  enzyme  from  mouse  liver  is  not  inhibited 
even  by  0.1  mM.  The  nature  of  the  inhibition  is  not  clear,  inasmuch  as 
aminoguanidine  is  also  a  derivative  of  hydrazine  and  might  act  by  attacking 
carbonyl  groups:  hydrazine  and  semicarbazide  are,  indeed,  potent  inhibitors 
of  diamine  oxidase  (Schuler,  1952;  Waton,  1956).  A  further  complication  is 
that  aminoguanidine  hydrolyzes  to  form  semicarbazide  and  eventually 
hydrazine.  It  behaves  chemically  more  like  a  hydrazine  than  a  guanidine, 
and  reacts  with  carbonyl  groups  without  being  hydrolyzed  (Lieber  and 
Smith,  1939).  It  is  also  possible  that  the  NHNHg  group  simulates  the 
CH2NH2  substrate  group,  as  in  the  monoamine  oxidase  inhibitors,  and  forms 
a  tight  bond  to  the  enzyme.  In  vivo  inhibition  of  diamine  oxidase  by  amino- 


DIAMINE  OXIDASE   (HISTAMINASE) 


363 


guanidine  was  demonstrated  by  Schayer  et  al.,  (1954)  by  injecting  labeled 
cadaverine  into  mice  and  determining  the  C^^Og  respired  (Fig.  2-7).  It  is 
much  more  effective  than  isoniazid  or  agmatine.  The  metabolism  of  hista- 
mine should  also  be  blocked  by  aminoguanidine.  This  was  shown  in  three 
psychiatric  patients  by  administering  labeled  histamine  and  finding  that 


100 


-1.5  -0.5 

LOG    INH    DOSE   (^g/G) 


+  0.5 


+  1.5 


+  2.5 


+  3.5 


Fig.  2-7.  Effects   of  diamine   oxidase   inhibitors  on  the  oxidation  of  cadaverine  in 
mice,  as  determined  by  the  formation  of  C'^Oj  from  labeled  cadaverine.  (Data  from 

Schayer  et  al.,  1954.) 


urinary  imidazoleacetate-C^^  is  reduced  by  aminoguanidine  at  0.1-1  mg/kg 
(Lindell  et  al.,  1960).  A  larger  fraction  of  the  histamine  is  excreted  as 
methylhistamine,  demonstrating  a  diversion  of  metabolic  pathways  by 
this  inhibitor. 

Diamine  oxidase  is  not  strongly  inhibited  by  monoamidines,  but  diami- 
dines,  diguanidines,  and  diisothioureas  of  the  proper  chain  lengths  are 
quite  potent  inhibitors  (Blaschko  et  al.,  1951).  The  data  for  these  series 
are  summarized  in  Fig.  2-8.  The  correlation  between  chain  length  and 
inhibition  is  not  nearly  so  clear  as  for  monoamine  oxidase  (Blaschko  and 
Duthie,  1945;  Blaschko  and  Himms,  1955),  and  in  some  cases,  as  the  dia- 
midines,  there  is  surprisingly  little  variation  of  inhibition  with  chain  length. 
For  diamidines  of  n  =  10-16,  one  wonders  if  the  binding  is  actually  to 


364 


2.  ANALOGS  OF  ENZYME  KEACTION  COMPONENTS 


Fig.  2-8.   Inhibition   of  pig   kidney   diamine  oxidase   with   cadav- 
erine  (5  niM)  as  the  substrate  and  all  the  inhibitors  at   1   mM. 
(From  Blaschko  et  ah,  1951.) 
NH, 


(A)    Monamidines 


H3C  —  (CHa)^     C 


NH 


(B)    Diamidines 


H,N 


NH, 


(C)    Dibenzamidines 


(D)    Diguanidines 


H2N 
HN 
H,N 


0-(CH2)„-0— ( 


NH, 


NH, 


W 
NH 


C-NH-(CH2)— HN-C^ 
HN  ^NH 


(E)   Diisothioureas 


H,N 


NH, 


C-S-   CH,L— S— C 

//  \\ 

HN  NH 


CARBOXYPEPTIDASE,    AMINOPEPTIDASES,    DIPEPTIDASES  365 

the  substrate  site  entirely,  or  possibly  to  two  substrate  sites,  or  even  to 
anionic  groups  outside  the  substrate  site.  The  enzyme  may  have  binding 
sites  for  the  imidazole  ring  of  histamine  since  imidazole  inhibits  11%, 
imidazolelactate  20%,  and  histidine  4%  at  6.7  mM  when  cadaverine  is 
3.3  mM  (Zeller,  1941).  Urate  also  competitively  inhibits  the  oxidation  of 
histamine,  but  rather  weakly.  An  excellent  review  of  the  structure-action 
relationships  among  the  amidine  derivatives  is  by  Fastier  (1962). 

CARBOXYPEPTIDASE,   AMINOPEPTIDASES, 
AND    DIPEPTIDASES 

Certain  aspects  of  the  inhibition  of  carboxypeptidase  by  substrate 
analogs  were  discussed  in  Volume  I  (page  292)  to  illustrate  how  certain 
interaction  contributions  could  be  estimated.  We  shall  now  attempt  to 
visualize  more  clearly  the  orientation  of  these  analogs  on  the  enzyme  sur- 
face. The  data  indicate  that  a  three-point  attachment  of  the  substrate 
is  necessary  for  catalysis.  The  enzyme  sites  may  be  indicated  as  follows 
(see  Fig.  2-9  for  hypothetical  orientation  of  substrate):  (A)  the  peptidatic 
site  contains  the  mechanism  of  the  electron  displacement  necessary  for  hy- 
drolysis and  is  probably  positively  charged,  (B)  the  cationic  site  is  a  positively 
charged  group  that  interacts  electrostatically  with  the  C00~  group,  and 
(C)  the  electrokinetic  site  is  perhaps  a  lipophilic  region  capable  of  reacting 
with  alkyl  or  phenyl  groups  by  dispersion  forces.  It  is  easy  to  see  why 
D-substrates  are  not  reacted  since  the  peptide  bonds  would  not  be  able  to 
approach  the  peptidatic  site.  There  is  also  an  enzyme  region  near  the  projec- 
tion direction  of  the  fourth  asymmetric  carbon  bond  that  sterically  prevents 
attachment  of  groups  larger  than  an  amino  group,  and  thus  the  D-isomers 
usually  do  not  bind  and  are  not  inhibitors.  Only  two-point  attachment  is 
necessary  for  inhibitors,  and  most  that  have  been  studied  bind  at  the  cationic 
and  electrokinetic  sites. 

The  relative  binding  energies  for  inhibitors  in  Table  1-6-26  were  calculated 
from  the  data  of  Smith  etal.  (1951).  Earlier  studies  by  Elkins-Kaufman  and 
Neurath  (1949)  provide  additional  information  on  the  competitive  inhibitors 
in  the  accompanying  tabulation.  It  is  interesting  that  D-phenylalanine  is 


Inhibitor 

(mM) 

Relative  —  AF  oi  binding 
(kcal/mole) 

^-Phenylpropionate 

0.062 

5.96 

Phenylacetate 

0.39 

4.83 

y  -  Phenylbutyrate 

1.13 

4.17 

D  -  Phenylalanine 

2 

3.82 

2?-Nitrophenylacetate 

2.5 

3.68 

366 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

A. 


Acyl-L-phenyl- 
alanine  (substrate) 


(3"'-^l 


L- Phenylalanine 


/^ X  NH3  Vc 


r    hV 


Benzoate 


\^       y  CHj 


D- Phenylalanine 


Ql 


Phenylacetate 


Propionate 


Q- 


3  -  Pheny  Ipr  opionat  e 
(hydrocinnamate) 


Valerate 


y- Pheny  Ibuty  rate  3-Indoleacetate 

Fig.  2-9.  Possible  orientations  of  substrate  and  inhibitors  at  the  active   center   of 
carboxypeptidase.  (A)  is  peptidatic  site,  (B)  is  cationic  site,  and  (C)  is  electrokinetic 
site.  The  molecular  configurations  are  only  approximate. 


CARBOXYPEPTIDASE,    AMINOPEPTIDASES,    DIPEPTIDASES  367 

a  better  inhibitor  than  L-phenylalanine  under  the  usual  conditions,  since 
the  NH3+  group  of  the  latter  is  repelled  by  the  positively  charged  pep- 
tidatic  site  while  in  the  former  it  does  not  encounter  serious  steric  in- 
terference (Fig.  2-9).  However,  /5-phenylpropionate  (hydrocinnamate)  is 
bound  more  tightly  than  D-phenylalanine  by  2.14  kcal/mole,  suggesting 
that  some  steric  repulsion  of  the  latter  analog  occurs.  iV- Substitution  in- 
creases the  repulsion  markedly  and  inhibitory  activity  is  lost.  A  comparison 
of  the  possible  orientations  of  some  inhibitors  in  Fig.  2-9  with  the  relative 
binding  energies  may  give  some  idea  of  the  structural  requirements  for 
potent  inhibition.  It  may  be  noted  that  benzylmalonate  is  an  effective 
inhibitor  but  is  somewhat  less  well  bound  than  /5-phenylpropionate;  this  is 
surprising  because  it  might  be  thought  that  additional  energy  would  be 
contributed  by  interaction  of  one  of  the  C00~  groups  with  the  peptidatic 
site.  Neither  cis-  nor  <rans-cinnamate  inhibits  and  it  was  suggested  that  the 
double  bond  restricts  the  orientation  of  the  ring  so  that  adequate  binding 
cannot  occur.  The  linearity  of  these  molecules  may  also  be  a  factor,  since 
the  active  site  is  probably  not  flat  as  is  implied  by  the  two-dimensional 
representations  in  the  figure. 

Competitive  inhibition  by  the  following  analogs  has  been  demonstrated 
more  recently:  3-indolepropionate,  e-aminocaproate,  (5-amino-w-valerate 
(Folk,  1956;  Greenbaum  and  Sherman,  1962),  y-aminobutyrate,  S-guani- 
dinovalerate,  argininate  (Folk  and  Gladner,  1958),  iV-acetyl-L-tyrosine, 
D-leucinyl-L-tyrosine,  glycyl-L-tyrosine,  and  other  dipeptides  (Yanari  and 
Mitz,  1957).  The  inhibition  apparently  sometimes  depends  on  the  substrate 
used;  for  example,  3-indolepropionate  inhibits  the  hydrolysis  of  carbo- 
benzoxyglycyl-L-phenylalanine  but  not  the  hydrolysis  of  a-iV-benzoyl- 
glycyl-L-lysine,  where  s-aminocaproate  exhibits  just  the  opposite  behavior. 

Relatively  little  work  has  been  done  on  analog  inhibition  of  dipepti- 
dases  and  aminopeptidases,  and  it  will  suffice  to  mention  a  few  isolated 
observations.  Yeast  dipeptidase  is  inhibited  by  various  amino  acids;  for 


Concentration  for 

Amino  acid 

50%    inhibition 

(mM) 

L-Leucine 

1.5 

L-Isoleucine 

1.8 

L-Tryptophan 

6.0 

L-Histidine 

8.0 

L-Leucinamide 

9.0 

L-Arginine 

18 

DL- Valine 

22 

DL-Phenylalanine 

22 

DL-Serine 

>100 

368  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

example,  L-leucine,  D-alanine,  and  glycine  inhibit  the  hydrolysis  of  alanyl- 
glycine  and  glycylglycine  (Grassmann  et  al.,  1935).  This  has  been  investi- 
gated more  quantitatively  by  Nishi  (1960)  and  his  results  are  summarized 
in  the  accompanying  tabulation  (substrate  is  glycylglycine  at  50  mM). 
These  inhibitions  are  competitive  with  respect  to  substrate  and  uncompe- 
titive with  respect  to  Co++.  Some  interesting  inhibitions  of  pig  kidney 
leucine  aminopeptidase  have  been  reported  by  Hill  and  Smith  (1957). 
The  hydrolysis  of  substrates  of  the  type  R— CH(NH3+)— CONH— R' 
depends  on  a  three-point  attachment  of  the  R  group,  the  NH3+  group, 
and  the  amide  N.  The  R  groups  interact  by  van  der  Waals'  forces;  further 
energy  is  contributed  from  the  hydrogen  bonding  of  water  molecules  dis- 
placed from  the  hydrophobic  surfaces.  The  inhibitions  given  in  the  follow- 
ing tabulation  are  for  L-leucinamide  as  substrate  at  50  mM  and  at  pH  8.50- 


Inhibitor 

Concentration 
{mM) 

%  Inhibition 

L-Leucine 

50 

45 

100 

61 

L-Leucinol 

50 

38 

100 

48 

Isocaproamide 

25 

41 

50 

56 

Isocaproate 

100 

47 

200 

65 

w-Hexylamine 

100 

0 

a-Ketoisocaproamide 

25 

21 

L-a-Hydroxyisocaproamide 

50 

68 

8.65.  Every  good  inhibitor  contains  an  R  group  that  should  give  nearly 
optimal  interaction  with  the  electrokinetic  site  of  the  enzyme;  in  addition, 
at  least  one  group  that  will  react  with  the  bound  Mn++  is  present. 


CHYMOTRYPSIN  AND  OTHER  PROTEOLYTIC  ENZYMES 

Chymotrypsin  hydrolyzes  various  amides  and  esters  of  the  general  type: 

NH— R2 
Rj —  CH2     CH 

bo— R3 

where  R^  represents  the  side  chains  of  amino  acids  (phenylalanine,  tyrosine, 
and  tryptophan  most  commonly  used),  Rg  is  an  acyl  group  (acetyl,  benzoyl, 


CHYMOTRYPSIN    AND    OTHER   PROTEOLYTIC    ENZYMES  369 

or  nicotinyl),  and  Eg  is  a  group  forming  either  an  amide  or  ester  bond. 
A  typical  substrate  is  benzoyl-L-tyrosinamide: 


NH— CO 
KO—(.        />— CH,— CH 

\\     //  Vo-NH 


o 


In  addition,  the  NH — R2  chain  may  be  replaced  by  H,  CI,  or  OH  groups. 
Only  the  derivatives  of  L-amino  acids  are  hydrolyzed.  The  R^  and  Rg 
groups  are  important  in  binding  to  the  enzyme  and  thus,  with  the  esteratic 
(paptidatic)  site,  one  may  again  visualize  a  three-point  attachment.  Analogs 
either  devoid  of  susceptible  amide  or  ester  bonds,  or  having  in  their 
place  bonds  resistent  to  hydrolysis,  are  often  inhibitory.  The  R^  group  is  the 
most  important  for  binding,  as  is  shown  by  the  strong  inhibitory  activity 
of  /5-phenylpropionate  (hydrocinnamate)  (Kaufman  and  Neurath,  1949). 
The  necessity  for  at  least  one  aromatic  ring  in  one  of  the  side  chains  was 
pointed  out  by  Neurath  and  Gladner  (1951).  Their  data  on  the  /^-substituted 
propionates  indicate  the  ring  groups  to  have  the  following  order  of  inhibitory 
activity: 

Indole  >  napthyl  >  phenyl  >  2,4-dinitrophenyl  >  cyclohexyl 

The  distance  between  the  COO"  group  and  the  Rj  group  is  also  of  impor- 
tance. The  inhibitions  are  summarized  in  Table  2-19.  The  weaker  binding 
of  cyclohexyl  derivatives  compared  to  phenyl  compounds  (0.5-1  kcal/mole) 
could  be  explained  by  either  the  smaller  polarizability  of  the  cyclohexyl 
ring  or  the  inability  of  the  cyclohexyl  ring  to  approach  the  enzyme  surface 
as  close  as  the  phenyl  ring.  The  strong  binding  of  the  indole  compounds 
was  explained  on  the  basis  of  the  enhancement  of  hydrogen  bonding  by  the 
ring  N.  In  fact,  Neurath  and  Gladner  interpreted  the  inhibitions  by  most 
of  the  analogs  in  terms  of  hydrogen  bonding.  Even  the  C00~  may  not 
interact  electrostatically  with  an  enzyme  cationic  group  since  the  binding 
energies  are  quite  low;  indeed,  it  may  serve  as  a  hydrogen  acceptor.  The 
equivalent  bindings  of 


.        /y — CH2CH2— COO  and  (^        /)— O-CH^CH,— OH 

l3-Phenylpropionate  2-Phenoxyethanol 

would  be  difficult  to  explain  otherwise;  however,  the  latter  compound  can 
act  as  a  hydrogen  donor  in  forming  a  hydrogen  bond.  These  two  com- 
pounds have  essentially  the  same  molecular  dimensions  but  the  electronic 
configurations  of  the  terminal  groups  differ. 


370  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Table  2-19 

Competitive   Inhibition   of   a-CnYMOXRYPSiN   by   Analogs  " 


Inhibitor 


Apparent  iC,  Relative  —  AF  oi  binding 

(mM)  (kcal/mole) 


3-Indolepropionate  2.5  3.57 

3-Indolebutyrate  3.6  3.35 

a-Naphthylpropionate  4.0  3.28 

2,4-Dinitro-^-phenylpropionate  5.3  3.08 

^-Phenylpropionate  5.5  3.07 

2-Phenoxyethanol  5.8  3.05 

y-Phenylbutyrate  14  2.53 

Cyclohexylpropionate  30  2.08 

Cyclohexylbutyrate  35  1.99 

Benzoate  42  1.88 

Phenylacetate  42  1 .  88 

a-Naphthylmethylmalonate  55  1 .  72 

a-Benzylmalondiamide  78.8  1.51 

Cyclohexylacetate  86  1.46 

"  The  substrate  is  acetyl-L-tyrosinamide  (K^  =  27  mM).  Experiments  at  pH  7.8 
and  25°.  (Data  from  Neurath  and  Gladner,  1951.) 

For  the  past  several  years  Niemann  and  his  associates  have  conducted 
excellent  quantitative  investigations  of  the  inhibition  of  chymotrypsin 
by  a  variety  of  analogs  and  their  results,  some  of  which  are  presented  in 
Table  2-20,  provide  a  basis  for  the  interpretation  of  the  binding  mechanisms. 
Although  final  conclusions  must  await  completion  of  their  work,  some  tenta- 
tive ideas  may  be  expressed. 

(A)  Despite  the  fact  that  the  derivatives  of  L-amino  acids  are  hydro- 
lyzed  by  chymotrypsin,  the  D-isomers  of  several  inhibitors  are  bound  on  an 
average  of  0.45  kcal/mole  more  tightly  than  the  corresponding  L-isomers 
(see  also  page  271).  Although  three-point  attachment  may  be  important  for 
substrate  binding,  it  is  evident  from  this  difference  and  other  data  that 
it  is  not  for  inhibitor  binding. 

{B)  Comparing  the  derivatives  of  the  three  amino  acids,  it  is  seen  that 
the  phenylalanine  and  tyrosine  analogs  are  equally  bound,  whereas  the 
tryptophan  analogs  are  bound  some  1.19  kcal/mole  more  tightly,  and  this  is 
probably  to  be  attributed  to  the  greater  affinity  of  the  enzyme  for  the 


CHYMOTRYPSIN   AND    OTHER   PROTEOLYTIC    ENZYMES 


371 


Table  2-20 
Competitive  Inhibition  of  q-Chymoteypsin  by  Analogs  " 


Inhibitor 


Apparent  K^       Relative  —  AF  oi  binding 
(mil/)  (kcal/mole) 


Tryptophan  series 
L-Tryptophanamide 
D-Tryptophanamide 
Acetyl-D-tryptophanamide 
Trifluoroacetyl-D-tryptophanamide 
Acetyl-L-tryptophanmethylamide 
Acetyl-D-tryptophanmethylamide 
Benzoyl-D-tryptophanamide 
Nicotinyl-D-tryptophanamide 
p-Methoxybenzoyl-D-tryptophanamide 
Acetyl-L  -  tryptophanate 
Acetyl-D  -tryptophanate 
Acetyl-D-tryptophan  isopropylester 
Tryptamine 
Acetyltryptamine 
Trifluoroacetyltryptamine 
Indole 

Indoleacetate 
Indolepropionate 
Indolebutyrate 
Indolepropionaraide 

Tyrosine  series 

Acetyl-D-tyrosinamide 

Trifluoroacetyl-D-tyrosinamide 

Chloroacetyl-D-tyrosinamide 

Nicotinyl-D-tyrosinamide 

Acetyl-L-tyrosinemethylamide 

Formyl-L-tyrosinemethylamide 

Nicotinyl-L-tyrosinemethylamide 

Benzoyl-L-tyrosinemethylamide 

Acetyl-L-t>Tosinate 

Fluoroacetyl-L-tyrosinate 

Chloroacetyl-L-tyrosinate 

Acetyl-D-tyrosine  ethyl  ester 

Nicotinyl-D-t>Tosine  ethylester 

D-Tyrosinehydroxamide 

Acetyl-D-tyrosinehydroxamide 


8.5 

2.93 

4.0 

3.40 

2.4 

3.72 

4.0 

3.40 

6.5 

3.10 

1.8 

3.89 

0.7 

4.46 

1.6 

3.96 

0.6 

4.55 

9.5 

2.87 

7.5 

3.01 

0.8 

4.39 

2.3 

3.74 

1.8 

3.89 

1.2 

4.13 

0.8 

4.34 

18 

2.47 

15 

2.58 

23 

2.32 

2.3 

3.74 

12 

2.72 

20 

2.40 

6.5 

3.10 

9 

2.90 

61 

1.72 

31 

2.14 

8.8 

2.92 

6.4 

3.12 

110 

1.36 

120 

1.30 

150 

1.16 

4.7 

3.30 

0.8 

4.39 

40 

1.98 

7.5 

3.01 

372  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Table  2-20  {continued) 


Inhibitor 


Apparent  K^      Relative  —  Zli^  of  binding 
(mM)  (kcal/mole) 


Phenylalanine  series 

Acetyl-D-phenylalaninamide  12  2 .  72 

Nicotinyl-D-phenylalaninamide  9  2.90 

Acetyl-D-phenylalanine  methylester  2.3  3.74 

Benzoate  150  1.16 

Benzamide  10  2 .  83 

Phenylacetate  200  0.99 

Phenylacetamide  15  2.58 

Phenylpropionate  25  2 .  26 

Phenylpropionamide  7.0  3 .  05 

Phenylbutyrate  60  1.73 

Phenylbutyramide  12  2.72 

"  The  values  of  K^  are  taken  from  work  at  pH  7.9.  A",  varies  with  the  pH  and  thus 
the  degree  of  ionization  may  be  of  importance  in  some  instances.  The  relative  binding 
energies  are  therefore  subject  to  some  error  but  may  provide  an  initial  basis  for  discus- 
sion. Various  substrates  have  been  used  in  the  different  studies  and  this  may  introduce 
further  uncertainties.  (Data  from  Foster  et  al.,  1955;  Foster  and  Niemann,  1955  a, 
b;  Lands  and  Niemann,   1959.) 

indole  ring  system.  This  is  confirmed  by  comparing  the  phenyl  and  indole 
acids  and  amides,  the  indole  derivatives  being  bound  0.77  kcal/mole  more 
strongly  on  the  average. 

(C)  Comparing  the  iV-substituted  Rg  groups  one  may  calculate  that  the 
order  of  binding  is: 

A  =  0.35  A  =  0.68 

(kcal/mole)  (kcal/mole) 

Benzoyl  >         nicotinyl  >  acetyl 

and  it  appears  that  the  acetyl  derivatives  are  better  inhibitors  than  the 
analogs  with  a  free  NH3+  group.  It  is  likely  that  these  groups  interact 
with  the  enzyme  surface  in  a  nonspecific  fashion. 

(D)  Analogs  with  a  CONH2  group  are  bound  around  1.10  kcal/mole 
more  tightly  than  those  with  a  free  C00~  group.  This  might  indicate  that 
the  peptidatic  site  is  in  an  electric  field  arising  from  surrounding  negative 
charges,  but  it  could  also  mean  that  hydrogen  bonds  between  the  pep- 
tide linkage  and  the  enzyme  are  important.  The  rather  tight  binding  of 


CHYMOTRYPSIN   AND    OTHER    PROTEOLYTIC    ENZYMES  373 

tryptaraine  indicates  also  a  more  favorable  field  for  positive  than  negative 
ionic  groups,  but  the  positive  charge  is  not  important  since  acetyltrypt- 
amine  is  bound  even  more  readily. 

Foster  and  Niemann  (1955  a)  determined  the  values  of  K^  for  several 
inhibitors  at  pH  6.9  and  7.9.  The  acetyltryptophanates  and  indolepropio- 
nate  are  bound  1.0-1.4  kcal/mole  more  tightly  at  pH  6.9  than  at  pH  7.9, 
whereas  the  binding  of  various  amides  is  not  significantly  affected.  This  is 
interpreted  as  pointing  to  the  development  of  a  negative  charge  in  the  vicin- 
ity of  the  active  site  as  the  pH  rises  from  6.9  to  7.9;  this  charge  would  repel 
the  negatively  charged  tryptophanates  and  other  carboxylates.  It  is  pos- 
sible that  the  lack  of  effect  of  pH  on  amide  binding  is  due  to  the  simulta- 
neous deprotonation  of  the  CONH3+  with  rise  in  pH  (p^^  =  7.5).  The  con- 
cept of  a  negative  charge  on  or  near  the  active  center  was  first  postulated 
by  Neurath  and  Schwert  (1960)  on  the  basis  of  the  suppressing  effect  of  a 
carboxylate  group  on  the  hydrolysis  of  an  adjacent  peptide  bond.  Whether 
this  enzyme  negative  group  participates  in  the  hydrolysis  along  the  lines 
suggested  by  Steam  (1935)  and  Vaslow  and  Doherty  (1953)  is  not  certain. 

The  nitrogen  analogs  of  substrates  are  usually  not  hydrolyzed  and  can 
act  as  inhibitors  (Kurtz  and  Niemann,  1961).  In  these  compounds  an  a- 
methine  group  is  replaced  by  a  N  atom,  or  an  or-methylene  group  by  an 
NH  group.  Thus  ethyl  l-acetyl-2-benzylcarbazate  is  an  analog  of  ethyl 

^COO— Et 

CH2— N 

NH-  COCH3 

acetyl-L-phenylalaninate  and  is  an  inhibitor,  with  K^  =  20  raM.  Also 
9)— CH2CH2— COO— CH3  and  (7^— NH— CH2— COO— CH3  are  substrates  of 
chymotrypsin  whereas  cp — CHg — NH — COO — C2H5  is  an  inhibitor  [K^  = 
6  m.M).  It  is  suggested  that  an  a-N=  or  a-NH —  group  leads  to  a  restriction 
of  the  rotation  around  the  bond  joining  it  to  the  CO  group  compared  to 
that  of  a  C — C  bond,  this  constraint  leading  to  loss  of  substrate  activity  or 
weakening  of  inhibitor  binding. 

Wallace  et  al.  (1963)  published  the  results  of  their  studies  of  the  interac- 
tions of  136  compounds  with  chymotrypsin,  and  it  is  seen  that  several 
substances  not  directly  related  to  amino  acids  are  fairly  potent  inhibitors; 
such  are  quinoline  and  the  methylquinolines,  hydroxyquinolines,  and  ami- 
noquinolines,  various  acridines,  a-naphthol,  and  the  naphthylamines,  all 
with  K^  values  less  than  1  mM .  They  summarized  their  conclusions  with  re- 
spect to  structure  and  inhibitory  activity  in  ten  postulates,  from  which  the 
following  comments  are  extracted.  Aromatic  compounds  are  more  effective 
than  the  corresponding  saturated  derivatives,  and  monosubstituted  benzene 
derivatives  with  polarizable  uncharged  groups  are  more  inhibitory  than  the 


374  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

parent  compounds;  both  facts  indicate  the  importance  of  polarization  of 
the  inhibitor  molecule  in  the  field  of  ionic  groups  on  the  enzyme.  The  ac- 
tive site  has  a  locus  for  interaction  with  the  aromatic  nucleus,  and  vicinal 
to  this  there  is  at  least  one  anionic  group;  the  orientation  of  the  inhibitor 
is  determined  by  the  interaction  of  the  polarizable  group  with  a  sublocus. 
Molecules  presenting  a  larger  planar  area  are  more  inhibitory,  and  it  seems 
that  the  site  at  which  the  aromatic  compounds  act  is  mainly  flat,  of  greater 
length  than  breadth,  and  not  straight  but  curved  along  the  enzyme  surface. 
The  different  loci  involved  in  the  interactions  perhaps  have  different  prop- 
erties; e.  g.,  it  was  postulated  that  the  active  site  may  be  hermaphroditic* 
in  that  one  locus  may  be  electron-deficient  and  another  electron-rich. 

The  chymotrypsin  reaction  proceeds  in  two  steps,  the  acetylation  of  the 
enzyme: 

EH  +  AcB  z^  (EH— AcB)  ^  (EAc-HB)  ^  EAc  +  HB 

and  the  solvolysis  of  the  acetylchymotrypsin: 

EAc  +  HOR  ^  (EAc-HOR)  ^  (EH— AcOR)  5=t  EH  +  AcOR 

The  competitive  inhibitors  tabulated  here  were  shown  by  Foster  (1961) 
to  block  the  acetylation  reaction  and  not  the  solvolysis.  The  inhibition  by 
indole  is  not  strictly  competitive  (Applewhite  et  al.,  1958). 


A'i 

Relative  —  AF  of  binding 

Inhibitor 

(mi/) 

(kcal/mole) 

Skatole 

0.5 

4.69 

Indole 

0.7 

4.43 

p-Nitrobenzoyl  -  d  -  try  ptophan 

1.5 

4.01 

Tryptamine 

2.0 

3.84 

p-Nitrobenzoyl-L-tryptophan 

2.3 

3.74 

Acetyl-D-tryptophan 

4.0 

3.41 

Acetyl-L-tryptophan 

6.0 

3.16 

D-Tryptophan 

10 

2.84 

L-Tryptophan 

20 

2.41 

The  relation  of  chymotrypsin  or  a  chymotrypsin-like  enzyme  to  hista- 
mine release  and  the  anaphylactic  reaction  was  examined  by  Austen  and 
Brocklehurst  (1960)  in  guinea  pig  lung  sensitized  to  ovalbumin.  Inhibitors 
such  as  /5-phenylpropionate,  indoleacetate,  and  indolepropionate  depress 
anaphylactic  histamine  release  more  than  50%  at  2.5  mM,  and  indole  and 
skatole  are  effective  at  1  mM  or  below.  Antigen  apparently  activates  some 
proteolytic  enzyme  necessary  for  the  release  of  histamine. 

*  This  terminology  connotes  an  entirely  new  way  of  looking  at  enzyme  catalysis. 


CHYMOTRYPSIN    AND    OTHER    PROTEOLYTIC    ENZYMES  375 

The  hydrolysis  of  a-benzoyl-L-argininamide  by  papain  is  inhibited  com- 
petitively by  carbobenzoxy-L-glutamate  (Kimmel  and  Smith,  1954;  Stockell 
and  Smith,  1957).  The  inhibition  is  very  sensitive  to  pH  between  3.9  and 
5,  the  inhibitory  activity  almost  disappearing  at  the  upper  end  of  this  range 
(see  Fig.  1-14-6).  The  y-carboxyl  group  has  a  pTiT,,  of  4.4  so  it  is  possible 
that  the  active  form  is  un-ionized: 

coo" 

/ 
HOOC— CH2CH2— CH  /^^ 

Vh^— COO-CH2 — (\       /) 

It  is  rather  surprising  that  the  ionization  of  this  group  should  be  so  im- 
portant in  the  binding,  especially  as  it  is  at  some  distance  from  the  other 
probable  binding  groups,  and  the  enzyme  as  a  whole  is  quite  positively 
charged  at  these  pH's  (isoelectric  point  near  8.75).  Perhaps  a  hydrogen- 
bonding  function  must  be  attributed  to  the  COOJI  group  instead.  Benzoyl- 
L-arginine  {K^  =  60  mM)  and  benzoyl-L-ar'gininamide  (iiC,  =  54  mM) 
also  inhibit  the  hydrolysis  of  benzoyl-L-arginine  ethyl  ester  (K,,,  =23  mM) 
by  ficin  at  pH  5.5  (Bernhard  and  Gutfreund,  1956). 

Beef  spleen  cathepsin  C  is  competitively  inhibited  by  various  amides, 
esters,  and  dipeptides  when  glycyl-L-tyrosinamide  is  the  substrate  (Fruton 
and  Mycek,  1956).  L-Phenylalanine  amides  and  esters  are  good  inhibitors 
(as  shown  in  the  accompanying  tabulation)  but  L-phenylalanine,  acetyl-L- 


Inhibitor 

(mM) 

L-Phenylalaninamide 

8.3 

L-Phenylalanine  ethyl  ester 

14 

L-Phenylalanyl-L-phenylalanine 

17 

L-Tyrosinamide 

22 

DL-Phenylalanylglycine 

25 

D-Phenylalaninamide 

68 

phenylalanine,  and  glycyl-DL-phenylalanine  are  inactive.  ^-Acetylation  or 
a  free  C00~  group  seems  to  prevent  binding.  The  interaction  between  the 
inhibitors  and  the  enzyme  may  involve  several  groups  with  possible  hydro- 
gen bonding  of  the  carbonyl  group  in  a  manner  similar  to  that  proposed  for 
chymotrypsin.  Decarboxylation  and  A^-acylation  of  amino  acids  can  lead 
to  inhibitors,  as  in  the  competitive  inhibition  by  tosylagmatine  of  thrombin 
(Ki  =  13.2  mM)  and  trypsin  {K,  =  3.45  mM)  (Lorand  and  Rule,  1961). 
This  inhibitor  markedly  slows  clotting  in  a  thrombin-fibrinogen  system  and 
in  whole  plasma. 


376  2.  ANALOGS  OF  ENZYME  KEACTION  COMPONENTS 

HEXOKINASES 

We  shall  now  turn  to  various  aspects  of  carbohydrate  metabolism  and 
begin  with  those  enzymes  responsible  for  the  initial  phosphorylation  of 
sugars.  Hexokinases  catalyze  the  reaction  between  two  types  of  substrate 
—  hexoses  and  ATP  —  and  analogs  of  either  may  inhibit.  Discussion  of  cer- 
tain particularly  important  glucose  analogs  (2-deoxy-D-glucose,  6-deoxy-6- 
fluoro-D-glucose,  and  related  compounds),  whose  actions  may  involve  not 
only  hexokinases  but  other  early  steps  in  carbohydrate  metabolism,  will 
be  postponed  to  the  following  section.  The  remaining  analog  inhibitors  may 
be  classed  as  (A)  hexoses,  (B)  hexose  phosphates,  (C)  glucosamine  and  deri- 
vatives, and  (D)  nucleotides  and  polyphosphates.  It  should  be  remembered 
that  the  values  of  K„,  and  K^  may  often  be  composite  in  that  the  sugar  or 
its  derivative  may  occur  in  solution  in  a  variety  of  forms  (for  example,  a- 
or  /^-isomers,  or  pyranose  or  furanose  rings).  These  forms  may  have  quite 
different  activities  and  different  individual  constants. 

Inhibition  by  Hexoses 

Most  hexokinases  are  not  specific  for  the  phosphorylation  of  one  sugar 
but  act  at  varying  rates  with  different  hexoses.  Thus  yeast  hexokinase 
phosphorylates  glucose  {K„i  =  0.16  mM),  fructose  {K„j  =  1.7  mM),  and 
mannose  (K,,,  =  0.1  mM);  brain  hexokinase  is  very  similar  (Slein  et  al., 
1950).  If  one  active  site  on  the  enzyme  is  responsible,  competition  between 
the  substrates  should  be  observed.  At  equimolar  concentrations,  glucose 
and  mannose  almost  completely  inhibit  the  phosphorylation  of  fructose 
(88-98%),  whereas  fructose  inhibits  the  phosphorylation  of  the  former 
hexoses  very  little  (15-18%),  which  would  be  expected  on  the  basis  of  their 
relative  iiC,„'s.  The  calculated  K/s  are  essentially  the  same  as  the  J^,„'s. 
The  inhibition  of  Schistosoma  fructokinase*  by  glucose  and  mannose  is  also 
quite  marked,  whereas  galactose  is  much  less  potent  (Bueding  and  Mac- 
Kinnon, 1955).  The  fructokinase  of  rat  intestinal  mucosa  shows  similar 
specificity,  the  following  inhibitions  being  observed  with  6  niM  inhibitor 
when  fructose  is  10  mM:  glucose  (90%),  mannose  (80%),  mannoheptulose 
(65%),  xylose  (50%),  allose  (0%),  and  galactose  (0%)  (Sols,  1956).  Ascites 
cell  glucokinase  is  inhibited  rather  well  by  talose  (jK",  =  3.2  mM)  and  altrose 
{K^  =  6  mM),  but  not  by  allose,  gulose,  or  idose  (Lange  and  Kohn,  1961  b). 
Yeast  glucokinase  is  competitively  inhibited  by  mannose  {K^  =  0.11  mM) 
(Fromm  and  Zewe,  1962).  Mannoheptulose  induces  a  temporary  diabetic 
state  by  blocking  the  uptake  of  glucose  into  the  tissues  through  inhibition 
of  glucokinase  (Coore  and  Randle,  1964). 

*  The  terms  "glucokinase,"  "fructokinase,"  etc.,  will  be  used  to  designate  kexo- 
kinases  when  the  corresponding  substrates  are  used  without  implying  that  the  en- 
zymes are  specific  for  these  substrates. 


HEXOKINASES  377 

Such  inhibitions  may  be  of  importance  in  the  metabolism  of  mixtures 
of  sugars.  Fructose  is  usually  phosphorylated  more  rapidly  than  glucose, 
but  in  mixtures  of  the  two  the  phosphorylation  of  fructose  is  markedly 
suppressed  and  essentially  only  glucose  is  metabolized.  Although  little 
is  known  of  the  nature  of  the  binding  of  these  hexoses  to  the  enzyme,  it 
would  appear  that  the  configurations  at  C-3  and  C-4  are  particularly  im- 
portant; for  example,  allose  differs  from  glucose  only  at  C-3  and  is  not 
readily  bound,  and  galactose  differs  from  glucose  only  at  C-4  and  is  much 
less  bound. 

Inhibition  by  Hexose  Phosphates 

Three  hexose  phosphates  have  been  found  to  be  fairly  specific  and  inter- 
esting inhibitors  of  hexokinases:  these  are  a-D-glucose-6-P,  a-L-sorbose-1-P, 
and  l,5-anhydro-D-glucitol-6-P.  Inhibiting  hexose  phosphates  should  be 
visualized  in  their  pyranose  or  furanose  forms,  since  it  is  likely  that  in- 
teraction with  the  enzyme  occurs  with  one  side  of  these  ring  structures. 
(See  formulas  on  page  378). 

Glyceraldehyde  has  been  known  for  many  years  to  be  an  inhibitor  of 
glycolysis  and  Lardy  et  al.  (1950)  in  a  study  to  determine  the  site  of  action 
found  that  L-glyceraldehyde  prevents  the  phosphorylation  of  glucose  or 
fructose  in  brain  extracts,  but  yet  has  no  direct  effect  on  the  hexokinase. 
This  paradox  was  resolved  by  showing  that  aldolase  catalyzes  the  conden- 
sation of  L-glyceraldehyde  with  glyceraldehyde-3-P  to  give  a  mixture  of 
D-fructose-1-P  and  L-sorbose-1-P.  The  latter  compound  was  found  to  be  a 
potent  inhibitor  of  hexokinase  (for  example,  0.08  mM  inhibits  67%  the 
phosphorylation  of  glucose  at  35  mM),  whereas  L-sorbose  and  L-sorbose-6-P 
are  inactive.  The  conventionally  written  structure  for  L-sorbose- 1-P 
(as  above)  bears  little  obvious  resemblance  to  D-glucose-6-P,  but  if  the 
former  structure  is  inverted  it  is  seen  that  the  molecules  are  identical  from 
one  side  and  differ  only  by  the  transposition  of  a  hydroxyl  group,  as  illus- 
trated by  Lardy  et  al.  with  molecular  models.  D-Glucose-6-P  is  the  product 

CH3— O— PO'a' 
O 


HO^ 


Q- L-Sorbopyranose- 1-P 

of  the  hexokinase  phosphorylation  of  glucose  and  is  an  inhibitor  of  the 
reaction  (see  below).  It  is  surprising  that  the  inhibition  by  L-sorbose-1-P 
is  not  competitive  with  glucose;  indeed,  as  the  glucose  concentration  is 
increased,  the  inhibition  becomes  somewhat  greater.  It  is  possible  that 


378 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


7^ 

/ 

"^o^ 

K 

pK 

\ 

^^^ 

—  O 

u- 

i 

\^ 

HEXOKINASES 


379 


competition  might  be  seen  if  rates  of  inhibition  in  the  presence  of  various 
concentrations  of  glucose  were  determined,  but  once  L-sorbose-1-P  has 
combined  with  the  enzyme  the  inhibition  is  essentially  irreversible.  It  is 
possible  that  glucose  forms  an  intermediary  tightly  bound  complex  with 
the  enzyme- ATP  during  the  reaction  and  that  L-sorbose-1-P  forms  a  similar 
complex  that  is  stable.  Actually  the  affinities  of  the  brain  enzyme  for 
D-glucose-6-P  (Kj  =  0.4  mM)  and  L-sorbose-1-P  {K^  =  0.7  mM)  are  close 
(Crane  and  Sols,  1954).  Other  hexokinases  may  not  be  so  susceptible  to 
L-sorbose-1-P,  since  Taylor  (1960)  found  only  slight  inhibition  by  0.5  mM  of 
glucose  uptake  by  Scenedesmus,  the  primary  transfer  site  being  hexokinase 
on  the  outside  of  the  membrane. 

A  closely  related  inhibitor  is  l,5-anhydro-D-glucitol-6-P  (1,5-D-sorbitan- 
6-P),  this  lacking  the  2-OH  group  in  L-sorbose-1-P  and  binding  somewhat 
less  tightly  {K^  =  1  mM)  to  the  brain  hexokinase  (Crane  and  Sols,  1954). 
The  nonphosphorylated  compound  is  a  very  weak  inhibitor  (Sols,  1956). 
Since  there  are  very  few  potent  and  specific  hexokinase  inhibitors,  Ferrari 
et  al.  (1959)  have  recently  investigated  l,5-anhydro-D-glucitol-6-P  in  some 
detail  to  see  if  it  might  be  useful  as  a  blocking  agent  of  this  enzyme  in  ho- 
mogenates.  It  is  stable  to  the  enzymes  attacking  D-glucose-6-P,  except  for 
hydrolysis  by  liver  glucose-6-phosphatase.  No  inhibition  of  glucose-6-P  de- 
hydrogenase is  evident,  but  it  inhibits  phosphoglucomutase  variably  (de- 
pending on  the  concentrations  of  glucose-1-P,  glucose-l,6-diP,  and  Mg++) 
and  phosphoglucose  isomerase  noncompetitively  at  higher  concentrations. 
At  6.25  mM  it  blocks  glucose  respiration  in  heart  homogenates  but  has 
no  influence  on  the  oxidation  of  glucose-6-P,  indicating  under  these  con- 
ditions a  rather  specific  inhibition  of  hexokinase. 

Brain  hexokinase  is  inhibited  by  glucose-6-P  whereas  yeast  hexokinase 
is  not  (L-sorbose-1-P  also  does  not  inhibit  the  yeast  enzyme),  and  the  inhi- 
bition has  been  found  to  be  noncompetitive  with  respect  to  both  glucose 
and  ATP  (Weil-Malherbe  and  Bone,  1951).  Inhibitions  by  various  hexose 
phosphates  have  been  studied  thorougly  by  Crane  and  Sols  (1953,  1954); 
the  accompanying  tabulation  summarizes  their  data.  The  following  are 
noninhibitory:  /?-D-glucose-l,6-diP,  D-mannose-6-P,  D-fructose-6-P,  D-fruc- 
tose-l,6diP,  D-arabinose-5-P,  D-ribose-5-P,  D-galactose-6-P,  a-glucose-l-P, 


Inhibitor 

imM) 

Relative 

!   —  JjP  of  binding 
(kcal/mole) 

a-D-Glucose-6-P 

0.4 

4.80 

a-D-Glucose-l,6-diP 

0.7 

4.46 

a-L-Sorbose-1-P 

0.7 

4.46 

l,5-Anhydro-D-glucitol-6-P 

1.0 

4.25 

a-D-Allose-6-P 

7.0 

3.05 

380  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

D-altrose-6-P,  glucuronate,  and  glucuronate-6-P.  It  may  be  noted  that  glu- 
cose is  the  only  hexokinase  substrate  that  forms  an  inhibitory  phosphate, 
indicating  the  importance  of  the  configuration  at  C-2  for  inhibition.  Thus 
glucose  phosphorylation  in  a  closed  system  slows  down  progressively  while 
that  of  mannose  is  linear,  a  phenomenon  which  may  be  significant  in 
regulating  the  rate  of  sugar  utilization.  In  a  purer  preparation  of  brain 
hexokinase.  Crane  and  Sols  confirmed  that  the  inhibition  is  not  formally 
competitive  but  that  a  reversible  EI  complex  is  formed.  Phosphorylation 
at  C-6  (or  at  C-1  in  the  sorbose  structure)  seems  necessary  for  inhibition; 
e.  g.,  glucose-1-P  and  1,5-anhydro-D-glucitol  lack  inhibitory  activity.  It  is 
interesting  that  the  inversion  of  the  phosphate-carrying  group  at  C-1  to 
form  /?-glucose-l,6-diP  abolishes  the  inhibition,  possibly  due  to  a  static 
interference  of  the  now  closely  apposed  phosphate  groups.  Inversion  of  the 
groups  on  C-2  (mannose-6-P),  C-3  (allose-6-P),  or  C-4  (galactose-6-P)  re- 
duces or  abolishes  the  inhibition;  it  was  felt  by  Crane  and  Sols  that  the 
configuration  at  C-3  influences  the  effect  of  an  adjacent  group  and  is  not 
directly  concerned  in  the  binding.  It  is  difficult  in  most  cases  to  decide  if 
the  change  in  affinity  on  inversion  of  the  groups  is  related  to  the  hydroxyl 
group  as  a  binding  site  or  as  producing  steric  hindrance;  thus  the  lack 
of  inhibition  by  galactose-6-P  could  be  due  either  to  the  loss  of  hydrogen 
bonding  through  the  OH  group  (occurring  in  glucose-6-P)  or  to  a  protru- 
sion of  the  OH  group  preventing  approach  of  the  pyranose  ring.  Compar- 
ison with  the  corresponding  deoxyglucose-6-P's  might  be  informative.  We 
cannot  do  this  for  C-4,  but  at  C-2  removal  of  the  OH  group  (2-deoxy- 
glucose-6-P)  abolishes  inhibition,  pointing  to  the  OH  group  as  a  binding 
site.  The  weak  inhibitory  activity  of  3-deoxyglucose-6-P  substantiates  the 
idea  that  the  3-OH  group  is  not  involved  in  binding.  The  retention  of 
inhibition  in  l,5-anhydro-D-glucitol-6-P  likewise  indicates  that  the  1-OH  is 
not  a  binding  site,  but  the  loss  of  inhibition  on  C-1  methylation  (a-meth- 
ylglucoside-6-P)  shows  that  steric  repulsion  occurs  when  the  C-1  group 
becomes  too  large.  One  may  conclude  that  binding  sites  are  at  the  2-OH 
and  6-phosphate  groups,  and  possible  at  the  4-OH  group. The  inhibitors 
thus  attach  to  a  different  set  of  enzyme  sites  than  the  substrates,  only 
C-4  being  common  to  both,  and  Crane  and  Sols  visualized  these  differences 
in  the  following  structures,  where  the  solid  circles  indicate  necessary  binding 
positions: 

►  H,OH  •H2— O— PO",' 

H  )r — \ 

OH 


Substrate  Inhibitor 


HEXOKINASES 


381 


Rat  intestinal  mucosa  hexokinase  is  inhibited  by  glucose-6-P  but  only 
about  one-tenth  as  readily  as  the  brain  enzyme  (Sols,  1956).  The  hexokinase 
of  Schistosoma  is  strongly  inhibited  by  glucose-6-P  when  glucose  or  mannose 
is  the  substrate,  but  fructose  phosphorylation  is  unaffected  (Bueding  and 
MacKinnon,  1955).  Ascites  tumor  hexokinase  behaves  like  the  brain  enzyme 
and  the  K,  for  glucose-6-P  is  0.4  mM  (McComb  and  Yushok,  1959).  Thus 
inhibition  of  various  hexokinases  by  glucose-6-P  has  been  observed,  but  the 
original  observation  that  the  yeast  enzyme  is  resistant  cannot  as  yet  be 
explained. 

Inhibition   by  D-Glucosamine  and   Derivatives 

Glucosamine  (2-amino-D-glucose)  is  phosphorylated  by  brain  hexokinase 
(Harpur  and  Quastel  1949)  and  it  was  postulated  by  Quastel  and  Cantero 
(1953)  that  it  might  be  carcinostatic  through  ATP  depletion.  However,  it 
also  competitively  inhibits  glucose  phosphorylation  and  any  carcinostatic 
activity  it  possesses  would  be  more  likely  related  to  this.  Maley  and  Lardy 
(1955)  thus  attempted  to  find  a  more  potent  inhibitor  among  the  A^-acylated 
derivatives  and  were  quite  successful,  as  shown  in  the  accompanying  tabu- 
lation. Furthermore,  these  derivatives  are  not  phosphorylated. 

Ki  (mM) 


Glucokinase 

Fructokinase 

iV^-(3,5-Dinitrobenzoyl)- 

0.011 

0.004 

A''-(m-Nitrobenzoyl)- 

0.033 

0.0084 

iV-(p-Nitrobenzoyl)- 

0.04 

0.05 

iV-Benzoyl- 

— 

0.036 

iV-(  w?  -  Aminobenzoyl)- 

0.15 

0.081 

i\/^-(p-Aminobenzoyl)- 

0.2 

0.11 

iV-Acetyl 

— 

0.46 

iV-Phenylacetyl- 

— 

0.86 

CH2OH 

nA 

/H 
\OH 
HON 

\ 

H 

1/ 

OH 


NH, 


CHoOH 


NH— CO 


O 


Q-D-Glucosamine 


iV- Benzoyl- o-D -glucosamine 


Before  considering  the  nature  of  this  inhibition,  let  us  examine  other 
hexokinases  to  determine  how  widespread  is  the  susceptibility.  The  fructo- 


382  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

kinases  of  Schistosoma  (Bueding  and  MacKinnon,  1955)  and  rat  intestinal 
mucosa  (Sols,  1956)  and  the  glucokinase  of  Spirochaeta  recurrentis  (P.  J.  C. 
Smith,  1960  b)  are  moderately  sensitive  to  glucosamine  (50-65%  inhibition 
by  6-10  mM)  and  more  sensitive  to  A^-acetylglucosamine  (75%  inhibition 
by  1-2  mM).  The  iC,  for  iV-acetylglucosamine  and  the  glucokinase  of  ascites 
tumor  cells  is  0.074  mM  (McComb  and  Yushok,  1959),  indicating  a  binding 
about  1  kcal/mole  tighter  than  for  glucose-6-P.  Furthermore,  both  glucos- 
amine and  iV-acetylglucosamine  inhibit  the  metabolism  of  glucose-C^*  and 
fructose-C^*  to  glycogen  and  COg  in  rat  liver  slices  (Spiro,  1958),  and  the 
A'^-(2?-nitrobenzoyl)  and  A^-(3,5-dinitrobenzoyl)  derivatives  inhabit  glucose 
uptake  by  Scenedesmus  (Taylor,  1960).  The  phosphorylation  of  glucosamine 
in  liver  extracts  is  competitively  inhibited  by  glucose  {K^  =0.11  mM), 
fructose,  A'^-acetylglucosamine,  and  hexose  and  glucosamine  phosphates, 
illustrating  mutual  interference  by  these  substrates  and  products  (McGar- 
rahan  and  Maley,  1962). 

In  all  these  instances  the  inhibitions  are  strictly  competitive  with  glucose 
or  fructose.  The  question  arises  as  to  why  the  iV-acylated  derivatives  are 
not  phosphorylated.  Maley  and  Lardy  (1955)  showed  by  molecular  models 
that  the  iV-acyl  groups  do  not  overlap  the  6-position  so  that  some  other 
explanation  must  be  sought.  It  was  suggested  that  the  A^-acyl  groups  might 
interfere  with  the  binding  of  ATP  to  the  enzyme,  but  it  is  also  possible  that 
they  shift  the  position  of  the  pyranose  or  furanose  rings  sufficiently  so 
that  the  6-position  is  not  favorably  oriented  for  phosphorylation.  It  may  be 
mentioned  that  no  carcinostatic  activity  was  noted  with  any  of  these  sub- 
stances when  tested  in  sarcoma-bearing  mice,  perhaps  due  to  the  hydrolysis 
of  the  iV-acyl  compounds  by  tissue  cathepsins. 

Kono  and  Quastel  (1962)  confirmed  the  glucosamine  inhibition  of  glyco- 
gen formation  in  rat  liver  slices  (50%  inhibition  by  around  0.8  mM)  and 
showed  there  to  be  no  depression  of  the  entry  of  glucose  into  the  cells.  Hexo- 
kinase,  phosphoglucomutase,  and  UDP-glucose  pyrophosphorylase  are  in- 
hibited quite  weakly  by  glucosamine,  significant  effects  being  exerted  only 
at  concentrations  over  20  mM.  UDP-glucose-glycogen  glucosy transferase 
in  inhibited  by  glucosamine  but  not  by  iV-acetylglucosamine,  which  inhi- 
bits glycogen  synthesis  as  does  glucosamine.  The  isolated  phosphorylase 
is  also  resistant  to  glucosamine.  Thus  the  enzymes  involved  in  glycogen 
formation  are  not  directly  inhibited  by  glucosamine  and  iV-acetylglucosa- 
mine.  However,  each  of  these  substances  stimulates  phosphorylase  activity 
in  slices  when  added  with  glucose.  Thus  the  explanation  for  the  reduced 
glycogen  formation  may  be  an  acceleration  of  glycogen  breakdown  and  not 
a  true  inhibition.  Silverman  (1963)  found  a  significant  reduction  of  glucose 
oxidation  by  glucosamine  in  the  rat  epididymal  fat  pad  in  the  presence  of 
insulin  and  felt  that  an  inhibition  of  glucokinase  could  not  entirely  account 
for  the  results,  so  that  one  must  assume  some  action  from  nonmetabolized 


HEXOKINASES  383 

products  formed  from  glucosamine,  possibly  glucosamine-P.  Glucosamine 
depresses  the  respiration  and  oxidation  of  pjTuvate  in  ascites  cells,  as 
does  glucose  (Crabtree  effect),  and  this  is  relieved  by  2,4-dinitrophenol 
(Ram  et  al.,  1963).  Since  this  is  presumably  related  to  the  phosphorylation 
of  glucosamine  and  the  effects  on  ADP-ATP  levels,  it  constitutes  another 
mechanism  whereby  glucosamine  can  alter  carbohydrate  metabolism. 

Inhibition  by  Adenine  Nucleotides  and  Polyphosphates 

The  reaction  rate  of  hexokinases  falls  with  time  due  to  the  accumulation 
not  only  of  glucose-6-P  but  also  of  ADP  (Sols  and  Crane,  1954).  Phosphate, 
pyrophosphate,  and  AMP  do  not  inhibit.  The  nature  of  the  ADP  inhibition 
appears  to  vary  with  the  source  of  the  hexokinase.  With  yeast  hexokinase 
the  inhibition  is  noncompetitive  with  respect  to  ATP  since  around  50% 
inhibition  is  produced  by  0.5  mM  ADP  at  all  levels  of  ATP  used  (Gamble 
and  Najjar,  1955),  and  with  Schistosoma  glucokinase  the  inhibition  actually 
increases  slightly  with  ATP  concentration  (Bueding  and  MacKinnon,  1955). 
The  inhibition  of  liver  fructokinase  by  ADP  is  reduced  slightly  by  increasing 
the  ATP  concentration  from  5  to  10  mM,  but  not  enough  to  indicate  pure 
competitive  inhibition;  Parks  et  al.  (1957)  stated  it  is  noncompetitive  but  it 
might  better  be  designated  as  mixed.  Echinococcus  fructokinase,  on  the 
other  hand,  is  inhibited  competitively  (Agosin  and  Aravena,  1959). 

Tripolyphosphate  (PgOjo^")  inhibits  the  fermentation  of  glucose  by  in- 
tact yeast  cells,  glycolysis  in  cell-free  extracts,  and  pure  hexokinase  (Vish- 
niac,  1950).  The  inhibition  of  hexokinase  is  quite  potent  when  (ATP)  =  3.75 
mM:  13%  at  0.47  mM,  31%  at  1.4  mM,  74%  at  4.7  mM,  and  93%  at  14 
mM  tripolyphosphate.  The  inhibition  is  reversed  by  both  ATP  and  Mg++. 
Wheat  germ  hexokinase  appears  to  be  more  resistant  to  tripolyphosphate, 
only  8%  inhibition  being  given  by  5  mM  (Saltman,  1953). 

Inhibition   by  Giucosone 

D-Glucosone  may  be  formed  from  D-glucose  by  mild  oxidation  (e.  g. 
with  Cu++)  at  C-2  but,  although  it  has  been  known  for  over  75  years,  its 
structure  is  not  completely  understood  (Becker  and  May,  1949).  The 
following  forms  are  possible  and  it  is  difficult  to  choose  between  them: 

CH,OH 

O 


a) 


384  2.  ANALOGS  OF  ENZYME  KEACTION  COMPONENTS 


OH  H  OH 

an)  (IV) 

There  is  some  evidence  that  the  enolic  tautomer  II  is  not  important  and  the 
behavior  with  enzymes  might  favor  structure  I.  Hynd  (1927)  at  St.  Andrews 
tested  D-glucosone  to  determine  if  it  could  counteract  insulin  hypoglycemic 
convulsions,  as  glucose  does,  but  found  that,  if  anything,  the  effect  of 
insulin  is  increased.  Glucosone  was  then  administered  to  normal  mice  giving 
toxic  symptoms  within  5  min  and  a  well-developed  insulin-like  reaction  in 
20  min.  The  lethal  dose  range  is  very  narrow,  2.4  mg/kg  being  nonlethal  and 
2.6  mg/kg  generally  lethal.  Glucose  injected  before  or  with  the  glucosone 
reduces  the  effects  somewhat.  Moribund  mice  following  lethal  doses  show 
an  elevation  in  blood  glucose  from  0.161  to  around  0.240  mg%;  thus  the 
symptoms  are  not  due  to  a  hypoglycemia.  Although  these  results  would 
point  to  glucosone  interference  with  the  utilization  of  glucose,  Hynd  un- 
fortunately assumed  that  glucosone  is  formed  from  glucose  by  the  action 
of  insulin  and  that,  indeed,  it  is  responsible  for  the  effects  of  insulin,  the 
raised  blood  glucose  levels  being  unexplained.  Similar  reactions  to  glucosone 
are  seen  in  several  species  (Herring  and  Hynd,  1928).  The  theory  that  insulin 
induces  glucosone  formation  was  made  untenable  by  Dixon  and  Harrison 
(1932),  who  found  no  glucosone  in  the  blood  during  insulin  convulsions. 

The  problem  rested  at  this  stage  for  20  years  and  then  was  taken  up  at  St. 
Andrews  (Bayne,  1952;  Mitchell  and  Bayne,  1952;  Johnstone  and  Mitchell, 
1953),  but  the  results  were  published  in  a  series  of  short  and  incomplete 
communications  without  adequate  data.  D-Glucosone  effects  in  mice  were 
not  seen  with  up  to  10  mg/kg  of  any  other  osone,  including  D-galactosone, 
D-arabinosone,  D-xylosone,  L-glucosone,  and  3-methyl-D-glucosone.  Turning 
to  yeast  glucose  fermentation,  they  found  no  inhibition  by  50  mM  D-gluco- 
sone but  marked  inhibition  at  200  mM,  whereas  L-glucosone  has  no  effect 
at  200  mM.  Becker  (1954)  reported  almost  complete  inhibition  of  the  aero- 
bic and  anaerobic  utilization  of  glucose  by  yeast  when  (glucosone)/(glu- 
cose)  =  5. 

It  was  realized  finally  by  Eeg-Larsen  and  Laland  (1954)  in  Oslo  that  the 
structural  similarity  of  glucosone  to  glucose  might  allow  the  former  to 
interfere  with  the  utilization  of  the  latter  by  blocking  its  phosphorylation. 
This  was  demonstrated  with  ox  brain  hexokinase;  glucosone  is  not  phospho- 
rylated  but  inhibits  glucose  phosphorylation  50%  at  0.35  mM  and  100% 
at  2.4  mM  when  glucose  is  2.4  mM.  They  concluded  that  the  inhibition  is 
noncompetitive,  but  the  small  range  of  glucose  concentrations  used  makes 


HEXOKINASES  385 

it  impossible  to  determine  the  type  of  inhibition;  actually  some  decrease 
in  the  inhibition  with  increasing  glucose  was  observed.  As  would  be  expected, 
glucosone  does  not  produce  a  Crabtree  effect  but  blocks  it  (Yushok  and 
Batt,  1957).  The  inhibition  of  glucose  fermentation  in  yeast  depends  on  the 
pH  and  the  buffer  system  present  (Hudson  rnd  Woodward,  1958).  At  pH 
6.5  in  phosphate  buffer  an  inhibition  of  73%  of  the  anaerobic  fermentation 
of  glucose  was  found  at  (glucosone) /(glucose)  =  2,  whereas  no  effect  was 
found  at  pH  3.5-4.5.  The  fermentation  and  phosphorylation  of  fructose  are 
inhibited  more  readily  than  with  glucose,  due  to  the  higher  ^,„  for  fructose, 
and  the  inhibitions  are  completely  competitive  with  K^  for  glucosone  around 
0.061  mM.  Marked  inhibition  of  anaerobic  glycolysis  in  rat  tissues  by  glu- 
cosone was  noted,  brain  being  much  more  sensitive  than  tumor  tissue;  in 
brain  complete  inhibition  occurs  with  (glucosone)/(glucose)  =  0.0067.  The 
susceptibility  of  brain  glycolysis  to  glucosone  is  certainly  much  greater  than 
of  any  hexokinase  studied  and  possibly  there  is  an  additional  site  of  action. 
In  any  event,  these  results  provide  sufficient  explanation  for  the  central 
toxic  actions  of  glucosone.  Despite  the  fact  that  glucosone  can  be  formed 
in  certain  organisms  (e.  g.,  molluscan  crystalline  styles  and  red  algae)  and 
can  be  metabolized  in  streptococci  and  mammals,  it  would  appear  that  it 
is  not  an  important  substance  in  intermediary  metabolism  and  is  not  gen- 
erally on  the  pathway  for  the  synthesis  of  glucosamine  (Becker  and  Day, 
1953;  Topper  and  Lipton,  1953;  Dorfman  et  at.,  1955;  Bean  and  Hassid, 
1956).  There  is  thus  no  evidence  that  glucosone  can  participate  in  the  reg- 
ulation of  carbohydrate  metabolism,  but  the  high  susceptibility  of  brain 
glycolysis  suggests  that  one  should  withhold  final  judgment  until  the  ab- 
sence of  glucosone  in  the  body  under  various  conditions  has  been  demon- 
strated. 

Inhibition  of  Phosphafructokinase   by  Cycle  Intermediates 

The  phosphofructokinase  from  sheep  brain  is  quite  potently  inhibited  by 
certain  cycle  intermediates  (see  accompanying  tabulation)  (Passonneau  and 
Lowry,  1963).  Although  this  is  not  strictly  analog  inhibition,  it  is  worth 


Inhibitor 

Ki  (mJ/) 

Citrate 

0.03 

cis-Aconitate 

0.1 

Isocitrate 

0.2 

Malate 

0.6 

Succinate 

1.5 

a-Ketoglutarate 

2.5 

Fumarate 

>10 

386  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

mentioning  because  of  the  implications  such  actions  have  for  a  feedback 
control  of  carbohydrate  oxidation.  A  rise  in  the  levels  of  the  inhibitory- 
intermediates  would  reduce  the  rate  of  formation  of  pyruvate.  The  steady- 
state  concentrations  of  the  cycle  intermediates  are  certainly  high  enough 
to  inhibit  significantly,  but  the  problem  of  compartraentalization  arises 
since  it  is  generally  assumed  that  the  tricarboxylates  particularly  are 
mostly  confined  to  the  mitochondria.  Whatever  the  significance  of  this 
type  of  inhibition,  it  emphasizes  the  importance  of  compartmentalization 
in  regulatory  control  of  metabolism,  a  factor  which  has  not  always  been 
taken  into  account. 

EFFECTS    OF   2-DEOXY-D-GLUCOSE 
ON    CARBOHYDRATE    METABOLISM 

An  inhibitor  capable  of  specifically  blocking  the  glycolytic  pathway 
would  not  only  be  a  valuable  tool  in  biochemical  investigation  but  might 
play  a  role  in  the  chemotherapy  of  certain  neoplasms.  Glucose  analogs, 
especially  those  entering  the  pathway  and  forming  inhibitory  intermediates, 
would  be  the  most  likely  candidates,  and  2-deoxy-D-glucose  (2-DG)  is  the 
most  interesting  and  best  understood  substance  of  this  type.  The  volume 
of  literature  during  the  past  10  years  on  this  analog  precludes  a  complete 
discussion  and  emphasis  will  be  placed  on  the  sites  and  mechanisms  of  the 
inhibition  in  the  glycolytic  pathway.  2-DG  was  first  examined  by  Cramer 
and  Woodward  (1952)  at  the  Franklin  Institute  in  the  course  of  searching 
for  carcinostatic  glucose  analogs,  and  they  found  that  it  does  indeed  produce 
some  regression  of  Walker  carcinoma  and  terminates  embryonic  develop- 
ment in  rats.  2-DG  differs  from  glucose  in  the  substitution  of  the  2-OH 
group  by  a  hydrogen  atom  and  may  be  represented  in  the  pyranose  form  as: 


2 -Deoxy-D -glucose 

Absorption,  Distribution,  and  Metabolism  of  2-DG 

It  will  be  well  to  discuss  the  uptake  and  fate  of  2-DG  in  cells  before  turn- 
ing to  the  metabolic  disturbance  produced.  2-DG  enters  most  cells  readily; 
this  may  involve  a  phosphorylation  at  the  membrane  in  some  cases,  but 
in  others  it  is  phosphorylated  only  after  entry.  Whatever  the  transport 
mechanism  there  is  usually  competition  between  2-DG  and  glucose,  the 


EFFECTS     OF    2-DEOXY-D-GLUCOSE  387 

uptake  of  2-DG  being  depressed  progressively  as  the  glucose  concentration 
is  increased  in  rat  diaphragm  (Nakada  and  Wick,  1956;  Kipnis,  1958) 
and  lymph  node  cells  (Helmreich  and  Eisen,  1959).  Nakada  and  Wick 
(1956)  showed  that  insulin  can  double  the  rate  of  2-DG  uptake  by  dia- 
phragm, and  Kipnis  and  Cori  (1959,  1960)  studied  this  in  greater  detail.  In 
normal  diaphragm  the  2-DG  taken  up  appears  as  2-deoxy-D-glucose-6-phos- 
phate  (2-DG-6-P),  the  rate  of  phosphorylation  being  apparently  greater 
than  the  rate  of  penetration.  Diabetic  diaphragm  takes  up  and  phosphory- 
lates  2-DG  at  a  reduced  rate  but  2-DG  does  not  accumulate  in  the  cells, 
indicating  the  penetration  is  still  rate-limiting.  Addition  of  insulin  acceler- 
ates the  uptake  and  some  free  2-DG  appears  in  the  cells  so  that  the  phos- 
phorylation is  not  increased  proportionately.  Epinephrine  does  not  in- 
fluence the  penetration  but  slows  phosphorylation  of  2-DG.  Transport  of 
2-DG  across  the  entire  diaphragm  is  slow,  being  about  one-fifth  the  rate 
for  glucose  and  one-twenty-fifth  the  rate  for  galactose  (Ungar  and  Psy- 
choyos,  1963).  It  is  possible  that  it  is  trapped  in  the  muscle  as  2-DG-6-P 
since  insulin  depresses  the  transfer.  The  uptake  of  2-DG  by  yeast  in  glucose- 
phosphate  medium  is  5-10  times  faster  aerobically  than  anaerobically; 
when  glucose  is  omitted  the  aerobic  uptake  of  2-DG  is  not  altered,  but  an- 
aerobically there  is  a  loss  of  2-DG  from  the  cells,  so  that  the  uptake  is  de- 
pendent on  aerobic  processes  and  probably  on  the  level  of  ATP  since  2,4- 
dinitrophenol  acts  like  anaerobiosis  (Kiesow,  1959).  Certain  fungi,  such  as 
Neurospora  crassa  and  Aspergillus  oryzae,  can  grow  with  2-DG  as  the  sole 
source  of  carbon  (Sols  et  al.,  1960  b)  but  not  as  rapidly  as  with  glucose; 
indeed,  growth  with  glucose  or  other  sugars  is  inhibited  by  2-DG.  E.  coli 
will  not  grow  with  2-DG  as  the  only  carbon  source  and  there  is  some  evi- 
dence that  it  does  not  penetrate  into  the  cells  (Gershanovich,  1963).  The 
evidence  for  the  lack  of  entrance  is  that  glycolysis  is  not  inhibited  in  intact 
cells  where  it  is  in  extracts.  2-DG  diffuses  across  the  intestinal  wall  but  is 
not  actively  transported  as  are  glucose,  1-DG,  and  3-DG,  indicating  the 
importance  of  the  2-position  in  transport  (Wilson  and  Landau,  1960), 
nor  does  2-DG  have  an  effect  on  the  short-circuit  current  through  the  in- 
testinal wall,  such  being  associated  with  transport  (Schultz  and  Zalusky, 
1964;  Barry  et  al,  1964). 

There  is  some  information  on  the  disposal  of  2-DG  in  intact  animals. 
Injected  into  rabbits,  it  is  rapidly  distributed  in  the  extracellular  space 
and  some  enters  the  tissues  in  eviscerated  and  nephrectomized  animals,  the 
uptake  being  markedly  stimulated  by  insulin  (Wick  et  al.,  1955).  Since  it 
produces  a  block  of  glucose  uptake  for  at  least  8  hr,  it  is  evident  that  little 
2-DG  is  metabolized  beyond  the  2-DG-6-P  stage  in  extrahepatic  tissues. 
This  is  confirmed  by  the  finding  that  little  or  no  C^^Oa  is  expired  following 
injection  of  2-DG-Ci*  (Wick  et  al,  1957).  Blood  levels  of  2-DG  are  more 
consistent  after  subcutaneous  injection  than  when  it  is  given  intraperito- 


388 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


neally,  due  to  erratic  absorption  from  the  peritoneum  (Ball  and  Saunders, 
1958).  After  subcutaneous  administration  there  is  a  blood  peak  at  15  min, 
after  which  there  is  a  gradual  fall  over  6  hr.  In  human  subjects  infused 
intravenously  with  50-200  mg/kg  of  2-DG  over  30-min  periods,  approxi- 
mately 30%  is  excreted  in  the  urine  (Landau  et  al.,  1958). 

The  pathways  of  2-DG  metabolism  have  not  been  completely  worked  out. 
The  accompanying  scheme  shows  some  of  the  reactions  encountered.  The 
phosphorylation  by  hexokinase  to  2-DG-6-P  would  seem  to  be  the  most 
important  reaction,  especially  as  2-DG-6-P  is  not  metabolized  in  most  cells 
and  tends  to  accumulate.  HeLa  cells  can  oxidize  2-DG-6-P  but  much  more 
slowly  than  glucose-6-P  (Barban  and  Schulze,  1961).  Hexokinases  for  the 

deoxy-disaccharides 


2-deoxy-D-glucose 


6 -deoxy-D-glucono  lactone 


^^  2-deoxy-D-gluconate 

\ 

2-deoxy-D-glucose-6-P  »- 2-deoxy-D-gluconate- 6- P 

i  \ 

(oxidized)  D-mannonate-6-P 

formation  of  2-DG-6-P  have  been  found  in  brain  (Sols  and  Crane,  1953), 
kidney,  intestine,  liver  (Lange  and  Kohn,  1961  a),  skin  (Brooks  etal.,  1959), 
diaphragm(  Kipnis  and  Cori,  1959),  HeLa  cells  (Barban  and  Schulze,  1961), 
ascites  carcinoma  cells  (McComb  and  Yushok,  1959;  Lange  and  Kohn,  1961 
b),  and  Neurospora  crassa  (Sols  et  al.,  1960  b).  The  Michaelis  constants  for 
2-DG  are  usually  higher  than  for  glucose  (see  tabulation)  but  the  rates  of 


Hexokinase  K^  {mM) 

Source 

Glucose 

2-DG 

Brain 

0.01 

0.024 

Ascites  carcinoma 

0.04 

0.069 

Intestine 

0.065 

0.09 

Kidney 

0.048 

0.04 

Liver 

0.04 

0.09 

phosphorylation  are  often  comparable.  It  is  interesting  that  a  strain  of 
HeLa  cells  resistant  to  2-DG  has  been  obtained,  and  they  are  defective  in 
hexokinase  or  contain  a  hexokinase  inhibitor;  the  phosphorylation  rates  for 
2-DG,  glucose,  fructose,  and  mannose  are  all  lower  than  normal  (Barban, 
1961).  Resistance  is  also  associated  with  a  5-  to  10-fold  increase  in  alkaline 
phosphatase  activity  and  this  may  partly  account  for  the  slower  rate  of 


EFFECTS    OF   2-DEOXY-D-GLUCOSE  389 

accumulation  of  2-DG-6-P  (Barban,  1962  a,  b).  How  much  of  the  2-DG  is 
oxidized  directly  is  generally  unknown,  but  the  glucose  oxidase  from 
As'pergillus  niger  oxidizes  it  fairly  well:  the  relative  rates  of  oxidation  are 
glucose  (100),  2-DG  (20),  3-DG  (1),  4-DG  (2),  5-DG  (0.05),  and  6-DG  (10) 
(Pazur  and  Kleppe,  1964).  The  direct  oxidation  of  2-DG  is  apparently 
catalyzed  by  a  variety  of  enzymes,  some  of  the  notatin  type  (Sols  and  de 
la  Fuente,  1957)  and  some  of  the  glucose  dehydrogenase  type  (Williams 
and  Eagon,  1959).  The  further  metabolism  of  2-deoxy-D-gluconate  probably 
varies  with  the  tissue  and  has  been  shown  in  the  scheme  above  for  skin 
(Brooks  et  at.,  1960).  Other  pathways  for  2-DG  metabolism  may  occur  in 
plants,  since  Kocourek  et  al.  (1963)  have  provided  evidence  for  (1)  /?- 
glucosidation  probably  on  C-6,  (2)  oxidation  on  C-1  to  form  2-deoxyhexo- 
nate  lactone,  and  (3)  epimerization  to  2-deoxygalactose  in  tobacco  plants 
taking  up  2-DG  through  the  roots.  The  last  reaction  involves  three  enzymes 
and  the  epimerization  occurs  in  a  complex  with  UDP.  The  abnormal  di- 
saccharide,  /5-D-fructofuranosyl-2-deoxy-D-glucose,  has  been  insolated  from 
the  excised  leaves  of  several  plants  following  incubation  with  2-DG  (Barber, 
1959).  It  is  possible  that  a  number  of  abnormal  polysaccharides  containing 
2-DG  will  eventually  be  found. 

Effects  of  2-DG  and  2-DG-6-P  on  Glycolytic  Enzymes 

2-DG  inhibits  the  utilization  of  glucose  and  other  sugars  in  many  organ- 
isms and  tissues,  and  we  shall  now  attempt  to  localize  the  site  of  this 
inhibition  in  the  early  phases  of  glycolysis.  We  must  consider  not  only 
2-DG  but  also  its  primary  metabolic  product,  2-DG-6-P,  as  inhibitors. 
The  most  likely  sites  for  inhibition  would  be  (1)  2-DG  on  hexokinases,  or 
(2)  2-DG-6-P  on  phosphoglucose  isomerase,  6-phosphofructokinase,  or  al- 
dolase with  respect  to  glucose  metabolism.  Since  2-DG  is  phosphorylated 
about  as  well  as  glucose  by  hexokinases  it  is  clear  that  some  competitive 
inhibition  would  be  observed  under  certain  circumstances,  a  suggestion  first 
made  by  Cramer  and  Woodward  (1952).  However,  this  would  appear  to 
be  generally  an  unimportant  factor  in  the  over-all  glycolytic  inhibition, 
since  2-DG  equimolar  with  glucose  does  not  inhibit  the  glucokinase  of  ascites 
cells  (Nirenberg  and  Hogg,  1958)  and  at  10  times  the  glucose  concentration 
does  not  inhibit  HeLa  cell  glucokinase  (Barban  and  Schulze,  1961).  The 
situation  may  be  somewhat  more  complex  in  certain  tissues,  however,  in- 
asmuch as  rat  liver  contains  two  glucose-phosphorylating  enzymes,  called 
glucokinase  and  hexokinase  (Walker  and  Rao,  1963).  The  sensitivities  of 
these  enzymes  are  quite  different  (see  accompanying  tabulation)  and  a 
kinetic  analysis  was  made,  the  hexokinase  being  studied  without  interfer- 
ence by  the  glucokinase  since  fetal  liver  contains  only  the  former.  The 
effect  of  the  various  inhibitors,  which  are  all  competitive,  varies  in  a  com- 
plex fashion  as  the  glucose  concentration  is  changed  because  of  the  vary- 


390  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


Ki  (mM) 

Inhibitor 

Rat  liver 

Rat  liver 

Guinea  pig 

glucokinase 

hexokinase 

hexokinase 

2-DG 

14 

0.3 

0.6 

D-Glucosamine 

0.8 

0.3 

0.2 

iV-Acetyl-D-glucosamine 

0.5 

0.2 

0.3 

Glucose  (K,„) 

10 

0.04 

0.03 

ing  importance  of  each  enzyme,  and  it  was  pointed  out  that  it  is  very  dif- 
ficult to  assess  the  effects  of  such  analogs  in  adult  liver. 

Most  of  the  emphasis  recently  has  been  placed  on  the  inhibition  of 
phosphoglucose  isomerase  by  2-DG-6-P.  This  inhibition  is  competitive  on 
the  enzyme  from  rat  kidney  (Wick  et  al.,  1957),  rat  muscle  (Ferrari  et  al., 
1959),  and  ascites  cells  (Nirenberg  and  Hogg,  1958).  The  inhibition  is  reas- 
onably potent  (see  tabulation  for  kidney  enzyme)  and  it  would  be  quite 


Glucose-6-P  2-DG-6-P 

{mM)  (mM) 


%  Inhibition 


I  0.5  9 

1  1  24 

0.5  1  79 


easy  for  the  2-DG-6-P  concentration  to  become  greater  than  the  glucose-6-P 
concentration  in  cells,  especially  as  the  former  is  usually  not  metabolized 
and  accumulates.  The  inhibition  of  the  skeletal  muscle  enzyme  is  very  simi- 
lar. Unfortunately  the  inhibitions  of  6-phosphofructokinase  and  aldolase 
by  2-DG-6-P  have  not  yet  been  adequately  examined,  so  one  is  left  with 
phosphoglucose  isomerase  as  the  site  of  the  primary  block,  the  conclusion 
of  Wick  et  al.  (1957).  However,  another  possible  site  for  inhibition  is  the 
membrane  transport  system  for  glucose,  as  suggested  by  the  work  of  Kipnis 
and  Cori  (1959),  and  this  might  be  by  either  2-DG  or  2-DG-6-P.  Furthermore, 
Nirenberg  and  Hogg  (1958)  reported  that  the  metabolism  of  fructose-1,6- 
diP  is  blocked  by  2-DG-6-P  and  stated  that  some  inhibition  must  occur 
after  the  phosphofructokinase  step  (which  could  be  on  aldolase).  Other 
inhibitions  on  enzymes  metabolizing  glucose  but  not  on  the  main  glycolytic 
pathway  may  be  mentioned.  Rat  liver  microsomal  glucose-6-phosphatase 
is  inhibited  weakly  by  2-DG  (Hass  and  Byrne,  1960),  HeLa  cell  glucose-6-P 


EFFECTS    OF    2-DEOXY-D-GLUCOSE  391 

dehydrogenase  is  inhibited  noncompetitively  by  2-DG-6-P  (Barban  and 
Schulze,  1961),  2-DG-6-P  competitively  inhibits  the  activation  of  rat  liver 
glycogen  synthetase  by  glucose-6-P  (Steiner  et  al.,  1961),  and  UDP6: 
a- l,4-glucan-o;-4-glucosy transferase  from  dog  muscle  is  inhibited  by  2-DG- 
6-P  with  Kj  =1.3  mM  (Rosell-Perez  and  Larner,  1964).  The  importance 
of  these  inhibitions  in  the  interference  produced  by  2-DG  on  glucose  me- 
tabolism is  not  understood. 

The  block  of  fructose  utilization  by  2-DG  may  well  present  a  different 
problem.  Fructose  phosphorylation  is  inhibited  much  more  readily  than 
glucose  phosphorylation,  presumably  due  to  the  lower  affinity  of  the  hexo- 
kinases  for  fructose  (Sols,  1956;  Nirenberg  and  Hogg,  1958;  Barban  and 
Schulze,  1961).  The  inhibition  of  phosphoglucose  isomerase  could  not  explain 
the  suppression  of  fructose  utilization  inasmuch  as  the  fructose  pathway 
bypasses  this  step.  In  rat  adipose  tissue  2-DG  has  very  little  effect  on  the 
metabolism  of  fructose  although  glucose  metabolism  is  quite  strongly 
depressed  (Fain,  1964).  With  glucose  at  2.8  mM  and  2-DG  at  1.4  mM,  the 
formation  of  COg  is  reduced  70%  and  of  fatty  acids  89%.  Nevertheless,  the 
stimulatory  effect  of  insulin  on  fructose  utilization  is  blocked  by  2-DG 
whereas  the  effects  of  insulin  on  glucose  are  unaltered.  Certainly  different 
tissued  and  organisms  must  have  various  transport  and  enzyme  systems  for 
the  metabolism  of  fructose,  so  one  should  not  expect  a  uniform  action  of 
2-DG.  The  enzymes  involved  in  the  metabolism  of  fructose- 1-P  or  fruc- 
tose-6-P  have  not  been  studied  with  respect  to  2-DG  inhibition. 

Effects  of  2-DG  on  Carbohydrate  Metabolism  and  Respiration 

Investigations  on  isolated  enzymes  have  indicated  an  important  block 
of  phosphoglucose  isomerase  by  2-DG-6-P  and  contributory  inhibition  of 
hexokinases  under  certain  conditions.  Let  us  now  turn  to  studies  on  carbo- 
hydrate metabolism  in  intact  cells  and  tissues  in  order  to  determine  if  the 
effects  of  2-DG  can  be  explained  adequately  on  this  basis,  or  to  accumulate 
evidence  of  blocks  elsewhere.  Anaerobic  glycolysis,  aerobic  glycolysis, 
and  glucose  respiration  are  inhibited  by  2-DG  but  to  very  different  degrees 
(Fig.  2-10).  Indeed,  respiration  is  inhibited  only  with  high  concentrations, 
usually  30-100  times  that  required  to  inhibit  anaerobic  glycolysis  compara- 
bly (Woodward  and  Hudson,  1954;  Tower,  1958),  so  that  some  workers 
have  reported  that  respiration  is  not  inhibited  (Fridhandler,  1959;  Taylor, 
1960).  In  the  case  of  sea  urchin  eggs,  the  inhibition  can  be  almost  completely 
counteracted  by  increasing  glucose  concentration  (Bernstein  and  Black, 
1959).  However,  glucose  must  be  present  when  the  2-DG  is  added  and  is 
ineffective  when  the  inhibition  has  developed.  The  respiration  of  guinea  pig 
skin  in  the  presence  of  various  substrates  is  inhibited  by  2-DG  moderately 
and  progressively  (see  accompanying  tabulation)  (Carney  et  al.,  1962). 
All  substrates  and  2-DG  were  20  mM.  There  is  no  effect  on  the  endogenous 


392 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


100 


50 


% 

INH 


ANAEROBIC 
GLYCOLYSIS 


0.1 


(2-06) 


Fig.  2-10.    Effects   of  2-DG   on   the   glucose   metabolism  in   cat  brain  slices. 
(From  Tower,  1958.) 

respiration.  The  small  effect  on  galactose  respiration  was  felt  to  be  due 
perhaps  to  a  different  hexokinase  being  used  for  the  phosphorylation  of 
galactose,  since  inhibition  is  exerted  by  2-  and  4-deoxy galactose. 


Substrate 

%  Respiratory  inhibition  at: 

0-2  hr 

22 

-24  hr 

Glucose 

12 

53 

Mannose 

10 

43 

Fructose 

8 

50 

Galactose 

3 

21 

PjTuvate 

0 

1 

Glycolysis  as  measured  by  the  formation  of  C^^Og  from  glucose-u-C^^ 
is  depressed  by  2-DG  in  diaphragm  (Nakada  and  Wick,  1956),  kidney  (Serif 
and  Wick,  1958),  lymph  node  cells  (Helmreich  and  Eisen,  1959),  and  adipose 
tissue  (Brooks  et  at.,  1961).  The  variation  of  inhibition  with  2-DG  concen- 
tration is  shown  in  Fig.  2-11  for  rat  kidney  slices.  Glycolysis  as  measured 
by  unlabeled  COg  or  lactate  formation  is  also  inhibited  in  yeast  (Cramer 
and  Woodward,  1952),  ascites  carcinoma  and  leukemic  cells  (Laszlo  et  al., 
1958),  and  cerebral  cortex  slices  (Tower,  1958).  The  depression  of  aerobic, 
and  anaerobic  glycolysis  in  tumor  tissue  is  counteracted  by  increasing  glu- 
cose concentrations  (Woodward  and  Hudson,  1954).  These  results  are  quite 


EFFECTS    OF    2-DEOXY-D-GLUCOSE 


393 


consistent  in  showing  an  inhibition  of  the  glycolytic  pathway  by  2-DG. 
However,  the  formation  of  CO2  from  glucose  is  not  always  depressed.  Frid- 
handler  (1959)  found  that  although  2-DG  inhibits  anaerobic  glycolysis 
in  rabbit  blastocysts,  the  rates  of  respiration  and  COg  formation  are  not 
significantly  affected.  The  formation  of  C^^Og  from  glucose-1-C^*  in  human 
fetal  liver  is  actually  increased  by  2-DG,  but  this  may  be  attributed  to  a 
partial  inhibition  of  glycolysis  (Villee  and  Loring,  1961).  In  ascites  carci- 
noma cells  2-DG  simultaneously  inhibits  the  glycolysis  of  glucose  and 


X 

INH 


100 

^^^--''^  -  DE ox  Y  -  D  -  GLUCOSE 

. 

^^^"^"^^^ 

50 

■    /^^^ 

< 

/              ^^ 6- DEOXY- 6- FLUORO-D-GLUCOSE 

10 


20 


30 


40 


50 


60 


(I) 


70 

mM 


Fig.   2-11.   Inhibition  of  the   formation   of  C^^Oa   from   glucose-u-C* 

by  glucose  analogs  in  rat  kidney  slices.  Glucose  was  10  vaM.  (From 

Serif  and  Wick,   1958.) 


increases  its  oxidation  (Christensen  et  al.,  1961),  while  the  disappearance  of 
glucose  from  the  medium  is  reduced  (Fig.  2-12).  The  increased  C^^Oj  formed 
from  glucose  coupled  with  the  depressed  glucose  uptake  must  be  taken  to 
mean  that  the  pathway  of  glucose  utilization  has  been  markedly  altered, 
i.  e.,  less  glucose  is  going  to  lactate  and  more  is  being  oxidized.  One  impor- 
tant factor  in  such  tissues  must  be  the  activity  of  the  pentose-P  pathway, 
which  is  apparently  not  directly  inhibited  but  is  indirectly  stimulated  by 
2-DG.  The  formation  of  C^^Og  from  glucose-6-C^*  in  calf  thymus  nuclei  is 
inhibited  essentially  completely  by  2-DG  (McEwen  et  al.,  1963  b).  However, 


394 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


the  C-l/C-6  ratio  in  brain  slices  remains  the  same  when  ghicose  uptake  is 
reduced  to  one  third  by  2-DG  (Tower,   1958). 

Since  hexose  uptake  into  cells  is  often  coupled  with  phosphorylation, 
one  would  expect  2-DG  to  inhibit  this  uptake  by  suppressing  kinase  activity 
directly  or  indirectly.  Lymph  node  cells  treated  with  2-DG  until  2-DG-6-P 
is  formed  and  then  washed  free  of  2-DG  do  not  accumulate  glucose,  fructose, 
or  mannose,  and  lactate  formation  is  markedly  reduced  (Helmreich  and 
Eisen,  1959).  The  uptake  of  glucose  into  chick  embryo  hearts  is  35-45% 


ISO 


0,    FORMED 


30 


60 


90 


(2-06) 


120 

mM 


Fig.  2-12.  Effects  of  2-DG  on  the  glucose  metabolism  in  Ehrlich  ascites  carcinoma 
cells.  Glucose-u-C*  was  10  mM.  Control  for  residual  glucose  taken  from  nonincubated 
flasks  with  no  2-DG.  Experiments  2  and  4  were  averaged.  (Data  from  Christensen 

et  ah,   1961.) 


reduced  by  40  mM  2-DG  (Modignani  and  Foa,  1963)  and  into  carrot  slices 
is  reduced  to  about  the  same  degree  by  10  mM  2-DG,  this  inhibition  not 
being  overcome  by  increase  in  glucose  concentration  (Grant  and  Beevers, 
1964).  2-DG  interferes  with  galactose  uptake  into  mouse  strain  L  cells  but 
not  potently  {K^  =  7.2  niM)  (Maio  and  Rickenberg,  1962).  It  is  quite  likely 
that  these  inhibitions  are  exerted  predominantly  by  2-DG-6-P.  On  the 
other  hand,  the  accumulation  of  a-methylglucoside,  which  is  transported 
into  E.  coli  by  the  glucose  carrier,  is  well  inhibited  by  1-DG,  poorly  by 
6-DG,  and  not  at  all  by  3-DG  (Hagihira  et  al,  1963).  Apparently  2-DG 
is  not  as  potent  an  inhibitor  here  as  1-DG. 

Another  mechanism  by  which  2-DG  could  alter  carbohydrate  uptake  and 
metabolism  is  by  changing  the  levels  of  Pj,  ADP,  and  ATP,  since  the  rates 


EFFECTS     OF    2-DEOXY-D-GLUCOSE  395 

of  hexose  phosphorylation  and  glycolytic  breakdown  are  controlled  by  these. 
In  brain  slices,  10  mM  2-DG  causes  a  50%  fall  in  creatine-P  and  almost 
complete  disappearance  of  the  adenosine  polyphosphates  (Tower,  1958). 
There  is  also  a  diversion  of  phosphate  to  the  stable  2-DG-6-P  so  that  a  cer- 
tain amount  of  phosphate  is  removed  from  glycolytic  participation,  as  also 
pointed  out  by  Kiesow  (1960  c).  Furthermore,  2-DG  has  been  shown  to 
inhibit  the  incorporation  of  P/^  into  ADP  and  ATP  in  ascites  carcinoma 
cells  70%  at  10  mM  (Greaser  and  Scholefield,  1960).  El'tsina  and  Beresot- 
skaya  (1962)  determined  P,  and  nucleotide  levels  in  tumor  cells  exposed  to 
11  raM  2-DG  and  found  marked  decreases  in  ATP  and  ADP  (see  accompa- 
nying tabulation).  On  the  other  hand,  rat  liver  and  kidney  slices  show  no 


Tumor  ( 

components 

Control 

2-DG 

Zahdel 

hepatoma 

ATP 

69.2 

7.7 

ADP 

28.9 

13.2 

AMP 

— 

11.7 

P. 

157 

39 

Sarcoma  37 

ATP 

74.7 

1.7 

ADP 

40.5 

5.4 

AMP 

14.2 

3.5 

P, 

140 

67 

significant  changes,  a  difference  attributed  to  variations  in  the  activity 
and  cellular  location  of  hexokinases.  McComb  and  Yushok  (1964  a)  also 
reported  marked  falls  in  ATP  in  ascites  cells  within  12  min  after  exposure 
to  2-DG,  and  a  65%  net  loss  of  the  cellular  adenine  nucleotides.  The  disap- 
pearance of  nucleotides  is  at  least  partly  accounted  for  by  the  phosphory- 
lation of  2-DG  by  hexokinase,  the  formation  of  AMP  mediated  by  adeny- 
late kinase,  the  deamination  of  AMP  to  IMP,  and  the  splitting  of  IMP 
to  inosine  by  5 '-nucleotidase  (McComb  and  Yushok,  1964  b).  They  observed 
a  steady  rise  in  inosine,  correlated  with  a  rise  and  subsequent  fall  in  IMP. 
Changes  in  nucleotide  levels  should  affect  various  phases  of  metabolism 
which  involve  these  substances,  and  this  is  well  seen  in  the  effects  of  2-DG 
on  the  oxidation  of  ethanol  in  yeast  (Maitra  and  Estabrook,  1962).  When 
2-DG  is  added  to  previously  starved  yeast  in  the  presence  of  ethanol, 
there  is  acceleration  of  respiration  and  the  oxidation  of  NADPH,  accompa- 
nied by  a  fall  in  ATP  with  an  elevation  of  ADP.  This  stimulatory  phase 
lasts  less  than  a  minute  and  is  followed  by  a  depressed  phase  characterized 


396  2.  ANALOGS  or  enzyme  reaction  components 

by  very  low  levels  of  P^.  During  the  oxidation  of  ethanol,  the  ATP/ADP 
and  P,/ADP  ratios  are  high;  when  2-DG  is  added  it  temporarily  augments 
respiration  by  lowering  these  ratios,  but  within  a  minute  enough  phosphate 
has  been  trapped  in  2-DG-6-P  to  cause  a  marked  fall  in  the  Pj/ADP  ratio 
(perhaps  from  17  to  1.5).  The  P,  may  now  be  so  low  that  it  limits  the  respi- 
ration. 

The  respiration  of  certain  tissues  is  diminished  by  the  addition  of  glucose, 
a  phenomenon  often  called  the  Crabtree,  or  reversed  Pasteur,  effect;  it 
is  particularly  evident  in  Ehrlich  ascites  carcinoma  cells  and  most  of 
the  studies  of  the  mechanisms  involved  have  been  on  these  cells.  It  has 
been  stated  that  an  acceleration  of  glycolysis  inhibits  the  oxidation  of 
pyruvate,  but  there  was  no  real  evidence  to  link  the  entire  EM  pathway 
with  respiratory  control.  The  effects  of  2-DG  are  thus  of  particular  impor- 
tance, since  it  is  phosphorylated  but  not  further  metabolized  to  any  extent. 
It  was  shown  that  2-DG  inhibits  respiration  to  about  the  same  degree  as 
glucose  (Ibsen  et  al.,  1958).  Respiration  and  pyruvate  decarboxylation  are 
reduced  50%  by  10  milf  2-DG  (Ram  et  al,  1963).  There  has  been  dis- 
agreement as  to  whether  glucose  and  2-DG  act  by  the  same  mechanism  or 
differently.  Let  us  briefly  compare  the  responses  to  these  sugars.  (1)  2-DG 
depresses  the  respiration  more  slowly  than  does  glucose.  Yushok  (1964) 
has  shown  in  a  group  of  sugars  that  the  rate  of  respiratory  inhibition  is 
correlated  with  the  rate  of  phosphorylation.  One  would  thus  expect  2-DG 
to  act  more  slowly  than  glucose,  so  this  does  not  constitute  a  real  difference 
in  action.  (2)  The  inhibition  by  glucose  is  released  when  it  is  all  glycolyzed 
but  the  inhibition  by  2-DG  remains  (Ibsen  et  al.,  1962;  Hofmann  et  al.,  1962). 
This  does  not  seem  to  me  to  be  valid  evidence  for  different  mechanism  of 
action.  (3)  The  addition  of  glucose  leads  to  the  formation  of  lactate  whereas 
2-DG  does  not  (Ram  et  al.,  1963).  This  is  what  would  be  expected,  of  course, 
but  emphasizes  that  glycolysis,  as  defined  classically,  is  not  necessary  for  the 
effect.  (4)  Glucose  at  10  mM  inhibits  the  respiration  40%,  2-DG  at  20  mM 
inhibits  it  48%,  and  both  together  inhibit  it  only  23%  (Wenner  and  Cereijo- 
Santalo,  1962).  This  was  interpreted  to  mean  that  the  inhibitory  mechanisms 
are  quite  different.  (5)  It  has  been  stated  that  amobarbital  prevents  the 
inhibition  of  respiration  by  2-DG  but  not  by  glucose  (Wenner  and  Cereijo- 
Santalo,  1962).  This  is  true,  however,  only  in  the  presence  of  succinate, 
since  the  endogenous  respiration  is  not  affected  by  either  glucose  or  2-DG 
in  the  presence  of  amobarbital  (there  is  very  little  to  be  affected).  (6)  The 
respiratory  inhibition  by  glucose  is  released  by  2,4-dinitrophenol,  but 
there  is  some  disagreement  as  to  the  effect  of  the  uncoupler  with  2-DG, 
Ibsen  et  al.  (1962)  stating  that  it  releases  the  inhibition  and  Ram  et  al. 
(1963)  stating  that  it  does  not.  The  latter  workers,  however,  did  not  feel 
that  this  is  evidence  for  different  mechanisms  and  were  inclined  to  attrib- 
ute the  differences  to  the  availability  of  glycolytic  intermediates.  (7)  Both 


EFFECTS    OF    2-DEOXY-D-GLUCOSE  397 

glucose  and  2-DG  reduce  the  ATP  level  immediately  and  the  ADP  level 
very  soon  (Ibsen  et  al.,  1962).  The  most  commonly  accepted  explanation 
of  the  Crabtree  effect  is  a  depletion  of  ADP,  since  the  rate  of  mitochondrial 
oxidation  in  ascites  cells  depends  on  the  level  of  ADP.  (8)  2-DG  causes  the 
loss  of  enzymes  from  the  ascites  cells  (measured  with  lactate  dehydrogen- 
ase) whereas  glucose  does  not  (Hofmann  et  al.,  1962).  Indeed,  glucose 
seems  to  antagonize  this  action  of  2-DG.  Whether  this  observation  has  any 
bearing  on  the  Crabtree  effect  is  not  known.  (9)  Both  glucose  and  2-DG 
still  exert  respiratory  inhibition  in  the  presence  of  sufficient  iodoacetate 
to  block  almost  completely  the  glycolytic  pathway  (Ibsen  et  al.,  1958; 
Wenner  and  Cereijo-Santalo,  1962).  It  seems  that  although  the  Crabtree 
effect  is  not  abolished  by  iodoacetate,  it  may  be  diminished.  If  the  initial 
phosphorylation  of  hexoses  is  responsible  for  the  Crabtree  effect,  iodoacetate 
would  not  be  expected  to  inhibit  it,  except  as  it  might  reduce  ATP  for  the 
kinase  reactions.  (10)  Glucosone  inhibits  the  Crabtree  effect  produced  by 
both  glucose  and  2-DG  (Yushok,  1964),  but  this  is  probably  the  result  of 
the  inhibition  of  hexokinase  by  glucosone.  Summarizing  these  results,  it 
would  appear  that  glucose  and  2-DG  inhibit  respiration  by  basically  the 
same  mechanism  and  that  this  is  related  to  their  phosphorylation.  It  is 
difficult  to  understand  the  diminished  Crabtree  effect  in  the  presence  of 
glucose  and  2-DG  together,  observed  by  Wenner  and  Cereijo-Santalo  (1962), 
but  this  should  be  investigated  further,  inasmuch  as  Ram  et  al.  (1963) 
stated  that  the  respiratory  inhibition  by  glucose  is  not  enhanced  by  2-DG, 
apparently  no  antagonism  being  noted. 

Glucose  and  2-DG  not  only  depress  the  oxidation  of  pyruvate  in  ascites 
cells  but  even  more  strongly  the  oxidation  and  C^'^Oa  formation  from  label- 
ed palmitate  (Sauermann,  1964).  Inhibition  of  palmitate  oxidation  occurs 
to  the  extent  of  around  80%  at  the  relatively  low  concentration  of  1.8 
mM  2-DG.  The  inhibitions  are  approximately  29%  for  acetate,  58%  for 
pyruvate,  and  92%  for  palmitate  at  18  mM  2-DG.  There  would  thus  appear 
to  be  some  effect  on  fatty  acid  oxidation  which  is  exerted  prior  to  the  uti- 
lization of  acetyl-CoA.  This  action  is  not  related  to  the  inhibition  of  glyco- 
lysis for  several  reasons,  including  the  demonstration  that  it  occurs  in  the 
presence  of  iodoacetate.  Some  effect  on  fatty  acid  oxidation  might  be 
expected  from  a  lowering  of  the  ATP  level,  but  it  was  felt  that  this  is  not 
the  entire  explanation  because  of  the  relatively  small  effect  on  acetate 
oxidation.  It  is  not  necessary,  however,  that  a  fall  in  ATP  should  affect 
acetate  and  palmitate  oxidations  equally.  One  awaits  with  interest  the 
elucidation  of  this  interesting  effect. 

The  Crabtree  effect  has  been  discussed  in  some  detail  because  it  illus- 
trates one  way  in  which  2-DG  can  alter  carbohydrate  metabolism  through 
the  alteration  of  levels  of  P,  and  the  adenine  nucleotides.  It  is  quite  possi- 
ble that  part  of  the  effect  of  2-DG  on  tissues  is  the  result  of  a  lowering 


398 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


of  available  ATP,  this  secondarily  reducing  hexokinase  activity,  and  the 
uptake  of  sugars.  The  greater  resistance  of  aerobic  glycolysis  and  respira- 
tion to  2-DG,  compared  to  anaerobic  glycolysis,  may  be  due  in  part  to  the 
higher  ATP  levels  aerobically.  The  differential  effects  on  the  uptakes  of 
different  hexoses  (Fig.  2-13)  might  be  explained  on  the  basis  of  whether  a 
hexose  is  actively  transported  or  not,  and  the  steady-state  rate  of  its 
phosphorylation  (e.  g.,  galactose  metabolism  is  not  depressed  as  much  as 
that  of  glucose),  and  the  different  affinities  of  the  hexoses  for  the  hexoki- 
nases. 


20 


15 


10 


5  ■ 


c  0^ 
(%  OF 
ADDED  ) 


^ 

--^ 

■"-..^GLUCOSE 

\ 

FRUCTOSE 

GALACTOSE 

^ 

-^-^^__ 

10 


15 


20 


[  2  -  DG  ) 


Fig.   2-13.    Inhibition   of  hexose   oxidation   by   2-DG    in   isolated  rat 
diaphragm.   Hexoses  were  10  mM.  (From  Nakada   and   Wick,    1956.) 


A  summary  of  the  sites  and  mechanisms  in  the  inhibition  of  carbohydrate 
utilization  by  2-DG  would  then  include:  (1)  primary  competitive  inhibition 
of  certain  hexokinases  by  2-DG,  (2)  possible  direct  interference  in  the  ac- 
tive transport  of  hexoses  into  the  cell,  (3)  inhibition  of  phosphoglucose 
isomerase  by  2-DG-6-P,  (4)  secondary  reduction  in  transport  and  hexose 
phosphorylation  through  depletion  of  ATP,  and  (5)  possible  inhibitions  by 
2-DG-6-P  of  glycolytic  enzymes  not  yet  examined. 


EFFECTS    OF    2-DEOXY-D-GLUCOSE  399 

Effects  of  2-DG   on   Various   Metabolic   Pathways 

Infusion  of  acetate-C^^OO"  into  rabbits  and  determination  of  the  respired 
C^^Og  were  done  prior  to  and  after  2-DG  injections;  no  alteration  of  acetate 
oxidation  was  observed  (Wick  d  al.,  1957).  The  total  COg  produced  decreas- 
ed, partly  due  to  the  hypothermia  brought  about  by  2-DG.  This  was  taken 
as  evidence  that  any  block  by  2-DG  is  previous  to  the  tricarboxylate  cycle. 
An  effect  of  2-DG  on  citrate  levels  in  ascites  carcinoma  cells  in  the  presence 
of  pyruvate  was  reported  by  Letnansky  and  Seelich  (1960).  The  citrate 
begins  to  rise  2  min  after  addition  of  18.3  mM  2-DG  and  eventually  reaches 
levels  definitely  higher  than  in  untreated  suspensions.  The  utilization  of 
pyruvate  is  also  depressed  around  50%.  These  observations  might  be  in- 
terpreted as  originating  from  some  action  on  the  cycle,  but  more  recently 
(Seelich  and  Letnansky,  1961)  it  was  shown  that  methylene  blue  reduces 
the  high  citrate  levels  and  promotes  pyruvate  utilization  in  the  presence  of 
2-DG.  It  was  postulated  that  this  is  caused  by  oxidation  of  NADPH  to 
NADP,  which  is  necessary  for  isocitrate  oxidation.  The  rise  in  citrate  may 
be  due  to  a  deficiency  of  NADP,  possibly  because  lactate  is  not  formed  in 
the  presence  of  2-DG  and  NADPH  is  not  oxidized,  and  also  due  to  a  drop  in 
ATP.  The  results  can  thus  be  satisfactorily  explained  on  the  basis  of  the 
mechanisms  previously  discussed  without  assuming  any  direct  action  on 
the  cycle. 

Lipogenesis  in  human  fetal  liver  from  glucose  is  depressed  by  2-DG, 
incorporation  of  l-C^*  being  lowered  33%  and  of  6-Ci*  18%  at  6.1  mM 
(Villee  and  Loring,  1961).  This  can,  of  course,  be  attributed  to  an  inhi- 
bition of  glycolysis.  Plasma  fatty  acids  in  man  rise  as  much  as  330% 
following  intravenous  infusion  of  2-DG  at  60  mg/kg,  and  this  might  be 
interpreted  as  a  depression  of  fatty  acid  synthesis  so  that  the  normal  equi- 
librium is  disturbed  and  fatty  acids  are  mobilized  from  the  tissues  (Laszlo 
et  al.,  1961).  On  the  other  hand,  there  is  evidence  that  2-DG  augments 
epinephrine  release,  so  that  this  could  be  partly  responsible. 

The  appearance  of  labeled  amino  acids  from  labeled  glucose  in  brain 
slices  is  also  depressed  by  2-DG  (Tower,  1958).  2-DG  not  only  prevents  the 
accumulation  of  glutamate  but  causes  a  profound  fall  in  the  intracellular 
level,  glutamine  decreasing  less  markedly.  It  is  believed  that  glutamate 
accumulation  is  important  in  K+  uptake  by  brain  and  it  was  found,  as 
expected,  that  this  is  inhibited  around  60%  by  10  mM  2-DG.  The  role  of 
such  changes  in  the  central  actions  of  2-DG  in  the  whole  animal  is  not  as 
yet  understood.  The  incorporation  of  L-valine-C^*  into  protein  in  ascites 
cells  is  almost  abolished  by  10  mM  2-DG;  since  this  may  be  reversed  by 
glucose,  it  is  likely  that  the  inhibition  is  related  to  glycolytic  suppression 
(Riggs  and  Walker,  1963).  The  DNA  content  of  a  rat  carcinoma  is  reduced 
by  feeding  2-5%  2-DG  in  the  diet  and  possibly  this  is  related  to  the  observed 
inhibition  in  growth  (Sokoloff  et  al.,  1956). 


400  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Effects  of  2-DG   on   Cell    Division   and    Growth 

The  growth  rate  of  E.  coli  is  depressed  at  least  50%  by  10  mM  2-DG 
but  the  differential  rate  of  inducible  /5-galactosidase  synthesis  is  not  altered 
(Cohn  and  Horibata,  1959).  Although  Neurospora  and  Aspergillus  can 
grow  slowly  on  2-DG  alone,  the  growth  is  inhibited  when  the  usual  sugars 
are  present  (Sols  et  al.,  1960  b).  The  mutiplication  of  influenza  virus  in 
chick  embryos  is  markedly  inhibited  by  2-DG  but  in  preliminary  studies 
there  was  no  evidence  that  the  development  of  lesions  in  mice  is  altered 
(Kilbourne,  1959).  Glucose  and,  to  some  extent,  pyruvate  are  able  to 
counteract  this  depression.  Sea-urchin  egg  cleavage  is  delayed  by  2-DG 
and  development  is  stopped  at  various  stages,  depending  on  the  concen- 
tration: 1-10  mM  prevents  gastrulation  and  the  eggs  reach  swimming 
blastulae,  100  mM  delays  first  cleavage  but  the  early  blastula  stage  is 
eventually  reached,  200  mM  causes  greater  cleavage  delay  and  develop- 
ment stops  before  the  blastula  stage  (Bernstein  and  Black,  1959).  Glucose 
can  counteract  the  cleavage  delay.  The  most  sensitive  tissue  investigated 
appears  to  be  chick  embryo  heart  fibroblasts,  since  the  mitotic  index  is 
reduced  from  2.65  to  0.15  by  1.52  mM  2-DG  (Ely  et  al,  1952).  Glucose  at 
approximately  equimolar  concentration  is  not  able  to  reverse  this  inhibition 
significantly.  The  migration  of  cells  in  the  2-DG-treated  cultures  is  also 
limited  and  the  cells  become  vacuolated.  These  observations  all  point  to  a 
general  growth-inhibiting  action  of  2-DG,  which  is  not  unexpected,  but 
there  has  been  little  study  of  differential  growth  effects. 

The  ability  of  tumors  to  derive  a  good  part  of  their  energy  for  growth 
from  glycolysis  prompted  the  original  study  of  2-DG  as  a  possible  carcino- 
static  agent,  but  surprisingly  little  work  has  been  done  on  this  aspect. 
It  has  been  established  that  glycolytic  inhibition  of  tumors  in  vivo  is  pos- 
sible, in  that  patients  with  chronic  myelogenous  leukemia  infused  with 
60  mg/kg  2-DG  show  35-40%  inhibition  of  glycolysis  in  the  leukemic  cells 
(Laszlo  et  al.,  1958).  The  growth  of  cultured  HeLa  cells  from  human  carci- 
noma is  inhibited  readily  by  2-DG,  5  mM  producing  essentially  complete 
depression  which  is  reversible  up  to  3  days  but  not  afterward  (Barban 
and  Schulze,  1961).  Glucose  or  mannose  will  counteract  the  growth  inhibi- 
tion. Cells  in  a  fructose  medium  are  inhibited  more  readily  than  when 
grown  on  glucose  or  mannose,  which  corresponds  to  the  greater  inhibition 
of  fructose  utilization  discussed  previously.  When  2-DG  is  added  at  2-5% 
to  the  diet  of  rats,  the  growth  of  carcinoma  G-175  in  reduced  (Sokoloff  et  al., 
1956).  Adult  rats  tolerate  this  dosage  well  but  the  growth  of  young  rats  is 
retarded.  The  2-DG  can  also  be  injected  subcutaneously  at  2-4  mg/kg/day. 
Mouse  sarcoma  is  similarly  affected.  The  inhibition  in  either  case  is  not 
very  marked.  Solid,  transplanted,  and  ascitic  tumors  in  mice  grow  more 
slowly  when  2-DG  is  administered  (Laszlo  et  al.,  1960).  A  modest  prolonga- 
tion of  the  survival  time  of  the  animals  was  noted.  Definite  carcinostatic 


EFFECTS    OF    2-DEOXY-D-GLUCOSE  401 

activity  in  vivo  has  thus  been  demonstrated  but  there  should  be  much  more 
work  to  establish  if  sufficient  differential  depression  can  be  achieved,  and 
other  types  of  neoplasm  should  be  studied. 

Effects   of  2-DG   on    Whole   Animals 

The  intravenous  infusion  of  2-DG  at  doses  of  50-200  mg/kg  in  cancer 
patients  produces  a  feeling  of  warmth,  flushing,  diaphoresis,  headache, 
drowsiness,  tachycardia,  a  rise  in  blood  glucose,  and  a  fall  in  white  cell 
count  (Landau  et  al.,  1958).  Hyperglycemia  has  been  noted  in  all  studies 
and  might  be  attributed  to  a  reduced  utilization  of  glucose  brought  about 
by  glycolytic  inhibition.  However,  other  factors  must  be  considered.  Brown 
and  Bachrach  (1959)  showed  that  the  rise  in  blood  glucose  from  2-DG  can 
be  partially  prevented  by  demedullation  of  the  adrenals,  indicating  that 
2-DG  may  stimulate  the  release  of  epinephrine.  Increases  in  urinary  cate- 
cholamines during  2-DG  infusion  have  also  been  noted  (Laszlo  et  al.,  1961). 
Hokfelt  and  Bydgeman  (1961)  felt  that  this  epinephrine  release  is  the 
primary  cause  of  the  hyperglycemia  and  showed  that  2-DG  can  deplete 
the  adrenals  of  half  their  epinephrine.  Spinal  transection  reduces  the  2-DG 
effect,  indicating  epinephrine  release  to  be  mediated  through  the  central 
nervous  system.  Pretreatment  with  dihydroergotamine,  which  blocks  the 
effects  of  epinephrine,  abolishes  the  2-DG  hyperglycemia  (Altszuler  et  al., 
1963;  Sakata  et  al.,  1963).  It  is  interesting  in  connection  with  the  possible 
effects  of  2-DG  on  the  central  nervous  system  that  2-DG  is  transported  from 
the  blood  into  the  cerebrospinal  fluid  faster  than  glucose;  there  is  also  com- 
petition between  glucose  and  2-DG  for  the  carrier  (Fishman,  1964).  Landau 
et  al.  (1958),  on  the  other  hand,  were  inclined  to  discount  the  role  of  epine- 
phrine since  no  rise  in  blood  pressure  is  observed  during  maximal  hyper- 
glycemia. The  hyperglycemia  probably  is  responsible  for  the  decreased 
stainability  and  degranulation  of  pancreatic  islet  cells  produced  by  2-DG, 
indicating  increased  activity  of  these  cells,  rather  than  a  direct  action 
(Hokfelt  and  Hultquist,  1961).  The  symptoms  listed  above  must  be  due 
in  part  to  the  restricted  glucose  utilization  caused  by  2-DG.  The  rise  in 
blood  glucose  must  tend  to  counteract  this  inhibition,  since  glucose  in- 
fusions reduce  mortality  from  2-DG,  but  not  sufficiently.  It  is  interesting 
that  the  anaphylactoid  reaction  to  dextran  and  ovomucoid  in  rats  is 
inhibited  by  2-DG  at  200  mg/kg  intravenously,  although  the  response  to  the 
histamine  releasers  is  not  affected,  indicating  an  important  role  of  glucose 
uptake  and  metabolism  in  certain  inflammatory  reactions  (Goth,  1959). 
The  LD50  in  mice  is  around  2.5  g/kg  for  single  intravenous  injections  and 
5  g/kg  subcutaneously  or  intraperitoneally  (Laszlo  et  al.,  1960).  The  animals 
survive  several  hours  to  a  day  in  most  cases  but  may  die  within  10  min  af- 
ter intravenous  injection.  The  major  toxic  effects  are  related  to  the  central 
nervous  system  and  are  weakness,  convulsions,  and  coma. 


402  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Effects  of  2-DG  on  the  Heart  * 

The  contractile  tension  of  rat  atria  is  progressively  depressed  by  10  mM 
2-DG  when  glucose  is  present  at  its  usual  concentration  of  5.5  mM;  the  inhi- 
bition is  25%  at  20  min,  50%  at  40  min,  and  70%  at  90  min  (A.  Gimeno  and 
M.  Gimeno).  If  pyruvate  is  present,  the  rate  of  depression  is  slower  and  the 
inhibition  is  30%  at  90  min;  if  pyruvate  is  added  at  30  min  when  the  depres- 
sion by  2-DG  is  around  40-45%,  there  is  partial  recovery  to  the  —  30% 
level;  if  glucose  and  pyruvate  are  present  and  2-DG  is  added  at  30  min, 
there  is  a  depression  to  the  —  30%  level  at  90  min.  The  atria  in  the  absence 
of  glucose  progressively  fail  so  that  at  30  min  the  contractile  tension  is 
50%  depressed  (actually  the  course  is  quite  similar  to  that  when  glucose 
and  2-DG  are  present),  but  with  2-DG  the  rate  of  fall  is  much  faster,  in- 
dicating that  2-DG  can  effect  the  endogenous  metabolism.  Addition  of 
glucose  at  30  min  when  the  depression  is  80%  or  more  results  in  partial 
recovery,  pyruvate  is  less  effective,  and  both  allow  return  to  near  the 
—  30%  level.  The  rapid  cessation  of  the  2-DG  depression  brought  about 
by  either  glucose  or  pyruvate  is  noteworthy.  The  contractile  levels  reached 
in  90  min  may  be  summarized  in  the  following  tabulation.  The  failure  of 


(I)  Glucose-free  and  glucose  added  at  30  min  —  0% 

(II)  Glucose-free  and  pyruvate  added  at  30  min  —15% 

(III)  Glucose  +  pyruvate  +  2-DG  (in  any  order)  —30% 
(IV^  Glucose  +  2-DG  (in  any  order)  -70% 

(V)  2-DG  alone  -90%  to  -100% 


pyruvate  to  maintain  normal  contractions  might  indicate  that  some  25-30% 
of  the  tension  is  dependent  on  glycolysis,  but  it  is  also  possible  that  2-DG  is 
interfering  in  some  way  with  the  utilization  of  pyruvate.  The  ability  of 
glucose  to  stimulate  30  min  after  depression  by  2-DG  at  10  mM  shows 
that  the  block  of  glycolysis  is  only  partial  or  that  glucose  is  acting  by  some 
other  mechanism.  The  atrial  depression  by  anoxia  (—  85%  at  10  min)  is 
accelerated  by  2-DG  (—  93%  at  5  min  and  —  100%  at  10  min),  and  re- 
covery upon  readmission  of  Og  is  much  less  when  2-DG  is  present  (J.  La- 
cuara).  Rat  ventricle  strip  contractility  is  not  affected  over  160  min  by 
0.5-2  milf  2-DG,  but  4  mM  causes  a  slow  depression  and  partially  prevents 
the  positive  inotropic  effect  of  ouabain  (E.  Majeski). 

*  Inasmuch  as  so  little  is  known  of  the  effects  of  2-DG  on  tissue  functions  and 
nothing  has  been  reported  relative  to  the  heart,  this  section  summarizes  briefly  some 
of  the  recent  work  done  in  our  department  and  not  yet  published  at  the  time  of  sub- 
mission of  the  manuscript. 


EFFECTS    OF    6-DEOXY-6-FLUORO-D-GLUCOSE  403 

It  is  very  interesting  that  the  atrial  depression  is  completely  unassoci- 
ated  with  demonstrable  changes  in  the  membrane  electrical  characteristics 
(E.  Ruiz-Petrich).  2-DG  at  10  mM  depresses  the  contractile  tension  around 
50%  at  20-30  min  under  the  conditions  of  these  experiments,  glucose  being 
present,  but  there  are  no  changes  at  all  of  the  resting  potential,  the  action 
potential  magnitude,  or  the  rates  of  depolarization  and  repolarization.  The 
addition  of  5  mM  pyruvate  rapidly  stopped  the  progression  of  the  contractile 
depression  and  allowed  slight  recovery,  again  without  detectable  altera- 
tions of  the  membrane  characteristics.  2-DG  is  the  only  inhibitor  with  which 
we  have  worked  that  is  able  to  affect  the  contractile  processes  so  selectively, 
all  other  inhibitors  decreasing  the  action  potential  duration  to  varying 
degrees  and  producing  other  correlated  changes  in  the  potentials.  If  2-DG 
acts  here  by  reducing  the  utilization  of  glucose  or  glycogen,  these  results 
would  point  to  a  close  relation  between  the  contractile  process  and  some  as- 
pect of  glycolysis  other  than  the  generation  of  ATP.  It  has  also  been  shown 
that  atrial  K+  influx  and  efflux  are  only  very  slightly  altered  by  11  mM 
2-DG,  and  that  intracellular  K+  is  unchanged  over  a  period  during  which 
the  contractile  activity  is  depressed  50%  (Chin,  1963). 

EFFECTS   OF  6-DEOXY-6-FLUORO-D-GLUCOSE 
ON    METABOLISM 

Modification  of  hexoses  at  the  6-position  should  interfere  with  their 
phosphorylation  by  hexokinase,  and  hence  any  inhibition  of  glucose  metab- 
olism observed  would  probably  not  be  exerted  beyond  the  hexokinase  step. 
Thus  inhibition  produced  by  6-substituted  sugars  should  be  simpler  than 
inhibition  by  2-substituted  sugars,  which  can  be  phosphorylated  and  may 
block  at  several  sites.  Brooks  et  al.  (1961)  compared  6-deoxy-D-glucose 
with  2-DG  on  the  oxidation  of  glucose-u-C^*  by  various  tissues  and  found 
it  to  be  somewhat  less  potent.  When  glucose  is  10  mM  and  6-DG  is  30  mM, 
the  inhibitions  of  C^'^Og  formation  are  43%  for  adipose  tissue,  23%  for 
kidney,  and  19%  for  diaphragm.  The  inhibition  in  adipose  tissue  is  reversed 
by  increasing  glucose  concentration  and  appears  to  be  competitive.  6-DG 
is  not  metabolized  by  adipose  tissue  and  no  C^^Oa  arises  from  6-DG-u-C^^. 
The  site  of  inhibition  was  considered  to  be  either  the  membrane  transport 
system  or  hexokinase.  It  has  been  found  that  galactose  transport  across 
the  intestine  is  inhibited  by  6-DG  (47%  at  5  mM)  and  that  the  transport  of 
6-DG  is  depressed  by  glucose  (Wilson  et  al.,  1960). 

The  replacement  of  the  6-0II  group  with  fluorine  to  give  6-deoxy-6- 
fluoro-D-glucose  (6-DFG)  leads,  as  expected,  to  an  interesting  inhibitor 
of  glucose  utilization.  The  original  idea  was  apparently  to  produce  a  gly- 
colytic inhibitor  with  fluorine  analogous  to  the  cycle  inhibitor  fluoroacetate. 
The  initial  work  was  done  by  Blakley  and  Boyer  (1955)  at  Minnesota;  they 


404  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

showed  that  fermentation  of  glucose  and  fructose  by  yeast  is  competitively 
inhibited  by  6-DFG.  The  apparent  constants  are  shown  in  the  following 
tabulation.  6-DFG  is  not  fermented  by  intact  cells  or  yeast  extracts.  In 


Glucose 

Fructose 

Yeast 

Km 

(milf) 

Ki 

{mM) 

Km 

{n\M) 

K, 

(mM) 

Baker's 
Brewer's 

1.8 
6.9 

7.3 

2.7 

5.0 

27 

3.3 

2.3 

extracts  6-DFG  has  essentially  no  effect  even  at  36  milf  and  hexokinase 
is  very  weakly  inhibited.  Glucose  utilization  by  rat  diaphragm  is  not  in- 
hibited as  well  as  yeast  fermentation,  and  6-DFG  is  not  taken  up  as  readily 
as  glucose  by  the  muscle  cells.  Glucose  oxidase  (notatin)  oxidizes  6-DFG  at 
about  3%  the  rate  of  glucose  oxidation  so  that  direct  oxidation  in  the 
tissues  is  probably  negligible.  It  was  concluded  that  the  rate-limiting  re- 
action for  glucose  utilization  is  different  in  intact  cells  and  extracts,  and 
that  6-DFG  probably  inhibits  the  membrane  transport  of  normal  hexoses. 

The  uptake  and  metabolism  of  6-DFG  may  vary  from  tissue  to  tissue. 
For  example,  although  6-DFG  is  not  actively  transported  into  rat  diaphragm 
and  insulin  has  no  effect  on  this  transfer,  it  is  well  transported  across  the 
intestinal  wall  whereas  2-DG  is  not  (Wick  et  at.,  1959;  Wilson  and  Landau, 
1960).  And,  although  6-DFG  is  not  metabolized  in  most  tissues,  a  particulate 
preparation  from  Aerobacter  aerogenes  is  able  to  oxidize  it  to  the  corres- 
ponding gluconate  (Blakley  and  Ciferri,  1961).  It  is  fairly  certain  that  the 
compound  is  quite  stable  and  that  release  of  fluoride  does  not  occur  suffi- 
ciently to  inhibit  glycolysis  (Serif  et  al,  1958). 

6-DFG  inhibits  the  formation  of  C^^Oa  from  glucose-u-C^^  in  liver,  kidney, 
and  adipose  tissue,  without  modifying  the  metabolism  of  acetate  or  lactate 
(Serif  and  Stewart,  1958;  Serif  and  Wick,  1958;  Serii  et  al,  1958).  In  general 
6-DFG  is  somewhat  less  potent  than  2-DG  (Fig.  2-11),  but  the  relative 
sensitivities  depend  on  the  tissue  studied.  Nevertheless,  in  the  eviscerated 
rat  6-DFG  is  able  to  inhibit  the  intracellular  transport  of  glucose  quite 
appreciably,  e.  g.,  42%  from  200  mg/kg  (Wick  et  al,  1959).  There  is  no  direct 
evidence  that  6-DFG  acts  elsewhere  than  on  membrane  transport  and  the 
comparable  inhibitions  produced  on  the  metabolism  of  glucose- 1-C^^,  glu- 
cose-6-C^*,  and  glucose-u-C^*  would  point  to  a  block  prior  to  the  glyco- 
lytic-pentose phosphate  shunt  division  (Serif  and  Wick,  1958).  However, 
hexokinases  from  other  than  yeast  have  not  been  adequately  tested.  Al- 
though the  toxicities  of  2-DG  and  6-DFG  have  not  been  directly  compared, 
Blakley  and  Boyer  (1955)  found  that  250  mg/kg  6-DFG  intraperitoneally 


INHIBITORS    OF    CARBOHYDRATE    METABOLISM  405 

in  rats  produces  toxic  symptoms  with  recovery.  Since  the  LD50  for  2-DG 
is  probably  around  10  times  this,  it  would  indicate  that  6-DFG  must  af- 
fect the  central  nervous  system  more  than  the  other  tissues  which  have 
been  studied. 

VARIOUS  ANALOG   INHIBITORS 
OF  CARBOHYDRATE    METABOLISM 

Numerous  inhibitions  of  enzymes  in  the  glycolytic  and  pentose  phosphate 
pathways  by  analogs  have  been  reported.  Some  of  these  may  be  important 
in  attempts  to  block  carbohydrate  metabolism  specifically  and  some  are 
undoubtedly  significant  in  mechanisms  regulating  the  rates  in  these  path- 
ways. We  shall  discuss  a  few  of  the  more  important  enzymes  and  inhibi- 
tions, no  effort  being  made  to  include  all  of  the  observations. 

Phosphorylases 

The  enzymes  involved  in  the  synthesis  and  phosphorolysis  of  polysac- 
charides, such  as  starch  or  glycogen,  are  very  stereospecific  with  respect 
to  substrates  and  inhibitors.  The  fairly  potent  inhibition  of  glycogen  phos- 
phorolysis by  the  product  glucose- 1-P  in  preparations  from  rabbit  muscle 
was  reported  by  Cori  et  al.  (1939),  7  mM  inhibiting  93%,  whereas  glucose- 
6-P  at  the  same  concentration  inhibits  only  17%.  The  reverse  reaction  of 
glycogen  synthesis  from  glucose-1-P  is  inhibited  competitively  by  glucose, 
and  to  a  lesser  extent  by  mannose,  galactose,  and  maltose,  but  the  ajffinity 
of  the  enzyme  for  these  sugars  is  low  since  30%  inhibition  occurs  when 
glucose  and  substrate  are  approximately  equimolar  at  17  m.M  (Cori  and 
Cori,  1940).  However,  the  lobster  muscle  phosphorylase  is  inhibited  only 
25%  by  250  mM  glucose  when  glucose-1-P  is  100  mM  (Cowgill,  1959). 

The  only  known  substrate  for  rabbit  muscle  phosphorylase  is  a-D-glu- 
cose-l-P  with  respect  to  glycogen  synthesis,  and  Cori  and  Cori  (1940)  had 
found  that  only  a-D-glucose  inhibits,  /?-D-glucose  having  little  if  any  ef- 
fect, so  the  question  of  a,  /5-specificity  was  studied  in  detail  on  the  crystal- 
line enzyme  by  Campbell  et  al.  (1952).  It  was  found  that  /?-D-glucose-l-P 
is  neither  a  substrate  nor  an  inhibitor,  and  that  whereas  a-methyl  glucoside 
inhibits,  /?-metliylglucoside  is  without  activity.  The  /?-anomers  appear  to 
have  no  affinity  for  the  enzyme.  Furthermore,  a  large  number  of  sugars 
and  derivatives  at  50  mM  were  found  to  be  inactive.  The  configurations  of 
the  hydroxyl  groups  on  all  the  positions  of  glucose  seem  to  be  necessary 
for  combination  with  the  enzyme.  A  pyranose  structure  seems  also  to  be 
a  requirement  for  inhibition,  since  sorbitol  and  inositol  are  without  effect, 
and  a  primary  alcohol  group  on  C-5  is  necessary  since  D-xylose  is  inactive. 
The  failure  of  fructose  to  inhibit  might  be  explained  in  several  ways:  (1) 
the  more  planar  furanose  form  is  dominant,  (2)  the  configurations  on  the 


406  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

C-1  of  D-glucopyranose  and  C-2  of  D-fructopyranose  are  different,  (3) 
there  is  a  primary  alcohol  group  on  the  C-5  of  glucose  but  not  on  the  C-6 
of  fructose,  and  (4)  the  /?-anomer  of  fructose  may  predominate. 

An  interesting  type  of  inhibition  on  the  phosphorolysis  of  starch  by 
monofluorophosphate  was  observed  by  Rapp  and   Sliwinski  (1956).  The 

0-  o- 

-O— P+— OH  -0— P+— F 

o-  o- 

Orthophosphate  Monofluorophosphate 

sizes  and  electronic  configurations  of  orthophosphate  and  monofluoro- 
phosphate are  quite  similar  so  that  interference  with  many  reactions  in- 
volving phosphate  might  be  anticipated.  The  inhibition  of  potato  phosphory- 
lase  is  completely  competitive  {Kj„  =  3.2  mM  and  K,  =  2.8  mM  calculated 
from  their  ijv-ijiS)  plot)  and  the  affinity  of  the  enzyme  for  the  two  substances 
corresponds  to  this  tectonic  resemblance.  The  inhibition  is  not  due  to  the 
release  of  fluoride  since  20  mM  fluoride  inhibits  only  6.4%  and  2.1  mM 
monofluorophosphate  inhibits  50%. 

Phosphoglucose  Isomerase 

This  enzyme  is  important  in  regulating  carbohydrate  metabolism  since 
its  activity,  along  with  other  factors,  determines  how  much  glucose-6-P 
enters  the  glycolytic  pathway;  in  other  words,  this  enzyme  represents  a 
branching  point  of  metabolism  in  the  terminology  of  Krebs.  We  have  seen 
that  inhibition  by  2-DG-6-P  is  probably  an  important  component  of  the 
mechanism  of  action  of  2-DG.  The  potent  inhibition  by  6-phosphogluconate 
is  particularly  interesting  because  this  substance  is  formed  from  glucose-6-P 
in  the  pentose  phosphate  pathway  and  could  determine  to  some  extent  the 
diversion  at  the  branching  point.  Parr  (1956, 1957)  reported  inhibition  of  the 
enzymes  from  blood,  liver,  muscle,  and  potato  and  found  it  to  be  competi- 
tive; in  the  reaction  glucose-6-P  ->  fructose-6-P,  1  mM  inhibits  75%  and 

fructose-6-P »^  glycolytic  pathway 

Glucose  »-  glucose-6-P 


-phosphogluconate ■*-  pentose- P  pathway 

2  mM  95%  (glucose-6-P  2  mM).  Similar  competitive  inhibitions  of  the  en- 
zymes from  yeast  (Noltmann  and  Bruns,  1959)  and  Trichinella  spiralis 
(Mancilla  and  Agosin,  1960)  have  been  noted.  Rabbit  muscle  phosphoglu- 
cose isomerase  may  be  even  more  sensitive  to  6-phosphogluconate,  since  the 


INHIBITORS    OF    CARBOHYDRATE    METABOLISM  407 

Ki  is  around  0.005  mM  (Kahana  et  al.,  1960).  Another  potent  inhibitor  that 
is  an  intermediate  in  the  pentose-P  pathway  is  erythrose-4-P,  for  which 
the  K^  is  0.0007-  0.001  mM  {K„,  for  fructose-6-P  is  0.08  mM)  (Grazi  et  al, 
1960).  Under  conditions  that  would  lead  to  an  accumulation  of  erythrose- 
4-P,  more  glucose-6-P  might  be  diverted  into  the  pentose-P  pathway  or, 
if  fructose-6-P  is  being  metabolized,  a  negative  feedback  effect  would  be 
exerted.  A  third  potent  inhibitor  is  glucosamine-6-P  (Wolfe  and  Nakada, 
1956),  which  on  the  T.  spiralis  enzyme  is  around  40  times  more  inhibitory 
than  6-phosphogluconate,  0.06  mM  inhibiting  64%  when  glucose-6-P 
is  5  mM  (Mancilla  and  Agosin,  1960).  The  deamination  of  glucosamine-6-P 
in  E.  coli: 

Glucosamine-6-P   ->  fructose-6-P  :f±  glucose-6-P 

thus  terminates  at  fructose-6-P  initially  because  of  the  inhibition  of  the 
second  reaction,  but  eventually  disappearance  of  glucosamine-6-P  relieves 
the  inhibition  and  glucose-6-P  is  formed,  another  example  of  metabolic 
regulation  through  inhibition. 

Aldolase 

This  enz^Tne  splitting  fructose- 1,6-diP  to  glyceraldehyde-3-P  and  dihy- 
droxyacetone-P  has  unfortunately  been  studied  very  little  from  the  stand- 
point of  inhibition  by  hexose  or  triose  phosphates.  Yeast  aldolase  binds  sev: 
eral  hexose  phosphates  quite  tightly  but  splits  them  very  slowly  (Richards 
and  Rutter,  1961),  as  may  be  seen  from  the  K/s  and  the  relative  reac- 
tion rates  (rate  with  fructose-l,6-diP  as  1)  in  the  accompanying  tabulation. 


Inhibitor  ,      ',  Pielative  rate 

(mM) 


L-Sorbose-l,6-diP 

0.13 

0.0014 

L-Sorbose-1-P 

0.2 

0.0002 

D-Fructose-l-P 

1.0 

0.0004 

D-Fructose-6-P 

3.8 

0.0001 

Muscle  aldolase  splits  these  analogs, niueh  more  readily.  The  aldolase  from 
rabbit  Ihuscle  was  found  by  Herbert  et  al.  (1940)  to  be  competitively 
inhibited  by  fructose-6-P  (31%),  fructose  (16.3%),  and  glucose  (5.4%),  the 
percentages  being  for  20  mM  inhibitor  and  fructose-l,6-diP.  The  difficulty 
in  interpreting  these  results  lies  in  our  ignorance  of  the  preferred  form  of 
the  substrate  (i.  e.,  a  or  j3,  p>Tanose,  furanose,  or  linear)  for  the  enzyme, 
and  the  distribution  of  the  substrates  and  inhibitors  among  these  forms 
under  the  experimental  conditions.  The  values  of  K^  may  be  quite  misleading 


408  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

because  the  inhibiting  form  could  represent  only  a  small  fraction  of  the  total 
concentration  of  the  inhibitor. 

Glyceraldehyde-3-Phosphate  Dehydrogenase 

A  new,  potent,  and  apparently  specific  inhibitor  of  this  enzyme  has 
been  found  and  is  likely  to  be  useful  as  a  blocking  agent  of  this  step  in 
glycolysis.  Glycolaldehyde-2-P  freshly  prepared  does  not  inhibit  glyceral- 
dehyde-3-P  dehydrogenase,  but  either  aging  the  preparation  or  allowing  it 
to  react  in  1  N  NaOH  for  a  short  time  leads  to  the  formation  of  a  potent 
inhibitor,  termed  tetrose-diP  by  Racker  et  al.,  (1959)  and  isolated  as  d- 
threose-2,4-diP  by  Fluharty  and  Ballou  (1959).  Both  the  d-  and  L-isomers 

CHO 


CHO 

"OgP— 0— C— H 

CHa— 0— P03=    — 

H— C— OH 

CH^— 0— PO 

ycolaldehyde-2-P 

D-Threose-2,4-diP 

of  the  inhibitor  were  synthesized  by  the  latter  workers  and  only  the  D-iso- 
mer  was  found  to  inhibit  strongly.  The  inhibition  is  reversible  but  non- 
competitive with  respect  to  glyceraldehyde-3-P;  the  inhibition  may  actually 
increase  with  NAD  concentration.  The  value  of  K,  for  rabbit  muscle  en- 
zyme is  0.0001-0.0002  milf.  D-Threose-2,4-diP  is  oxidized  by  the  enzyme 
but  only  as  much  as  the  NAD  present.  Fluharty  and  Ballou  postulated 
that  it  might  react  with  a  site  other  than  the  normal  catalytic  site  for 
oxidation  of  glyceraldehyde-3-P  but  Racker  etal.,  showed  spectroscopically 
that  a  stable  acyl-enzyme  complex  is  formed,  probably  with  the  SH  group 
known  to  be  involved  in  the  catalysis  and  the  binding  of  NAD.  The  sub- 
strate reactions  might  be  written  as: 

Enzyme-NAD  -f  Glyceraldehyde-3-P  :f±  phosphoglyceryl-enzyme-NADH 
Phosphoglyceryl-enzyme-NADH  +  P  5±  glycerate-l,3-diP  +  enzyme-NADH 

whereas  the  reaction  with  the  inhibitor  is: 

Enzyme-NAD  +  threose-2,4-diP  :^  diphosphothreonyl-enzyme-NADH 

Normally  an  inorganic  phosphate  is  transferred  from  a  second  site  to  the 
phosphoglyceryl-enzyme  to  form  glycerate-l,3-diP,  but  the  location  of  the 
2-phosphate  group  on  the  inhibitor  is  such  as  to  occupy  this  second  site 
so  that  no  phosphate  can  enter  the  reaction.  This  is  thus  an  example  of 
an  inhibition  in  which  a  substance  enters  the  reaction  sequence  in  the  same 


INHIBITORS    OF    CARBOHYDRATE    METABOLISM  409 

manner  as  the  substrate,  forming  a  stable  complex  with  the  enzyme,  but  is 
unable  to  complete  the  sequence. 

The  use  of  D-threose-2,4-diP  to  block  glycolysis  in  intact  cells  or  tissues 
will  probably  not  meet  with  general  success  due  to  the  poor  penetration. 
Extracts  of  ascites  tumor  cells  are  inhibited  readily,  but  glycolysis  in  intact 
ascites  or  HeLa  cells  is  unaffected  (Racker  et  al.,  1959).  Preparations  of 
glycolaldehyde-2-P,  containing  the  tetrose-diP,  inhibit  the  growth  of  some 
bacteria  and  not  others,  probably  depending  on  the  degree  of  penetration. 
The  possibility  was  considered  that  the  inhibitor,  can  be  formed  intracellu- 
larly  but  so  far  appropriate  precursors  have  not  been  found. 

If  photosynthesis  involves  the  reversal  of  the  glycolytic  sequence, 
D-threose-2,4-diP  should  inhibit,  and  it  has  been  found  that  the  total  C^^Og 
fixation  in  sonically  ruptured  spinach  chloroplasts  is  reduced  57%  by  this 
analog  at  0.1  mM  (Park  et  al.,  1960).  The  photoreduction  of  3-phospho- 
glycerate  is  inhibited  and  this  substance  accumulates  in  the  presence  of  the 
inhibitor.  This  inhibition  is  due  primarily  to  an  action  on  glyceraldehyde-3-P 
dehydrogenase  leading  to  a  deficiency  of  ribulose-l,5-diP,  the  CO2  acceptor. 
Carboxy  dismutase  is  inhibited  only  slightly  by  0.1  mM  D-threose-2,4-diP 
but  higher  concentrations  inhibit  appreciably.  This  inhibitor  may  thus  play 
a  role  in  photosynthetic  studies. 

Keleti  and  Telegdi  (1959)  examined  various  glyceraldehyde-3-P  dehy- 
drogenases and  found  inorganic  phosphate  to  stimulate  the  activity  at  low 
concentrations  but  to  inhibit  progressively  above  20-30  mM.  The  inhi- 
bition of  the  yeast  enzyme  is  competitive  with  both  glyceraldehyde-3-P 
and  NAD.  Taylor  et  al.,  (1963)  also  found  phosphate  inhibition  of  the 
hydrolysis  of  p-nitrophenylacetate  by  glyceraldehyde-3-P  dehydrogenase, 
the  muscle  enzyme  being  much  more  sensitive  than  the  yeast  enzyme. 
Indeed,  the  muscle  enzyme  is  56%  inhibited  already  at  10  mM  phosphate. 
Phosphate  inhibitions  of  the  glycolytic  enzymes  have  seldom  been  consid- 
ered as  playing  a  role  in  the  regulation  of  carbohydrate  metabolism. 

Enolase 

Enolase  catalyzes  the  reaction 

D-2-phosphoglycerate  :^  phosphoenolpyruvate 

and  various  analogs  of  these  substances  inhibit  competitively  (see  tabula- 
tion) (Wold  and  Ballou,  1957).  The  following  substances  do  not  inhibit: 
D-lactate,  D-glyceraldehyde-3-P,  dihydroxyacetone-P,  and  glycerol-2-P.  An 
inhibitor  must  have  a  carboxyl  and  a  phosphate  group,  and  the  distance 
between  them  is  of  some  importance,  although  not  critical.  The  3-methyla- 
tion  of  2-phosphoglycerate  does  not  lead  to  much  reduction  in  binding,  but 
3-methylation  of  the  3-phosphoglycerate  reduces  the  afiinity  by  approxi- 


410 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


Inhibitor 

{mM) 

D-Phospholactate 

0.35 

^-Hydroxypropionate-P 

0.45 

D  -  3  -  Phosphogly  cerate 

0.45 

D-er;/<Aro-2,3-Dihydroxybutyrate 

-2-P 

0.60 

D-er2/<7»ro-2,3-Dihydroxybutyrate-3-P 

3.3 

COO- 

HCOP03H- 
H2COH 

D-2-Phosphoglycerate 


coo- 

I 

CH, 

H2COPO3H- 

/9-Hydroxypropionate-P 


coo- 

HCOH 

H2COPO3H- 
D  -3  -  Phosphogly  cerate 

coo- 

HCOPO3H- 
HCOH 

CH3 

D-er7/<A/-o-2,3-Dihydroxy- 
butvrate-2-P 


COO- 

HCOP03H- 
CH3 

D-Phospholactate 

coo- 

HCOH 
HCOPO3H- 

CH3 

T>-erythro-2,Z-T)'\\ry- 
droxybutyrate-3-P 


mately  1.2  kcal/mole.  This  would  indicate  that  2-  and  3-phosphates  fit 
comparably  as  long  as  there  is  no  bulky  group  on  the  3-position,  but  when 
there  is  it  sterically  interferes  with  the  bending  of  the  3-phosphate  to  fit  the 
active  site. 

Glucose   and    Glucose-6-P   Dehydrogenases 

Beef  liver  glucose  dehydrogenase  is  inhibited  strongly  and  competitively 
by  glucose-6-P  (Strecker  and  Korkes,  1952)  and  fructose- 1,6-diP  (Brink, 
1953  a),  the  K-b  being  0.0025  mM  and  0.062  mM,  respectively.  The  K^ 
for  glucose  is  around  31  mM  at  pH  7,  so  that  if  this  represents  a  dissociation 
constant  the  phosphorylated  compounds  are  bound  much  more  tightly 
(around  5.8  kcal/mole).  The  rat  liver  enzyme  is  similarly  inhibited:  glucose- 
6-P  at  0.015  mM  inhibits  78%  when  glucose  is  200  mM  (Metzger  et  al, 
1964).  Glucose-1-P  also  inhibits  but  is  about  one-tenth  as  effective  as  glu- 
cose-6-P.  Ribose-5-P  and  fructose-6-P  are  also  less  inhibitory  (Brink,  1953  a). 
The  potency  of  the  glucose-6-P  inhibition  is  surprising  and  it  is  quite  likely 
that  it  may  be  important  in  metabolic  regulation  or  conserving  glucose  for 
phosphorylation.  Not  enough  is  known  about  this  enzyme  or  the  reaction 
mechanism  to  speculate  on  the  nature  of  the  interaction  of  these  inhibitors. 


INHIBITORS    OF    CAEBOHYDRATE    METABOLISM  411 

Yeast  glucose-6-P  dehydrogenase  is  inhibited  competitively  by  glucos- 
amine-6-P  but  in  this  case  the  affinity  for  the  inhibitor  is  not  as  great 
as  for  the  substrate  {K,,,  =  0.058  mM,  and  K^  =  0.72  mM)  (Glaser  and 
Brown,  1955).  Neither  mannose-6-P  nor  N-acetyl-D-gkicosamine-6-P  inhi- 
bits. Phosphate  inhibits  the  yeast  enzyme  rather  weakly  but  it  is  competitive 
with  respect  to  NADP.  The  enzyme  from  Prototheca  zopfii,  however,  is  inhi- 
bited 13%  by  0.07  mM  phosphate,  43%  by  0.14  mM,  and  70%  by  0.7 
mM  (glucose-6-P  =  3  mM  and  NADP  =  0.67  mM)  (Ciferri,  1962).  There 
needs  to  be  much  more  study  of  the  inhibition  of  such  enzymes  if  we  are 
to  understand  the  controlling  factors  of  carbohydrate  metabolism. 

Phosphopentose  Isomerases 

The  phosphoribose  isomerase  of  alfalfa  is  j^otently  inhibited  by  5-phospho- 
ribonate  and  much  less  so  by  a  variety  of  related  substances  (as  shown  in 
the  accompanying  tabulation)  (Axelrod  and  Jang,  1954).  Since  ribose-5-P 


Inhibitor 

Concentration 
{mM) 

%  Inhibition 

5-Phosphoribonate 

0.13 

50 

Glucose-6-P 

11.3 

32 

Phosphate 

25 

49 

Ribose 

11.5 

0 

Ribose-3-P 

12.5 

0 

was  2.5  mM,  the  5-phosphoribonate  is  bound  more  tightly.  However,  even 
19  mM  5-phosphoribonate  does  not  inhibit  the  growth  of  Leuconostoc  mes- 
enteroides  or  Lactobacillus  arabinosus  in  glucose  medium;  this  could  mean 
that  the  pentose  phosphate  pathway  is  not  necessary  for  growth  of  these 
organisms  or  that  the  inhibitor  does  not  penetrate  readily.  The  Ophiodon 
elongatus  muscle  enzyme  is  also  inhibited  by  5-phosphoribonate  but  less 
potently  (28%  inhibition  at  0.5  mM  and  45%  at  5  mM)  (Tarr,   1959). 

The  phosphoarabinose  isomerase  of  Propionibacterium  pentosaceum  is  not 
inhibited  by  D-ribose,  D-xylose,  D-arabinose,  L-arabinose,  ribose-5-P,  and 
glucose-6-P  at  10  times  the  concentration  of  arabinose-5-P  (Volk,  1960), 
and  the  phosphopentose  isomerase  of  EcJiinococcus  granvlosiis  hydatid  cysts 
is  not  inhibited  by  glucose,  fructose,  glucosamine,  glucose-6-P,  fructose- 
6-P,  and  mannose-6-P  at  4  mM  (Agosin  and  Aravena,  1960).  However,  the 
latter  enzyme  is  inhibited  51%  by  1.2  mM  dihydroxyacetone-P  and  54%  by 
4  mM  dihydroxyacetone,  so  that  conditions  favoring  accumulation  of  these 
substances  might  suppress  the  alternate  pentose-P  pathway,  which  could 
be  important  in  the  results  obtained  with  iodoacetate  or  D-threose-2,4-diP. 


412  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Glucose-6-Phosphatase 

The  hydrolysis  of  glucose-6-P  in  rat  liver  preparations  (probably  microso- 
mal) is  inhibited  by  glucose  with  a  K^  around  29  mM  (Langdon  and  Weak- 
ley, 1957).  This  inhibition  is  noncompetitive  with  respect  to  glucose-6-P, 
and  it  was  postulated  by  Segal  (1959)  that  the  glucose  competes  with  the 
second  substrate,  water,  for  its  site;  the  transfer  of  phosphate  could  occur 
either  to  water  or  another  molecule  of  glucose.  If  this  is  true,  incorporation 

(1)  Enzyme  +  glucose-6-P  -^  enzyme-gIucose-6-P 

(2)  Enzyme-glucose-6-P  :^  enzyme-P  +  glucose 

(3)  Enzyme-P  +  water  ^  enzyme  +  Pt 

(4)  Enzyme-P  +  glucose  5=^  enzyme-glucose-6-P 

of  C^*  from  glucose-C^*  into  glucose-C^'*-6-P  should  occur  and  this  was  dem- 
onstrated. It  was  shown  that  with  the  appropriate  rate  constants  the  re- 
actions (l)-(4)  (reaction  (4)  is  of  course  the  reverse  of  (2)  and  is  included  to 
visualize  competition  between  water  and  glucose)  lead  to  noncompetitive 
kinetics.  A  study  of  the  inhibition  of  this  incorporation  by  Hass  and  Byrne 
(1960)  showed  that  glucose  is  the  most  potent  inhibitor  of  various  sugars 
tested. 

Miscellaneous  Inhibitions 

Numerous  other  analog  inhibitions  of  enzymes  involved  in  carbohydrate 
utilization  have  been  reported,  some  of  which  have  been  summarized  in 
Table  2-21.  These  inhibitions  along  with  those  previously  discussed  point 
to  many  mechanisms  for  the  control  of  glucose  metabolism.  The  complex 
interplay  between  all  the  sugars  and  their  phosphorylated  derivatives  with 
respect  to  the  inhibition  of  various  enzymes  in  the  different  available  path- 
ways must  always  be  borne  in  mind  in  work  on  intact  cells.  Many  of  these 
enzymes  are  also  inhibited  to  varying  degrees  by  inorganic  phosphate;  hence 
the  level  of  phosphate  can  also  be  a  regulating  factor.  Enzymes  inhibited 
by  phosphate  include  phosphodeoxyribomutase,  phosphoribose  isomerase, 
phosphoglucose  isomerase,  triosephosphate  isomerase,  glucose-6-P  dehydro- 
genase, enolase,  glyceraldehyde-3-P  dehydrogenase,  phosphorylase,  trans- 
aldolase,  transketolase,  glucose-6-phosphatase,  and  ribulose-P  carboxylase. 
In  many  instances  appreciable  inhibition  is  exerted  by  5-20  niM  phosphate. 
An  interesting  study  of  phosphate  inhibition  of  transaldolase  was  made  by 
Bonsignore  et  al.  (1960)  in  which  the  following  reactions  were  examined: 

Fructose-6-P  +  glyceraldehyde   ->  glyceraldehyde-3-P  +  fructose 
Sedoheptulose-7-P  +  glyceraldehyde-3-P  ->  fructose-6-P  +  erythrose-4-P 

Phosphate  inhibits  the  first  reaction  competitively  with  respect  to  fruc- 


INHIBITORS    OF    CARBOHYDRATE    METABOLISM 


413 


P5 


-     -r.     m 


O  3 


P-I       (=^ 


O 


K 


M 

O 

Pm 

^ 

cc 

ci 

ci 

*^ 

Ph 

O 

i 

ci 

i 

CB 

o 

& 

CO 

o 

o 

0) 

^ 

+i 

o 

o 

5£ 

-^ 

o 

rt 

o 

O 

' 

3 

X5 

ro 

oj 

i-f 

t^ 

o 

O 

P5 

-tj 

<B 

e 

]5 

C 

•§ 

Xi 

'> 

g 

di 

o 

e 

P5 

m 

^ 

c 
■ft 


O     O 


'^ 

t3 

c 

s 

cS 

0) 

0) 

o 

o 

>. 

t>-. 

o 

O 

^ 

^ 

w 


^ 

^ 

Ph 

'? 

<a 

<ii 

« 

I> 

Ph 

rH 

t^ 

fij 

3 

fe 

o 

C/3 

'S 

H 

^ 

_o 

o 

e 

'bi 

"m 

o 

^ 

5 

»-H 

D 

o 

o 

"a 

cc 
O 

ST 

p 

o 

o 

O 

03 

O 

0^ 
O 
o 

0) 
CO 

'o 

1 

"o 

o 

o 

01 

^ 

r^ 

j2 

O 

^ 

^ 

13 

3 

Q 

Ph 

Ph 

'S 

■>> 

T. 

Lh 

o 

iC 

m 

m 

fe 

cc 

CO 

'^^ 

O 

X 

XI 

M 

fej 


Ph 


Q 


P5 


O 


-7=^.        O   -3 


W 


O 


414  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

tose-6-P  but  inhibits  the  second  reaction  noncompetitively  with  respect  to 
sedoheptulose-7-P  or  glyceraldehyde-3-P.  Since  the  ^/s  are  roughly  the 
same  for  all  inhibitions  (around  60  vnM),  a  single  binding  site  for  phosphate 
was  postulated.  Phosphate  prevents  the  formation  of  the  transaldolase- 
dihydroxyacetone  complex  but  does  not  interfere  with  the  transfer  of  the 
dihydroxyacetone  to  its  acceptor. 

Multivalent  anions  in  general  are  inhibitors  of  glycolysis.  The  anaerobic 
formation  of  lactate  from  glucose  in  pigeon  hemoly sates  is  depressed  84% 
by  40  mM  sulfate,  64%  by  20  mM  phosphate,  100%  by  4.2  mM  oxalate, 
and  48%  by  0.1%  ribonucleate  (Dische  and  Ashwell,  1955).  The  reactions 
from  glucose  -^  lactate  are  inhibited  more  strongly  than  from  fructose-1,6- 
diP  — >  lactate;  therefore  there  is  inhibition  previous  to  fructose-l,6-diP.  The 
sequence  from  glucose  -^  glyceraldehyde-3-P  is  inhibited  strongly  by  sulfate 
and  ribonucleate.  Aldolase  is  inhibited  about  20%  by  40  mM  sulfate  but  is 
not  affected  by  oxalate.  It  was  concluded  that  there  are  three  sites  of  action, 
the  strongest  on  hexokinase,  next  on  glyceraldehyde-3-P  dehydrogenase, 
and  the  last  possibly  on  pyruvate  kinase,  the  inhibitions  being  competitive 
with  ATP  or  NAD.  Actually  little  positive  evidence  was  provided  for  these 
sites  and  other  possibilities  are  just  as  likely;  furthermore,  competition  with 
hexose  phosphate  and  glycerol  phosphate  is  also  possible.  In  addition  the 
complexing  of  Mg++  by  these  anions  must  be  considered  since  several  gly- 
colytic enzymes  are  activated  by  this  ion. 

It  will  suffice  to  mention  four  additional  interesting  examples  of  analog 
inhibition  on  this  phase  of  metabolism.  Cataractogenic  sugars  and  polyols 
inhibit  lens  mutarotase  while  noncataractogenic  sugars  do  not;  the  K,'s  are 
4  mM  for  galactose,  15  mM  for  xylose,  and  100  mM  for  sorbitol  (Keston, 
1963).  Despite  the  weak  inhibitory  activity  of  sorbitol,  there  is  a  large 
amount  in  the  lens  in  certain  conditions,  such  as  diabetes  (perhaps  around 
30  mM).  Mannose  is  quite  toxic  to  honeybees;  of  bees  offered  1  M  mannose 
solution,  50%  were  dead  in  90  min  and  over  90%  in  3  hr  (Sols  et  al.,  1960  a). 
It  was  found  that  bees  have  a  hexokinase  very  active  toward  mannose 
coupled  with  a  negligible  amount  of  phosphomannose  isomerase,  so  that 
mannose  not  only  may  interfere  with  phosphorylation  of  glucose  and  fruc- 
tose, but  many  accumulate  as  mannose-6-P,  which  could  disturb  glycolysis 
in  a  number  of  ways.  Xylose  appears  to  be  in  some  manner  a  specific  inhibi- 
tor of  photosynthesis,  since  Chlorella  propagation  is  not  inhibited  by  0.5- 
1.5%  xylose  when  glucose,  fructose,  or  mannose  is  present  (thus  it  is  not 
inherently  toxic),  but  under  photosynthetic  conditions  the  cells  rapidly  lose 
their  color  and  ability  to  divide,  an  effect  that  can  be  reversed  by  glucose 
(Hassall,  1958).  It  was  postulated  that  xylose  may  compete  with  xylulose- 
5-P  for  an  enzyme  in  the  transketolase  pathway  and  block  photosynthesis; 
on  the  other  hand,  a  phosphorylated  product  may  be  the  active  inhibitor. 
Analogs  without  the  usual  hexose  structure  may  also  inhibit  glycolysis  and 


GLYCOSIDASES  415 

the  pentose-P  pathway.  Sahasrabudhe  et  al.  (1960)  in  looking  for  carcino- 
static  analogs  found  that  thiophene-2,5-dicarboxylate,  which  might  be  con- 
sidered as  an  analog  of  substances  such  as  ribose-5-P,  inhibits  the  formation 
of  C^^Og  from  glucose- 1-C^*  and  glucose-6-C^'*  around  43%  (concentration 
unspecified)  in  tumor  tissue,  and  suppresses  the  growth  in  vivo  of  rat  sar- 
coma. No  evidence  was  given  as  to  the  site  of  action  and  so  the  assumption 
is  tenuous,  but  the  concept  of  using  heterocyclic  substances  analogous  to 
the  furanose  and  pyranose  structures  may  be  important. 

GLYCOSIDASES 

This  large  group  of  enzymes  hydrolyzing  the  glycosidic  bonds  of  simple 
glycosides,  oligosaccharides,  and  polysaccharides  has  been  studied  for  many 
years  and  it  is  not  surprising  that  numerous  instances  of  analog  inhibition 
have  been  observed.  Most  of  the  reports,  although  important  in  themselves, 
do  not  lend  themselves  to  interpretations  on  the  molecular  level;  some  of 
the  data  have  been  briefly  summarized  in  Table  2-22.  Most  of  the  inhibi- 
tions are  relatively  weak  but  a  few  could  definitely  be  significant  in  meta- 
bolic regulation.  The  a,  (3  configuration  of  the  inhibitor  is  seen  to  be  im- 
portant in  some  instances.  However,  the  enzymes  are  seldom  completely 
specific  for  the  a-  or  /?-forms  and  the  affinities  often  diff"er  very  little,  as 
with  maltose  transglucosylase  (amylomaltase)  from  E.  coli  where  (see  ac- 
companying tabulation)  the  binding  difference  between  the  u-  and  /5-glu- 


Inhibitor 

(mM) 

Relative 

(k 

-  AF    of   binding 
cal/mole) 

/3-Methylmaltoside 

2.5 

3.70 

a-Methylglucoside 

8 

2.98 

^-Methylglucoside 

10 

2.84 

a-Phenylglucoside 

10 

2.84 

/S-Phenylglucoside 

30 

2.16 

cosides  is  0.14-0.68  kcal/mole  (Wiesmeyer  and  Cohn,  1960).  It  appears  that 
the  hydroxyl  groups  on  C-2,  C-4,  and  C-6  of  ring  A  are  involved  in  the 

CH,OH  CH,OH 

-Ott  XT  J O 


ring  A 


416 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


O 


H       K 


P5 


^-1 


s^ 


ft  ^  rS 

c 


g 

SI 

C 


S  ^ 


c 

-S3 
o 
O 

C 
>> 


^   ^     C     fl 


o  o 


w 


O   Q 


rs  "c 


o  o 


o  o    o  o  o 


(B  03  -u  o  M     o 

O  O  O  C  «',     -'-' 

^  ^  le  c«  '^    5      , 

O  O  O  S  ^  tR    «x 


"?     w 

o 

^ 

<c 

'm 

2   o 

fl 

"m 

^ 

OJ 

O 

i5     o 

a 

2 

o 

s 

O 

eg 

CO 

O 

C 

3 

'35 
o 

& 

O 

o 

a) 
O 
'n 

CO 

'be 

i 

=i.   a 

'ai. 

H^ 

H 

^ 

1— 1 

S 

H 

a 

P5 

Cl^ 


GLYCOSIDASES 


417 


W 


w 


TS 

lO 

a 

Oi 

c3 

^i;;;^ 

2 

T3 

!» 

K} 

c< 

g 

c 

<» 

o 

s 

^ 

o 

J 

D5 

in  ic  »o  >c 


—  ^  -^  c 


o  o  o 


O    ><    CK    S    O 


0 

0 

0 

<» 

03 

C 

^ 

">> 

0 

8 

CI 

^ 

0 

£ 

r5 

_s 

0 

^ 

5 

0 

0 

0 

r5 

© 

0 

c« 

a 

0 

0 

0 

w; 

C 

C 

'2 

"0 

c 

O 

3 

0 

0 

Ch 

0 

^ 

Q 

0 

■4i 

_o 

0 

0 

0 

0 

_o 

0 

0 

■3 

S 

3 

02 

<» 

_3 

'S 

OJ 

-C 

"S 

^ 

0 

0 

C 

0 

Pn 

0 

0 

0 

O     ra     o     o 


^  -=;        ^    O 


O    ^3    S    M 


0 

^ 

_5 

'S 

^ 

^ 

03 

O! 

">> 

'>. 

0 

-»^ 

03 

■t^ 

03 

03 

0 

03 
0 

a 

8 

c 

8 

^ 

<^ 

ff 

<1 

s 

Is 

0 

^ 

S 

fe; 

eS 

3 
ft* 

'00 

o 

o 

O 
B 


Q::^. 

^^ 

^ 

10 

0 

0 

03 

n3 

s 

S 

43 

« 

^ 

4 

c 

0       M 

o3 

t-i       0 

(^ 

->J      C3 

-S 

t?  -2 

>> 

^^ 

c 

03 

03 

r2 

r-< 

ft 

0 

P 

-u 

^  s, 


03      cS 
CO 


C? 


►«;     S 


418 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


P5 


1^  '2 


.  o   ^ 


-^ 


H        2 


2  Xh 


O     O     O     O     O     (M 


■S    1^ 


03     £ 


o    'Ti 


y   o 


>■- 

j3 

<x> 

0) 

73. 

'w 

"ft 

o 

o 

-u 

s 

o 

4^ 

riS 

g 

'rt 

t*5 


t*3 


OJ  rfr 


^^(5^3^^^:^    53 


o 


GLYCOSIDASES  419 


.s  .s  ^  ^ 


^  Pm  ^ 


<N    O    O  ^  ■* 

O    t^    O  fO  o 


Q         O  O  O  O         O  O 


I        03 

-5  'cS 


c5    K-^      c3 


jg.a:        fe;=';^ra         fe;''<lj<t3 


(M 

o 

6 

c 
o 

T3 

o 

a 

4^ 

0 

o 

^ 

c 

c^ 

<u 

X 

o 

s 

03 

o 

T3 

o 

0)    y< 


^     •    o 


^:2 


fl    -=    O  G      cS      Qi 

2  -^-l  2  ->.  = 


OJ       03 


•<S    J 


e8 

r-2 

eS 

S 

cs    as 

r2 

M      03 

«i  '3 

■  'S 

^S 

«i.  'O 

2--S 

'?>;§ 

>>    eS 

>»    c6 

-»^     m 

4J        r- 

*^        CO 

©     O 

43      S 

®      O 

o     o 

■^      eS 

O      o 

<  ,B 

*?      O 

<    ^ 

fe,  '^ 

fe;  ^ 

)^      ^ 

420 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


P5 


!^ 


-  a 


>>  o  S, 

c 


C-  -H 


o 


o 


o 


o 


o 


-u 

s 

IS 

o 

o 

C<1 

O 

c 

'^ 

5 

o 

'2 

3 

6 

o 

o 

o 

c3 

Is 

tj 

o 

'2 

eS 

^ 

a 

'a 

'bjb 

0) 

's 

Is 

« 

c 

CO 

O 

o 

< 

a 

o 

X 

o 

c 
o 

o 

-2 

c3 

X 
O 

t3 

o 
o 

c    ce    (» 

£  ■>.  a 


a 


be 

CO 

oi. 

c3 
"2 

T. 

'c 

'a 

o 

a 

< 

S 

o  o  o  o  o  o 


03     M^     ^ 


_3 

^ 

cc 

O 

o 

^ 

$ 

a) 

o 

0) 

2 

o 

95 

^ 

3 

cS 

3 

5 

<^ 

a 

s 

><1 

a 


GLYCOSIDASES 


421 


S 

s 

o 

c 

^    O 


^ 


.""    --I 


O     >M     05     C5 
CO     ■<*     IC     -^ 


1^    05    ic    00 

>o    CC    Oi     t^ 

^     1— I     Oq 


•c:.    c 


o 


aj     o     o'     © 


O       c«       O       t<        -*i 

-T  1  2  §  3 

o  O  fi<  oj    s 


> 

c 

CO    55 


o 


o 


t3     <D    aj 

2  ^  :« 


te 

73 

Xh 

eS 

C 

■ ■ 

43 

c3 

c 

Ph 

ro 

73 

'J 

c 

"s 

CO 

cS 

eS 

45 

4m 

o 

-C 

o 

v 

3 
1-5 

S 

(1h 

Cm 


•^    .-^     O    IC    t-    IC 


o   o   o   o 


73     Ti 


h 

3 

_s 

0) 

K 

OJ 

13 

o 

^ 

O 

-l-i 

^Ji 

"bX) 

"m 

Q 

o 

o 

Eh 

o 

15 

i 

o 
S 

5 

3 

o 

0 

■^i. 

P^ 

a 

■ai. 

c? 

=Ci. 

_o 

43 

'-3 

^ 

'2 

.,— ^ 

;s 

^ 

_c 

a 

^ 

o~" 

c 

bl 

_o 

OJ 

'-3 

73 

f« 

G 

v 

o 

aJ 

fl 

_> 

o 

■-13 

o 

-2 

ft 

cS 

bi 

c 

-kJ 

o 

oc 

o 

42 

i=; 

3 

o 

01 

c 

(X> 

11 

43 

H 

o 

C 

^ 

_o 

■-!3 

b 

o 

'2 

<» 

:a 

> 

.g 

•^ 

'-3 

4J 

o 

& 

cc 

g 

^ 

-fi 

O 

c3 

o 

o 

•  ^ 

11 

73 

o 


73 

a 

c3 


'bi  T) 


.2    73 


a  -i5 


42      4) 

-I 


ft  42 


4) 


.o    <u 


422  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

binding  since  alteration  of  the  positions,  as  in  a-methylmannoside  or  a- 
methylgalactoside,  abolishes  the  inhibition,  and  omission,  as  in  a-methyl- 
xyloside,  also  prevents  binding.  The  substitution  of  a  methyl  group  on  C-1 
of  ring  B  (/?-methylmaltoside)  does  not  interfere  with  the  binding,  since 
this  analog  has  approximately  the  same  affinity  for  the  enzyme  as  the  sub- 
strate maltose.  Alteration  of  the  glycosidic  link  from  the  o;-l,4  in  maltose 
to  the  /?-l,4  in  cellobiose  does  not  result  in  an  inhibitor,  and  other  changes 
(as  in  trehalose,  sucrose,  dextran,  or  lactose)  likewise  reduce  the  affinity. 
The  requirements  for  binding  may  be  summarized  as  (1)  a  glucose-like  con- 
figuration of  hydroxyl  groups  in  ring  A,  (2)  a  glycosidic  link  of  the  a- 1,4 
type,  and  (3)  a  widely  variable  glycosidic  group. 

An  a-mannosidase  of  Streptomyces  griseus  with  a-phenylmannoside  as  the 
substrate  is  inhibited  competitively  (probably)  by  several  sugars  and  gly- 
cosides (Hockenhull  et  al.,  1954  c)  and  the  relative  binding  energies  have 
been  estimated  on  this  basis  (see  tabulation),  although  the  reliability  of 


Inhibitor 

%  Inhibition  at 
50  mM 

Relative  —  Zli^  of  binding 
(kcal/mole) 

a-Methylmannoside 

96 

5.23 

Mannose 

83 

4.25 

Cellobiose 

82 

4.20 

Maltose 

73 

3.88 

a-Methylglucoside 

14 

2.15 

Xylose 

11 

1.98 

Arabinose 

10 

1.91 

Sucrose 

7.5 

1.72 

Glucose 

5 

1.45 

Fructose 

5 

1.45 

Ribose 

4 

1.31 

Mannitol 

2 

0.87 

Rhamnose 

0 

— 

these  figures  is  quite  low  due  to  variations  between  experiments.  The  man- 
nose  configuration  seems  to  confer  strong  binding,  as  expected,  but  the  high 
inhibitory  activities  of  maltose  and  cellobiose  {a  and  /5  glucosides,  respec- 
tively), especially  in  view  of  the  weak  inhibitions  produced  by  a-methyl- 
glucoside  and  glucose,  are  unexpected  and  perhaps  indicate  that  these  sub- 
stances are  oriented  on  the  enzyme  surface  in  a  different  way  than  the 
substrate,  although  there  is  no  necessity  to  postulate  an  irreversible  com- 
plex as  did  Hockenhull  and  his  co-workers.  The  active  centers  for  such 
enzymes  probably  possess  several  binding  groups  for  the  hydroxyls  arrang- 
ed in  a  certain  pattern;  glycosides  other  than  the  substrate  could  conceivably 


GLYCOSIDASES  423 

interact  satisfactorily  with  this  pattern  by  appropriate  translation  or  ro- 
tation of  the  molecules  (indeed,  the  other  sides  of  these  roughly  planar 
molecules  could  possibly  fit  the  pattern  in  some  cases). 

One  of  the  most  interesting  studies  of  analog  inhibition  was  reported 
by  Halvorson  and  Ellias  (1958)  for  the  a-glucosidase  of  Saccharomyces 
italicus,  the  data  from  which  are  given  in  the  accompanying  tabulation. 


Inhibitor 

{mM) 

Relative  —  AF  oi  binding 
(kcal/mole) 

a-Phenylgliicopyranoside* 
Glucose 

0.3 
1.2 

5.00 
4.15 

a-Butylglucopyranoside 

a-Ethylthioglucopyranoside 

Turanose* 

1.9 

4 

9.5 

3.86 
3.40 

2.87 

Isomaltose 

11 

2.78 

Xylose 
Sucrose* 

12 
23 

2.72 
2.32 

/S-Methylmaltoside* 

a-Ethylglucopyranoside 

Maltose* 

37 

45 
81 

2.04 
1.91 
1.55 

a-Ethylthioglucofuranoside 

86 

1.51 

Arabinose 

110 

1.36 

a-Methylglucopyranoside 

400 

0.56 

Several  of  the  substances  (marked  by  *)  are  also  substrates;  maltose,  in 
fact,  is  probably  the  natural  substrate.  Inversion  of  hydroxyl  groups  on 
C-2  and  C-4,  and  substitution  or  oxidation  at  C-6,  abolish  the  affinity. 
Increasing  the  size  of  the  aglycone  groups  increases  the  affinity.  The  rather 
potent  inhibition  by  glucose  is  surprising  and  indicates  that  the  alkyl  agly- 
cone groups  actually  reduce  the  affinity.  This  might  mean  that  the  C-1 
hydroxyl  group  can  be  involved  in  the  binding;  a  small  substituent  prevents 
this  but  the  binding  increases  as  the  alkyl  group  is  lengthened. 

Inhibition   of  (3-Glucuronidases   by   D-Glucaro-1,4-lactone 
and  Related  Compounds 

These  widely  distributed  enzymes  catalyze  the  hydrolysis  of  /5-glucur- 
onides  occurring  naturally  in  plants  and  animals,  and  in  addition  may  play 
a  role  in  mucopolysaccharide  metabolism.  They  are  not  involved  in  the 
glucuronide  syntheses  of  detoxification,  these  reactions  being  catalyzed  by 
glucuronyl  transferases.  There  is  evidence  that  /^-glucuronidases  are  related 
in  some  manner  to  growth  and  the  activities  of  certain  tissues,  and  it  was 


424 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


to  establish  their  role  in  metabolism  that  Karunairatnam  and  Levvy  (1949) 
searched  for  specific  inhibitors.  The  most  potent  of  the  analogs  tested  is 
glucarate  (K^  =  0.06  mM,  and  K,^,  =  3.5  for  phenyl-/?-glucuronide)  but 
subsequent  studies  in  different  laboratories  showed  great  variability  in  ac- 
tivity. The  problem  was  solved  by  Levvy  (1952),  who  found  that  the  ac- 
tive inhibitor  is  glucaro-l,4-lactone,*  which  is  formed  from  glucarate  and 
occurs  to  varying  extents  in  different  preparations.  The  tabulation  shows 


Inhibitor 

(mM) 

Relative  —  AF  oi  binding 
(kcal/mole) 

Glucaro- 1 ,4-lactone 

0.00054 

8.87 

3-Methylglucaro- 1 ,4-lactone 

0.13 

5.50 

Glucarate 

0.17 

5.33 

Glucaro-3,6-lactone 

0.48 

4.70 

Glucuronate 

1.6 

3.96 

Galacturonate 

6.0 

3.14 

Galactarate  (mucate) 

6.0 

3.14 

the  inhibitor  constants  obtained  on  mouse  liver  /^-glucuronidase  hydrolyz- 
ing  phenolphthalein-/?-glucuronide,  and  the  particular  potency  of  the  glu- 
caro-1,4-lactone  is  evident."^  No  inhibition  at  1  mM  is  observed  with  man- 
narate,  mannaro-l,4-3,6-dilactone,  4-methylglucuronate,  and  ascorbate,  or 
at  10  mM  with  3-methylglucarate,  glucurone,  3-methylglucuronate,  man- 
nuronate,  mannurone,  and  2-keto-L-gulonate.  Heat  and  acid  treatments  of 
glucarate  and  galactarate  increase  the  inhibitory  activity  greatly  and  it  is 
possible  that  the  dicarboxylates  are  completely  inactive.  The  inhibition  by 
glucaro- 1,4-lactone  is  competitive  and  the  high  potency  probably  due  to 
the  structural  similarity  between  the  inhibitor  and  the  /5-glucofuranuronide 
form  of  the  substrate.  The  hydroxyl  configuration  in  the  furan  ring  is  very 
important  since  the  glucaro-3,6-lactone  is  bound  much  less  tightly  to  the 
enzyme,  and  the  3-OH  must  also  be  involved  since  methylation  reduces 
the  binding  by  over  3  kcal/mole.  These  rather  large  energy  differences  argue 


*  This  substance  has  also  been  called  saccharo- 1,4-lactone,  1 ,4-saccharolactone, 
glucosaccharo- 1,4-lactone,  and  other  names.  The  generic  name  for  the  dicarboxylic 
acids  derived  from  sugars  is  saccharic  acid  and  that  specifically  from  glucose  is  glu- 
cosaccharic  acid.  However,  it  seems  that  glucaric  acid  is  used  most  commonly  today 
and  this  terminology  is  more  consistent  with  the  naming  of  the  monocarboxylic  acids, 
so  that  the  lactones  will  here  be  named  accordingly. 

t  The  figures  in  this  and  other  reports  for  inhibitors  such  as  glucarate,  galactarate, 
glucuronate,  and  related  acids  may  reflect  to  varying  degrees  the  presence  of  lactones, 
since,  due  to  the  high  potency  of  the  lactones,  only  small  amounts  need  be  present. 


GLYCOSIDASES 


425 


5 

1 

o 

o- 

/ 

'^ 

SB 

-o 

murom 

tone 

one) 

/ 

^ 

r-O 

O— A. 
O- 

fur; 
lac 
cur 

O 

\ 

-^o 

-K 

-Gluco 
3,6- 
(glu 

\ 

o- 

-K 

1 

B3 

oa 

k" 

s 

o 

2 

s 

a 

o 

as 
u- 

s 
o 

1 

-u- 

EC 
1 

-o- 

a 

o 

1 

-u- 

O 

1 

-o- 

8 

-o 

1 

1 

o 

1 

1 

3    4) 

1 

'o    O     X    O     O    O 

5    I     I     I     I    x" 
o-u-u-u-u— u 

I    I    I    I 

X     O    X     X 
X 


-22 

8 

o- 

X 

o 

1 

-u- 

X 
X    o 

1    1 

-u-u- 

?8 

-u-u 

3    <U 

5^ 

1 

X 

1    i 

O    X 

X 

1 

X 

* 

2 

'? 

o 

a 

rt 

u 

0) 

3 

v 

OiS. 


426  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

against  van  der  Waals'  forces  being  a  major  factor  in  the  hydroxyl  inter- 
actions, and  suggest  hydrogen  bonding.  The  terminal  carboxylate  group 
also  participates,  since  glucaro-3,6-lactone  is  much  more  inhibitory  than 
glucurone.  (See  formulas  on  page  425). 

Some  inhibitor  constants  for  various  /^-glucuronidases  are  shown  in  Table 
2-23.  Not  enough  /^-glucuronidases  have  been  tested  with  glucaro-l,4-lac- 
tone  to  determine  the  variations  in  susceptibility,  but  from  the  limited 
data  it  appears  that  the  animal  enzymes  are  quite  sensitive  whereas  the 
plant  enzyme  baicalinase  is  much  less  readily  inhibited.  It  is  clear  that 
glucaro-l,4-lactone  is  one  of  the  most  potent  inhibitors  known  and  that 
none  of  the  other  analogs  so  far  examined  on  /^-glucuronidase  is  comparable, 
although  galactaro-l,4-lactone  is  undoubtedly  a  very  effective  inhibitor.  It 
is  interesting  that  the  /?-galacturonidases  of  limpet  and  preputial  gland  are 
approximately  as  susceptible  to  these  lactones  as  are  the  /^-glucuronidases 
(Marsh  and  Levvy,  1958). 

Limpet  or-glucuronidase  shows  quite  a  different  pattern  of  inhibition  (see 
accompanying  tabulation  for  inhibitions  at  5  mM  with  phenyl-a-glucur- 


Inhibitor 

%  Inhibition 

Relative   —  AF  of  binding 
(kcal/mole) 

Glucuronate 

68 

3.73 

Mannuronate 

43 

3.09 

Galacturonate 

28 

2.68 

Glucurone 

9 

1.84 

Xylono- 1 ,4-lactone 

22 

2.48 

Glucono- 1 ,5-lactone 

7 

1.67 

Arabono- 1 ,4-lactone 

4 

1.31 

Glucono- 1 ,4-lactone 

3 

1.13 

a-Glucuronate- 1  -phosphate 

55 

3.39 

/?-Glucuronate- 1  -phosphate 

22 

2.48 

Menthyl-a-glucuronide 

55 

3.39 

Menthyl-/S-glucuronide 

6 

1.57 

Borneol-a-glucuronide 

41 

3.04 

Veratroyl-/3-glucuronide 

9 

1.84 

onide  at  1  raM  with  relative  binding  energies  calculated  on  the  basis  of 
competitive  inhibition).  a-Glucuronides  are  more  inhibitory  than  /5-glucur- 
onides,  as  expected,  but  no  particularly  potent  inhibitors  were  found  {Kj^ 
for  glucuronate  is  0.54  mM  and  for  galacturonate  4.7  mM).  The  following 
do  not  inhibit:  mannurone,  glucaro-l,4-lactone,  mannono-l,4-lactone,  and 
galactarate.  The  difference  between  the  a  and  /?  enzymes  with  respect  to 
glucaro-l,4-lactone  is  especially  striking. 


GLYCOSIDASES 


427 


A 


00    -^    (M    —    O    -* 

•    c 

O    -H    CD    —   O    O 


Id 

_^    CO     <M     C5     05 

rZT  o  c-i  t^  ^  CO 

c 


c    c    c    c 


ooooco-*-^-^  — 


'H  :2 


O  GO  CO  ., 

O  Oi  o  (^i 

o  o  o  «^ 

O  O  O  O    >0    CO 

o  o  o  o   ^   -<t 


P   -:,   ci:i 

>»  =  -^  ^ 

■*^    Si    A      lA 


03      03      eS    ^^    -*^ 


o  o  o  o  o  o 


>    a 


a  O 


o  ^    p 

>  :§  o 


3      S      O    ^ 


T3  7:= 
«    o 


-r    TTn    =5     3    -T      _.    -Q. 


^     Is      «      C3      5 


^     6     rt    "ci     5     g     P 


2  ^  "3  ^  le  -2  Is 
S  O  a  O  O  O  O 


0) 

-r) 

r2 

'6 

'S 

*2 

"S 

o 

o 

o 

^ 

tH 

«H 

3 

3 

3 

O 

CJ 

o 

_3 

'S 

"3i 

bC 

tiC 

'S 

^^ 

oi. 

^ 

**?- 
%. 

o 

3 
o 

s 

'S 

-a 
o 
u 

>> 

^ 

N 

"^ 

"eg 

'm 

c 

j3 

<ai. 

■>* 

^ 

Bf 

T. 

ec 

'o 

"o 

3 
a 

3 
<» 

3 

3 

4) 

a 

Ph 

Pl, 

PlH 

M 

M 

oi. 

<ai. 

^ 

r- 

r2 

« 

'3 

13 

jil 

43 

-►i 

4i 

^ 

J3 

a. 

ft 

"o 

o 

3 

3 

05 

(» 

43 

^ 

Ph 

M 


GQ 


is 

^ 

+3 

OJ 

^ 

3 

'^^ 

<u 

-ti 

^ 

e 

cS 

^ 

05 

^ 

3     C 


03 

^ 

(C 

3 

"^  .^ 

^ 

O 

ai 

3 

0) 

-tj 

-S 

-0 

CJ 

c 

<A 

^ 

03 

.s 

OJ 

3 

' — V 

Q 

"3 

-3       CO 

2  "o 


^H 


cS      o 


f*      cS 


eS    73 


c 

<1) 

o 

ft 

cS 

3 

frt 

o 

4J 

Tl 

t^ 

tS 

ft 

428  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

It  was  previously  stated  that  /^-glucuronidases  do  not  seem  to  partici- 
pate in  glucuronide  synthesis  in  tissues  and  the  evidence  is  mainly  the 
lack  of  inhibition  by  glucaro-l,4-lactone  usually  observed  (Levvy  and 
Marsh,  1960).  For  example,  Lathe  and  Walker  (1958)  found  no  effects  of 
2  mM  glucaro-l,4-lactone  on  the  formation  of  o-aminophenylglucuronide 
or  bilirubin-glucuronide  in  liver  suspensions.  However,  Fishman  and  Green 
(1957)  reported  that  glucarate  potently  inhibits  glucuronyl  transfer  (50% 
inhibition  by  0.05  mM),  and  Sie  and  Fishman  (1954)  found  glucarate  and 
glucurone  to  inhibit  glucuronide  synthesis  in  liver  slices.  It  is  not  definitely 
known  if  these  inhibitions  relate  to  a  /^-glucuronidase  or  a  transferase,  but 
the  question  of  the  role  of  the  /^-glucuronidases  in  glucuronide  synthesis 
should  probably  be  left  open. 

The  urinary  bladder  cancer  that  occurs  in  people  employed  in  certain 
industries  may  be  related  to  the  formation  of  glucuronides  of  aromatic 
amines  in  the  body  and  their  subsequent  hydrolysis  by  urinary  /5-glucur- 
onidase.  The  oral  administration  of  inhibitory  analogs  to  patients  is  a 
possible  approach  to  the  prevention  of  such  cancers.  Boy  land  et  al.  (1957) 
found  that  the  urinary  enzyme  can  be  inhibited  by  administration  of  glu- 
conate and  glucarate  at  a  dose  of  10  g/day,  but  the  most  potent  inhibition 
is  produced  by  glucaro-l,4-lactone,  73%  inhibition  occurring  from  1.5-2 
g/day  and  90%  inhibition  from  4  g/day.  Nevertheless,  these  inhibitions  are 
less  than  expected  and  the  urinary  pH  was  reported  later  to  be  an  important 
factor,  the  inhibition  decreasing  as  the  pH  rises  above  6  (Boyland  et  al., 
1959).  Inhibition  of  liver  /^-glucuronidase  in  mice  by  the  administration 
of  glucaro-l,4-lactone  at  doses  from  50  to  800  mg/kg  (17%  to  76%)  was 
found  by  Akamatsu  et  al.  (1961).  The  maximal  inhibition  occurs  at  30-60 
min  and  after  2-4  hr  most  of  the  activity  has  returned.  Similar  results 
were  found  in  rat  liver  and  kidney.  Since  /^-glucuronidase  and  /?-iV-acetyl- 
glucosaminidase  hydrolyze  products  from  chondroitin  and  hyaluronate  and 
may  participate  in  the  metabolism  of  connective  tissue  mucoproteins,  and 
since  certain  tumors  have  relatively  large  amounts  of  these  enzymes,  Carr 
(1963)  administered  the  two  inhibitors  —  glucaro-l,4-lactone  and  2-acet- 
amido-2-deoxygluconolactone  —  to  mice  bearing  Tumor  2146.  At  150  mg/kg 
these  substances  are  nontoxic  and  cause  regression  of  the  tumors.  It  was 
suggested  that  the  inhibitors  prevent  the  penetration  of  the  tumor  cells 
through  the  intercellular  cement  substance,  but  this  is  admittedly  only  a 
tenuous  hypothesis.  Whatever  the  explanation,  it  represents  an  interesting 
approach  to  tumor  chemotherapy. 

Inhibition   of  Various  Glycosidases   by  Glucono-  and   Glucaro   Lactones 

The  inhibition  of  the  /^-glucuronidases  by  the  saccharo-l,4-lactones  sug- 
gested that  the  /^-glycosidases  might  be  inhibited  by  the  aldonolactones, 
and  this  was  found  to  be  so  by  Conchie  (1953,  1954).  A  sheep  rumen  /?- 


PYRUVATE  METABOLISM  429 

glucosidase,  possibly  involved  in  the  digestion  of  cellulose,  is  strongly  inhi- 
bited by  glucono-l,4-lactone  (^,  =  0.094  mM)  and  glucono-l,5-lactone 
(^.  =  0.091  milf )  in  a  competitive  manner,  the  affinity  of  the  enzyme  for 
these  analogs  being  about  10  times  that  for  the  substrate  o-nitrophenyl- 
/?-glucoside,  whereas  gluconate  itself  has  no  action.  Ox  liver  /?-galactosidase 
is  inhibited  much  more  potently  by  the  galactono-  and  fucono-l,5-lactones 
than  the  corresponding  1,4-lactones,  and  this  can  be  readily  explained  on 
the  basis  of  the  relationship  to  substrate  configuration  (Lewy  et  al.,  1962). 
A  remarkable  degree  of  specificity  is  exhibited  by  the  a  and  /?  glycosidases, 
inhibition  usually  resulting  only  from  the  corresponding  aldonolactone  in 
a  number  of  enzymes  from  different  sources  (Conchie  and  Lewy,  1957). 
For  example,  limpet  a-mannosidase  is  inhibited  markedly  by  mannono- 
1,4-lactone  but  not  by  the  glucono-,  galactono-,  arabono-,  or  xylonolactones. 
On  the  other  hand,  galactono- 1,4-lactone  is  specific  for  /?-galactosidase 
(Conchie  and  Hay,  1959).  Cellulytic  rumen  enzymes  are  inhibited  to  vary- 
ing degrees  by  glucono- 1,4-lactone,  depending  on  the  substrate  chain  length; 
hydrolysis  of  cellobiose  is  inhibited  99%,  of  cellotriose  90%,  and  of  cellote- 
traose  75%  by  0.5  mM  (Festenstein,  1959).  Glucono- 1,4-lactone  also  inhibits 
a  yeast  debranching  isoamylase  (Gunja  et  al.,  1961)  and  a  mammalian 
thioglycosidase  (Goodman  et  al.,  1959),  so  that  this  type  of  inhibition  ap- 
pears to  be  widespread. 

Very  potent  and  specific  inhibitions  are  exerted  on  N-acetyl-/?-glucos- 
aminidase  and  N-acetyl-/?-galactosaminidase  by  the  corresponding  lactones 
(Marsh  and  Lewy,  1957;  Findlay  et  al,  1958).  The  former  enzyme  from 
epididymis  is  inhibited  by  N-acetylglucosaminolactone  competitively,  with 
K-  =  0.000072  mil/,  but  the  limpet  enzyme  is  less  sensitive,  the  K^  being 
0.027  mM.  The  N-acetyl-a-glucosaminidase,  on  the  other  hand,  is  inhibited 
much  less.  The  acetyl  group  is  essential  for  the  inhibition  and  cannot  be 
replaced  by  other  acyl  groups.  Certainly  the  use  of  the  corresponding  lac- 
tones for  specific  and  potent  inhibition  of  the  glycosidases  has  been  one  of 
the  most  successful  endeavors  in  the  application  of  analogs. 


PYRUVATE   METABOLISM 

The  usefulness  of  a  direct  and  specific  inhibitor  of  pyruvate  utilization 
is  obvious  and  the  natural  occurrence  of  certain  pyruvate  analogs  makes 
this  field  of  inhibitor  study  an  important  one,  but  relatively  little  work 
has  been  done.  Fluoropyruvate  will  be  taken  up  with  other  fluorinated 
compounds,  such  as  fluoroacetate,  in  a  separate  chapter.  Phenylpyruvate 
is  the  best  studied  of  the  other  pyruvate  analogs.  It  is  readily  formed  from 
phenylalanine  and  is  usually  decarboxylated  to  phenylacetate,  although 
some  may  be  reduced  to  phenyllactate.  In  phenylketonuria  (phenylpyruvic 
oligophrenia),  the  accumulation  of  phenylpyruvate  could  be  responsible  for 


430  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

some  of  the  disturbances  by  interfering  with  pyruvate  metabolism,  but  no 
work  has  been  done  on  the  metabolic  changes  in  tissues  resulting  from 
such  levels  of  the  analog.  Hydroxyphenylpyruvate  is  similarly  formed  from 
tyrosine. 


(.        ))—  CH2—  CH-  COO" 
Phenylalanine 


The  formation  of  acetoin  from  pyruvate  in  Streptococcus  fecalis  is  inhibited 
13%  by  1  mM  and  75%  by  10  mM  phenylpyruvate,  whereas  the  formation 
of  acetoin  from  acetoacetate  is  unaffected  by  10  mM  (Dolin  and  Gunsalus, 
1951).  Pyruvate  oxidase  is  inhibited  to  about  the  same  degree.  Several 
anaerobic  pyruvate  pathways  are  inhibited  by  phenylpyruvate  in  several 
bacteria  and  yeast,  including  the  formation  of  acetoin,  the  phosphoroclastic 
reaction,  and  decarboxylation,  whereas  the  oxidative  metabolism  of  pyru- 
vate is  not  so  readily  affected  (Watt  and  Werkman,  1954).  The  concentra- 
tions of  pyruvate  and  phenylpyruvate  used  (120  mM)  are  unfortunately 
too  high  to  be  physiologically  significant,  but  further  study  on  extracts 
of  Aerobacter  aerogenes  showed  competitive  inhibition  of  acetoin  formation 
with  K,„  =  123-197  mM,  and  K^  =  0.59-1.1  mM,  so  that  in  this  case  phen- 
ylpyruvate is  bound  to  the  enzyme  much  more  tightly  than  pyruvate. 
The  reduction  of  hydroxypyruvate  to  glycerate  by  glycerate  dehydrogenase 
from  spinach  is  inhibited  by  phenylpyruvate  (5%),  pyruvate  (34%),  and 
bromopyruvate  (52%)  at  10  mM  (Holzer  and  Holldorf,  1957).  These  inhi- 
bitions are  competitive  but  rather  weak. 

The  most  pertinent  study  with  respect  to  blocking  an  important  pyruvate 
pathway  is  that  of  Gale  (1961)  on  yeast  pyruvate  decarboxylase,  which 
reports  the  inhibitions  given  in  Table  2-24.  The  following  compounds  are 
inactive:  pyruvic  ethyl  ester,  oxalacetate,  propionate,  phenyllactate,  phen- 
ylalanine, acetamide,  oxamate,  and  oxalate.  One  might  infer  that  (1)  the 
C=0  group  is  necessary  for  inhibition  (reduction  or  substitution  abolishes 
activity),  and  (2)  the  COO"  group  is  necessary  for  strong  inhibition  (amides 
and  esters  inactive).  The  nature  of  the  R  group  in  R — CO — COO"  can  vary 
quite  widely  and  it  is  difficult  to  correlate  structure  with  activity;  for 
example,  it  is  surprising  that  ketomalonate  is  bound  so  well  and  chloro- 
pyruvate  relatively  poorly,  and  that  oxanilate  is  bound  so  very  weakly. 


PYRUVATE  METABOLISM 


431 


Table  2-24 
Inhibition  of  Yeast  Pyruvate  Decarboxylase  by  Analogs" 


Inhibitor 


Structure 


Concen- 
tration 
(mM) 


Inhibition 


Relative 
activity 


Glyoxylate 


o-Nitrophenyl- 
pyruvate 


Ketomalonate 


/)-Hydroxyphe- 
nylpyruvate 


Phenylpyruvate 


Chloropyruvate 


2,  3-Butanedione 


O 

II 
H-C-COO 


i> 


CH, 


O 

II 

c-coo 


o 

II 

ooc-c-coo 


HO 


CHj—  C-COO' 


O 

II 

CH,—  C-COO 


O 

II 
CI— CH,— C-COO 


O     O 

II      II 
HoC     C     C     CHo 


0.023 

33 

0.45 

95 

0.09 

61 

0.14 

70 

0.23 

71 

0.34 

82 

0.45 

62 

1.1 

71 

1.1 

68 

2.3 

80 

0.45 

27 

2.3 

73 

32 


17 


12 


2.9 


56 


1.0 


0.14 


Oxanilate 


a-Ketoglutarate 


O 

II 
NH- C-COO' 


OOC-CH,— CH,— C— COO 


27 


31 


24 


0.05 


0.012 


o  Pyruvate  concentration  4.5  mW  and  preincubation  with  inhibitor  15  mln.    Relative  inhibitory 
activity  calculated  from  the  formula  t/(l  -  i)(I)  and  roughly  would  be  inversely  proportional 
to  Kf  for  noncompetitive  Inhibition.    (From  Gale,  1961.) 


The  inhibition  by  ketomalonate  is  more  surprising  when  one  considers  that 
oxalate  and  oxalacetate  are  inactive.  The  kinetics  of  the  inhibitions  pro- 
duced by  the  five  most  potent  substances  were  studied  in  greater  detail 
and  a  typical  noncompetitive  mechanism  was  established.  However,  pre- 
sence of  the  substrate  prevents  development  of  the  inhibitions,  suggesting 
that  the  inhibitors  combine  at  the  substrate  site.  These  results  indicate  an 
irreversible  or  pseudoirreversible  type  of  inhibition,  and  it  was  indeed  de- 
monstrated that  the  inhibitions  are  all  progressive  with  time.  Phenylpy- 
ruvate is  the  only  inhibitor  whose  effects  are  even  partially  reversible  by 


432  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

dialysis.  It  was  considered  that  the  true  inhibitors  are  the  aldehydes  cor- 
responding to  the  keto  acids,  and  it  was  shown  that  COg  is  evolved  from 
phenylpyruvate  and  p-hydroxyphenylpyruvate  but  not  from  the  other 
three  inhibitors.  Although  this  does  not  completely  invalidate  the  aldehyde 
proposal,  it  makes  it  unlikely.  On  the  other  hand,  it  is  possible  that  the 
tight  complex  is  formed  with  some  intermediate  prior  to  decarboxylation. 
Hydroxypyruvate  is  decarboxylated  by  yeast  decarboxylase  at  a  rate 
1/76  that  of  pyruvate  and  strongly  inhibits  the  pyruvate  reaction  (Holzer 
et  al.,  1955  d).  It  was  thought  that  hydroxypyruvate  might  exert  some 
regulatory  function  on  pyruvate  metabolism  in  yeast.  Formaldehyde  and 
acetaldehyde  inhibit  pyruvate  decarboxylase  but  the  mechanism  may  not 
be  competitive  with  substrate  (Bauchop  and  Dawes,  1959).  The  oxidation 
of  pyruvate  in  rat  kidney  slices  is  inhibited  around  70%  by  20  mM  meso- 
tartrate,  but  not  at  all  by  d-  and  DL-tartrates  (after  correction  for  the  inhi- 
bition of  endogenous  respiration)  (Quastel  and  Scholefield,  1955).  This 
inhibition  is  completely  reversed  by  fumarate  and  malate.  However,  when 
mitochondria  were  examined  it  was  found  that  there  is  little  direct  effect  on 
pyruvate  oxidation  but  a  rather  strong  inhibition  of  or-ketoglutarate  oxida- 
tion, which  is  progressive  with  time.  In  the  presence  of  bicarbonate,  pyru- 
vate is  oxidized  and  this  is  inhibited  by  meso-tartrate.  It  was  assumed 
that  the  inhibition  is  upon  the  incorporation  of  CO.2,  and  it  was  stated  that 
meso-tartrate  is  a  specific  inhibitor  of  pyruvate  oxidation  in  slices  of  rat 
kidney  cortex,  a  conclusion  that  would  seem  to  be  unjustified  since  so  few 
systems  were  tested. 

LACTATE  METABOLISM 

A  specific  inhibitor  of  lactate  dehydrogenase  would  be  useful  in  studying 
the  effects  of  a  block  of  glycolytic  lactate  formation  and  for  determining 
the  role  of  this  enzyme  in  the  functioning  of  tissues.  One  of  the  most  in- 
teresting lactate  dehydrogenase  inhibitors  is  oxamate,  first  reported  by 
Hakala  et  al.  (1953),  who  stated  that  it  is  the  most  potent  inhibitor  of  many 
lactate  and  pyruvate  analogs  examined.  The  inhibition  is  competitive  with 

O 

+H3N— C— coo- 

Oxamate 

respect  to  pyruvate,*  noncompetitive  with  respect  to  lactate  and  NAD, 
and  uncompetitive  with  respect  to  NADH  (Papaconstantinou  and  Colo- 

*  Examination  of  the  double  reciprocal  plot  reveals  that  the  situation  is  not  purely 
competitive  but  mainly  so. 


LACTATE    METABOLISM  433 

wick,  1957;  Novoa  et  al.,  1959).  The  substrate  and  inhibitor  constants  have 
been  summarized  by  Papaconstantinou  and  Colo  wick  (1961  a),  as  shown 
in  the  accompanying  tabulation;  if  the  KJs,  are  dissociation  constants, 


Source 


A',„  (pyruvate)         K^  (oxamate) 
{mM)  (mif) 


Beef  heart  0.137  0.0374 

Ascites  carcinoma  0.212  0.0563 

Rabbit  muscle  0.302  0.10 


oxamate  is  bound  around  0.77  kcal/mole  more  tightly  than  pyruvate.  The 
difficulty  in  interpreting  K^  is  that  oxamate  complexes  with  the  apode- 
hydrogenase,  E,  with  E-NAD,  and  with  E-NADH,  the  dissociation  con- 
stants being  different  for  each  (Novoa  et  al.,  1959).  The  values  for  three 
different  pH's  are  shown  in  the  accompanying  tabulation.  Such  ternary 


K 

i  (mM)  at: 

Complex 

pH  6.40 

pH  8.45 

pH  9.70 

E-oxamate 

0.10 

0.78 

20 

E-NAD-oxamate 

0.069 

0.57 

11 

E-NADH-oxamate 

0.026 

0.17 

1.6 

complexes  have  recently  been  demonstrated  by  ultracentrifugal  separation, 
1  mole  of  oxamate  being  bound  for  each  mole  of  NAD  or  NADH  (Novoa 
and  Schwert,  1961).  The  dissociation  constant  for  the  E-NADH-oxamate 
complex  determined  ultracentrifu gaily  is  0.011  mM  at  pH  7.4.  No  evidence 
could  be  found  for  a  complex  with  the  apoenzyme  alone;  whether  such  a 
complex  occurs  or  not,  the  inhibition  is  mainly  due  to  the  ternary  complexes. 
The  lactate  dehydrogenase  from  human  liver  and  heart  is  also  inhibited  by 
oxamate  (Plummer  and  Wilkinson,  1963).  The  reduction  of  2-ketobutyrate 
is  inhibited  more  strongly  than  pyruvate  reduction,  presumably  because 
2-ketobutyrate  is  bound  to  the  enzyme  less  tightly,  but  the  reduction  of 
succinic  semialdehyde  by  a  rat  brain  lactate  dehydrogenase  is  inhibited 
to  the  same  degree  as  the  reduction  of  pyruvate,  namely,  88%  by  0.1  mM 
(Fishbein  and  Bessman,  1964).  Not  all  lactate  dehydrogenases  are  sensitive 
to  oxamate;  that  from  L.  mesenteroides  is  inhibited  only  50%  by  7  mM 
(Papaconstantinou  and  Colowick,  1961  a).  Oxamate  specifically  inhibits  the 
L(  +  )-lactate  dehydrogenase  and  does  not  affect  d(  — )-lactate  dehydroge- 


^- 


434  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

nase  up  to  8  raM  (Dennis  and  Kaplan,  1960).  The  active  site  of  the  l(  +  )- 
lactate  dehydrogenase  was  represented  as  containing  a  cationic  group  for 
electrostatic  interaction  with  the  C00~  group,  and  a  hydrogen  bonding 
group  which  interacts  with  the  OH  group  of  lactate  and  the  NHg  group  of 
oxamate  (it  might  also  bond  to  the  CO  or  enolized  COH  group  of  oxamate). 
The  role  of  the  amino  group  of  oxamate  on  the  binding  is  not  known,  nor 
can  the  importance  of  the  enolic  tautomer  of  oxamate  be  evaluated.  The 
marked  decrease  in  inhibition  between  pH  8.45  and  9.70  (see  tabulation 
above)  could  indicate  the  deprotonation  of  an  amino  group,  but  it  could 
be  on  the  enzyme  as  well  as  the  inhibitor. 

The  aerobic  lactate  production  in  human  leucocytes  is  inhibited  less  than 
25%  by  10  mM  oxamate,  even  in  broken  cell  suspensions,  suggesting  that 
lactate  dehydrogenase  is  not  solely  responsible  for  lactate  formation,  al- 
though the  sensitivity  of  the  leucocytic  enz^^me  to  oxamate  has  not  been 
examined  (McKinney  et  al.,  1955).  The  effects  of  oxamate  on  tumor  cell 
metabolism  and  growth  have  been  studied  thoroughly  by  Papaconstantinou 
and  Colowick  (1957,  1961  a,  b).  Anaerobic  glycolysis  in  ascites  carcinoma 
cells  is  inhibited  50%  by  8  mM  oxamate  and  aerobic  glycolysis  is  similarly 
depressed.  This  may  indicate  that  oxamate  does  not  penetrate  into  cells 
readily,  since  the  K^  of  0.0563  mM  for  ascites  cell  lactate  dehydrogenase 
would  lead  one  to  expect  a  greater  effect  at  this  concentration.  The  inhi- 
bition of  anaerobic  glycolysis  decreases  with  time  due  to  the  accumulation 
of  pyruvate,  whereas  no  accumulation  of  pyruvate  occurs  aerobically,  in- 
dicating that  oxamate  has  little  effect  on  pyruvate  oxidase  (an  18%  inhi- 
bition of  pyruvate  oxidation  by  10  mM  oxamate  was  observed).  A  decrease 
in  the  inhibition  anaerobically  with  time  was  also  noted  in  Tetrahymena 
pyriformis  (Warnock  and  van  Eys,  1963).  The  growth  of  HeLa  cells  is 
completely  inhibited  by  40-80  mM  oxamate  and  this  is  paralleled  by  de- 
creases in  glucose  uptake  and  lactate  formation,  so  that  lactate  dehydro- 
genase appears  in  some  manner  to  be  essential  for  the  growth  of  these  cells 
(assuming  that  oxamate  acts  specifically  on  lactate  dehydrogenase).  It  was 
proposed  that  oxamate  might  be  a  useful  inhibitor  for  selectively  blocking 
glycolysis  in  mammalian  cells.  Mice  can  tolerate  quite  large  doses  (1  g/kg), 
however.  A  block  of  glycolysis,  of  course,  refers  here  only  to  an  inhibition 
of  lactate  formation,  and  the  formation  or  utilization  of  pyruvate  should 
not  be  significantly  affected,  so  it  would  seem  that  aerobic  glucose  meta- 
bolism, at  least  with  respect  to  the  generation  of  energy,  would  be  resistant 
to  oxamate. 

A  number  of  disturbing  observations  have  appeared  which  cast  some 
doubt  on  the  simple  concept  that  oxamate  specifically  inhibits  lactate  de- 
hydrogenase. Leached  HeLa  cells  restored  to  normal  conditions  actively 
extrude  Na+  and  accumulate  K+;  these  processes  are  inhibited  50%  and 
77%,  respectively,  by  38  mM  oxamate  (Wickson-Ginzburg  and  Solomon, 


LACTATE    METABOLISM  435 

1963).  Inasmuch  as  the  conditions  were  aerobic  here,  it  is  difficult  to  un- 
derstand how  an  inhibition  of  lactate  dehydrogenase  would  account  for  the 
marked  effects  observed,  unless  there  is  an  elevation  of  the  NADH/NAD 
ratio  due  to  the  prevention  of  pyruvate  reduction,  this  slowing  the  oxidation 
of  glyceraldehyde-3-P  and  reducing  the  generation  of  ATP.  This  does  not 
seem  to  occur  in  Ehrlich  ascites  carcinoma  cells  inasmuch  as  oxamate  stim- 
ulates the  formation  of  C^^Og  from  glucose-6-C^*  without  affecting  that 
from  glucose- 1-C^'*  (Christensen  and  Wick,  1963).  The  stimulation  is  thus 
associated  only  with  oxidation  through  the  cycle  and  there  is  no  effect  on 
the  fraction  going  through  the  pentose-P  pathway.  More  pyruvate  enters 
the  cycle  since  less  goes  to  lactate.  The  effects  of  oxamate  on  glucose  utiliza- 
tion will  depend  for  one  thing  on  how  rapidly  pyruvate  can  be  oxidized. 
The  Crabtree  effect  is  abolished  almost  completely  by  40  uiM  oxamate,  i.e., 
in  the  presence  of  glucose,  oxamate  stimulates  the  respiration  in  ascites 
cells  (Papaconstantinou  and  Colowick,  1961  a).  Simultaneously,  glucose  up- 
take is  depressed  40%  and  lactate  formation  70%.  In  view  of  the  conclusion 
about  the  nature  of  the  Crabtree  effect  in  the  section  on  2-DG,  it  would 
seem  that  inhibition  of  lactate  dehydrogenase  could  not  be  responsible  for 
this  effect  of  oxamate.  It  is  possible  that  oxamate  diverts  more  pyruvate 
into  the  cycle  and  hence  stimulates  respiration  under  these  conditions,  but 
this  would  not  be  a  true  abolition  of  the  Crabtree  effect.  One  would  not 
expect  glucose  respiration  to  be  depressed  by  oxamate  in  any  case  if  the 
only  action  is  on  lactate  dehydrogenase,  but  31%  respiratory  depression 
is  produced  by  10  mM  oxamate  in  guinea  pig  alveolar  macrophages  (Oren 
et  at.,  1963).  Despite  the  statements  relative  to  the  specificity  of  oxamate, 
it  must  be  admitted  that  very  few  enzymes  or  metabolic  pathways  have 
been  studied.  It  is  quite  likely  that  certain  phases  of  amino  acid  metabolism 
might  also  be  inhibited,  since  oxamate  could  be  considered  as  an  amino 
acid  analog.  It  will  also  be  noted  that  all  the  effects  discussed  in  this  para- 
graph were  produced  by  oxamate  at  the  high  concentration  of  10  raM 
or  above. 

Oxalate  often  inhibits  lactate-metabolizing  enzymes  as  potently  as  does 
oxamate  and  similarly  forms  ternary  complexes  with  lactate  dehydrogenase 
and  NAD.  However,  it  differs  from  oxamate  in  being  competitive  with 
lactate  instead  of  pyruvate;  this  has  been  shown  on  beef  heart  lactate 
dehydrogenase  (^,  =  0.015  mM  at  pH  6.7)  (Novoa  et  al,  1959),  yeast 
D-lactate  dehydrogenase  (^,  =  0.007)  (Labeyrie  and  Stachiewicz,  1961), 
yeast  D-lactate  cytochrome  c  reductase  {K^  =  0.0016  mM)  (Nygaard,  1961 
b),  and  yeast  D-hydroxy  acid  dehydrogenase  (/if,  =  0.0025  mM)  (Boeri  et 
al.,  1960),  although  the  inhibition  seems  to  be  uncompetitive  on  the  lactate 
dehydrogenase  oi  Propionihacterium  pentosaceum  (Molinari  and  Lara,  1960). 
The  complex  kinetics  have  been  treated  in  detail  by  Novoa  et  al.  (1959), 
but  it  is  still  rather  puzzling  that  oxalate  inhibits  all  these  enzymes  so 


436  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

potently  and  competes  with  lactate  rather  than  pyruvate.  Tartronate  and 
malonate  also  inhibit  competitively  with  lactate  and  noncompetitively  with 
respect  to  pyruvate,  although  by  no  means  as  strongly  as  oxalate,  and 
Ottolenghi  and  Denstedt  (1958)  concluded  that  pyruvate  and  lactate  react 
with  different  sites  on  the  enzyme  surface,  but  this  is  not  necessary,  as 
Novoa  et  al.  (1959)  have  shown,  since  the  configurations  of  the  active  sites 
on  E-NAD  and  E-NADH  are  different.  The  five  lactate  dehydrogenase 
(LD)  isoenzymes  from  human  tissues  are  inhibited  to  different  degrees  by 
oxalate  at  0.02  vaM  (see  accompanying  tabulation)  (Emerson  et  al.,  1964). 


Isoenzyme 

0/ 

/o 

Inhibition 

LDi 

70 

LD^ 

64 

LD3 

56 

LD4 

46 

LD, 

32 

There  are  A  and  B  monomers,  the  B  monomer  being  more  sensitive  to 
oxalate.  LD^  is  a  pure  B  tetramer,  LD5  is  a  pure  A  tetramer,  and  the  others 
are  intermediate  in  the  proportion  of  A  to  B.  This  is  an  illuminating  example 
of  how  enzymes  may  respond  differently  to  inhibitors  by  reason  of  varied 
composition. 

There  are  several  interesting  studies  on  lactate  dehydrogenases  from 
which  deductions  on  the  nature  of  the  active  site  and  certain  interaction 
energies  may  be  derived.  Dikstein  (1959)  found  that  yeast  lactate  dehydro- 
genase is  not  inhibited  potently  by  monocarboxylates:  the  concentrations 
for  50%  inhibition  are  6000  vaM  for  formate,  2300  mM  for  acetate,  800  milf 
for  propionate,  300  mM  for  butyrate,  120  mM  for  valerate,  and  10  mM  for 
caprylate.  By  plotting  log  (1)50  against  the  number  of  carbon  atoms  in  the 
aliphatic  chains  he  obtained  a  straight  line,  from  the  slope  of  which  it  was 
possible  to  calculate  that  the  transfer  of  a  methylene  group  from  the  solvent 
to  the  enzyme  surface  involves  an  over-all  energy  change  of  0.5  kcal/mole. 
It  was  assumed  that  the  interaction  energy  between  the  COO"  group  and 
an  enzyme  cationic  group  could  be  calculated  from  the  intercept  of  this 
line,  but  it  is  doubtful  that  at  the  high  concentrations  used  the  residual 
energy  can  be  attributed  with  certainty  to  ion-ion  interactions,  since  non- 
specific effects  cannot  be  excluded. 

The  D-of-hydroxy  acid  dehydrogenase  of  yeast  is  competitively  inhibited 
by  several  dicarboxylates  (see  accompanying  tabulation)  (Cremona,  1964). 
Monocarboxylates  inhibit  much  more  weakly  and  the  inhibitions  are  usually 
not  purely  competitive.  Oxalate  is  bound  approximately  3.7  kcal/mole  more 


LACTATE    METABOLISM  437 

tightly  than  malonate,  suggesting  that  there  are  two  cationic  groups  quite 
close  on  the  enzyme. 


Inhibitor  Ki  (mil/) 


Oxalate 

0.0025 

Tartronate 

0.84 

Malonate 

0.95 

L-Malate 

1.05 

a-Ketoglutarate 

1.4 

The  D-lactate  dehydrogenase  of  yeast  studied  by  Boeri  et  al.  (1960)  is 
inhibited  potently  by  oxalate  {K^  =  0.0025  mM),  moderately  by  malonate 
(^ .  =  0.9  mM),  and  not  at  all  by  16  mM  succinate  or  fumarate,  indicating 
the  importance  of  the  position  of  the  second  C00~  group.  The  enzyme  is 
inactivated  gradually  by  EDTA  and  reactivated  with  Zn++.  Although  this 
does  not  prove  that  Zn++  is  the  normal  metal  ion  involved,  it  points  to  the 
possibility  of  chelation  between  a-hydroxy  carboxylates,  or  dicarboxylates, 
with  an  enzyme-bound  metal  ion.  The  configuration  around  the  a-carbon 
atom  is  important  since  L-lactate  binds  very  weakly  to  the  enzyme  {K^  = 
=  62  mM). 

The  lactate  cytochrome  c  reductases  of  yeast  are  flavoproteins  and  are 
competitively  inhibited  by  a  variety  of  lactate  analogs  (Nygaard,  1961  a, 
b,  c).  Fatty  acid  inhibitions  lead  to  the  calculation  of  0.37  kcal/mole  for 
the  interaction  of  methylene  groups  with  the  enzyme  and  2.80  kcal/mole 
for  the  C00~  group  in  the  case  of  the  D-lactate  cytochrome  c  reductase, 
and  of  0.37  kcal/mole  and  0.88  kcal/mole,  respectively,  for  the  L-lactate 
cytochrome  c  reductase.  Neither  eftzyme  is  inhibited  by  dicarboxylates, 
with  the  exception  of  oxalate,  and  from  the  differences  in  inhibition  patterns 
it  is  probably  safe  to  assume  that  the  active  sites  of  the  two  enzymes  are 
quite  different. 

Some  analog  inhibitors  of  lactate  dehydrogenases  need  only  be  listed  for 
reference:  pyruvate  (Green  and  Brosteaux,  1936;  Das,  1937  b;  Neilands, 
1952;  Labeyrie  and  Stachiewicz,  1961),  bromopyruvate,  chloropyruvate, 
hydroxypyruvate  (Busch  and  Nair,  1957),  phenylpyruvate  (Dikstein,  1959), 
a-ketoglutarate  (Boeri  et  al.,  1960),  malate  (Boeri  et  al.,  1960;  Busch  and 
Nair,  1957),  tartronate  (Green  and  Brosteaux,  1936;  Lehmann,  1938),  tar- 
trate (Labeyrie  and  Stachiewicz,  1961),  mandelate  (Lehmann,  1938),  a, 
y-diketovalerate  (Meister,  1950),  benzenesuLfonate  (Baptist  and  Vestling, 
1957),  mercaptoacetate,  mercaptosuccinate,  or-mercaptopropionate,  of-mer- 
captobutyrate,  and  a-mercaptovalerate  (Chaffee  and  Bartlett,  1960).  The 
benzenesulfonates,  where  the  COO"  group  is  replaced  by  a  SO3"  group. 


438  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

and  the  mercapto  fatty  acids,  where  the  a-OH  is  replaced  by  an  a-SH 
group,  are  particularly  interesting  and  deserve  further  study  to  determine 
their  specificity  on  lactate  metabolism. 

The  oxidation  of  glycolate  is  catalyzed  by  glycolate  oxidase  and  the 

CHO-CHO-  +  NADH  +  H+  ;fi  CH^OH-COO-  +  NAD+ 
Glyoxylate  Glycolate 

reverse  reaction  by  glyoxylate  reductase,  both  enzymes  being  found  in 
plants.  This  reaction  is  similar  to  the  pyruvate  ±5  lactate  interconversion 
and,  indeed,  L-lactate  is  slowly  oxidized  by  the  oxidase.  Since  these  enzymes 
bear  some  relationship  to  those  involved  in  lactate  metabolism  due  to  this 
similarity,  it  is  not  inappropriate  to  discuss  them  at  this  point,  particularly 
as  Zelitch  at  the  Connecticut  Agricultural  Experiment  Station  has  reported 
some  very  interesting  inhibitions  by  analogs.  Glyoxylate  reductase  is  ap- 
parently not  especially  susceptible  to  analogs:  the  following  inhibitions  were 
observed  with  16.5  mM  glyoxylate  and  10  mM  analogs  —  phenylglyoxy- 
late  21%,  oxalacetate  22%,  oxamate  31%,  and  pyruvate  56%  (Zelitch, 
1955).  Glycolate  oxidase,  on  the  other  hand,  is  well  inhibited  by  a-hydrox- 
ysulfonates.  Competitive  inhibition  was  found  with  hydroxymethanesul- 
fonate  {K,  =  0.0018  mM),  a-hydroxyethanesulfonate  (^,;  =  0.0023  mM), 
and  sulfoglycolate  {K^  =  0.0021  mM);  since  K,,^  =  0.38  mM,  these  analogs 
are  reasonably  potent  (Zelitch,  1957).  Rabbit  muscle  lactate  dehydrogenase 
is  inhibited  comparably,  D-glycerate  dehydrogenase  less  strongly,  and  ma- 
late  dehydrogenase  not  at  all,  so  that  some  specificity  toward  enzymes 
oxidizing  a-hydroxy  acids  is  evident.  The  inhibition  of  lactate  oxidation 
by  the  glycolate  oxidase  is  inhibited  very  strongly  because  lactate  is  bound 
less  tightly  to  the  enzyme.  Another  competitive  inhibitor  of  comparable 
potency  is  of-hydroxy-2-pyridinemethanesulfonate  (Zelitch,  1959)  which 
has  been  used  in  most  of  the  in  vivo  work  apparently  because  it  is  more 
effective  in  cells,  although  sulfoglycolate  w^ould  seem  to  act  very  similarly 
(Zelitch,  1958). 

OH 

I 

CH3— CH— SO3  H0-CH2— so; 

Q-Hydroxyethanesulfonate  HydroxymethanesuKonate 


9"  ^N       OH 

I  /^  \       I 

'OOC  — CH-SO;  \\      A>— CH-SO3 

Sulfoglycolate  ■^^'^^^!'u''^'K       . 

pyridinemethanesulfonate 


PHOSPHATASES  439 

Normal  tobacco  leaves  contain  around  0.5-1  //mole/g  wet  weight  glyco- 
late  and  this  level  can  be  increased  as  much  as  10-fold  by  placing  them  in 
solutions  containing  the  a-hydroxysuLfonates  (Zelitch,  1959).  Since  glyco- 
late  is  formed  photosyntheticaUy,  marked  accumulation  occurs  only  in  the 
light.  The  concentration  of  a-hydroxy-2-pyridinemethanesulfonate  for  max- 
imal inhibition  is  near  10  m.M,  indicating  that  penetration  into  the  cells 
is  quite  poor.  At  higher  concentrations  of  inhibitor  the  glycolate  level  falls, 
due  presumably  to  inhibition  of  glycolate  formation;  in  fact,  even  at  10  mM 
there  must  be  some  inhibition  since  photosynthetic  incorporation  of  C^Og 
is  inhibited  about  33%.  The  pattern  of  C^*  distribution  is,  however,  more 
markedly  altered;  in  controls,  glycolate-C^^  accounts  for  5.5%  of  the  labeling 
but  in  inhibited  leaves  it  is  almost  50%.  This  is  a  good  example  of  a  spe- 
cific analog  inhibitor  useful  in  studying  the  importance  of  an  intermediate 
in  a  complex  metabolic  pathway,  and  valuable  information  may  result 
from  more  detailed  studies  on  photosynthesis. 


PHOSPHATASES 

These  enzjines  are  commonly  inhibited  by  the  products  of  the  hydroly- 
sis. We  have  already  noted  the  inhibitions  of  phospho-L-histidinol  phos- 
phatase, O-phosphoserine  phosphatase  (page  270),  and  glucose-6-phos- 
phatase  (page  442)  by  dephosphorylated  products,  and  there  are  other 
examples  related  to  analog  inhibition.  Orthophosphate  also  frequently  in- 
hibits: The  following  may  be  cited  as  instances  of  well-marked  inhibition 
of  different  types  of  enzyme  —  calf  intestinal  phosphatase  (Schmidt  and 
Thannhauser,  1943),  mouse  liver  acid  phosphatase  (Macdonald,  1961), 
mouse  liver  pyrophosphatase  (Rafter,  1958),  calf  brain  carbamyl  and  acyl 
phosphatases  (Grisolia  et  al.,  1958),  sweet  potato  phosphatase  (Ito  et  al., 
1955),  and  E.  coli  alkaline  phosphatase  (Garen  and  Levinthal,  1960).  An- 
other class  of  inhibitor  comprises  the  phosphates  that  are  substrates.  Thus 
sweet  potato  phosphatase  with  phenyl  phosphate  as  substrate  is  competi- 
tively inhibited  by  /^-glycerophosphate  {Kf  =  2  mM),  pyrophosphate  (K,  = 
=  0.33  mM),  metaphosphate  (^,  =  3.2  mM),  and  ATP  {K^  =  0.67  mM), 
all  of  which  are  also  substrates  (Ito  et  al.,  1955).  Likewise  the  E.  coli 
phosphatase  with  p-nitrophenyl  phosphate  as  substrate  is  competitively 
inhibited  by  uridine  phosphate  {K^  =  0.044  raM),  guanosine  phosphate 
{K^  =  0.046  mM),  /^-glycerophosphate  {K^  =  0.05  vnM),  glucose- 1 -phos- 
phate {K^  =  0.063  mM),  and  adenosine-5'-phosphate  (Z,;  =  0.093  mM) 
(Garen  and  Levinthal,  1960). 

More  interesting  are  the  inhibitions  by  various  anions  that  may  be  con- 
sidered as  analogs  of  either  phosphate  or  the  substrate  phosphates.  Thus 
arsenate  (Garen  and  Levinthal,  1960;  Ito  et  al.,  1955;  Macdonald,  1961), 
borate  (Ito  et  al.,  1955),  and  silicate  (Umemura  et  al.,  1961)  inhibit  various 


440  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

phosphatases  about  as  potently  as  phosphate  and  probably  combine  with 
the  enzymes  in  a  similar  manner.  Certain  carboxylates  (oxalate,  malonate, 
malate,  citrate,  glucarate,  gluconate,  lactate,  and  others)  have  been  found 
to  be  inhibitory,  but  most  of  these  are  not  remarkably  effective,  probably 
binding  to  enzyme  cationic  groups  to  various  degrees  or  chelating  with 
metal  ions.  However,  ( +  )-tartrate*  is  a  very  potent  inhibitor  of  certain 
phosphatases,  as  first  shown  by  Abul-Fadl  and  King  (1949)  and  confirmed 
by  Anagnostopoulos  (1953),  and  is  so  much  more  active  than  other  anions 
that  the  mechanism  has  been  investigated  in  several  excellent  studies. 

The  phylogenetic  relationships  of  (  +  )-tartrate  inhibition  and  the  vari- 
able susceptibilities  of  phosphatases  from  different  tissues  are  quite  inter- 
esting. Abul-Fadl  and  King  (1949)  found  that  although  prostatic  acid  phos- 
phatase is  very  sensitive,  no  effects  are  exerted  on  the  acid  phosphatases  of 
erythrocytes  or  plasma,  and  Anagnostopoulos  (1953)  noted  no  inhibition 
with  the  phosphatases  of  mustard,  wheat  germ,  or  Aspergillus.  Kilsheimer 
and  Axelrod  (1958)  investigated  the  effects  of  (  +  )-tartrate  on  the  phospha- 
tases from  many  sources  and  found  that  at  20  mM  the  inhibitions  vary 
from  0  to  93%.  They  concluded  that  animal  phosphatases  are  more  suscep- 
tible than  plant  phosphatases,  and  that  bacterial  phosphatases  are  generally 
resistant.  It  was  suggested  that  (  +  )-tartrate  may  be  of  taxonomic  use  in 
those  more  primitive  organisms  where  it  is  difficult  to  decide  whether  they 
are  plants  or  animals. 

The  earliest  work  demonstrated  that  the  inhibition  is  stereospecific,  nei- 
ther ( —  )-tartrate  nor  meso-tartrate  exerting  appreciable  effects  on  prostatic 
phosphatase,  and  this  has  been  confirmed  in  all  the  recent  studies.  In  the 
series  of  phosphatases  tested  by  Kilsheimer  and  Axelrod  (1958)  it  was  ob- 
served that  very  few  of  the  enzymes  are  affected  by  (— )-tartrate  and 
these  are  inhibited  only  slightly.  The  iiT/s  were  determined  by  London  et 
at.  (1958)  as  0.13  mM  for  (  +  )-tartrate  and  93  mlf  for  (-)-tartrate  on 
prostatic  acid  phosphatase.  It  is  strange  that  they  found  meso-tartrate  to 
be  an  effective  inhibitor  {K^  =  0.4  mM)  in  contrast  to  all  other  work. 
The  acid  phosphatase  from  Neurospora  crassa  is  inhibited  completely  by 
12  mM  (  +  )-tartrate  but  is  unaffected  by  50  mM  (  — )-tartrate  or  meso- 
tartrate  (Kuo  and  Blumenthal,  1961).  In  all  cases  the  inhibition  by  (  +  )- 
tartrate  is  competitive  with  respect  to  substrate.  The  inhibition  is  thus  more 
marked  with  /^-glycerophosphate  as  the  substrate  than  with  phenyl  phos- 
phate or  p-nitrophenyl  phosphate,  since  the  former  substrate  is  bound  less 
tightly  (Nigam  at  al,  1959;  Kuo  and  Blumenthal,  1961). 

Tartrate  and  related  inhibitors  have  been  used  to  map  the  active  site 
of  prostatic  phosphatase  (London  et  al.,  1958).  The  K/s  and  estimated 

*  There  has  been  confusion  in  the  nomenclature  of  the  tartrates  and  the  dextro- 
rotatory (  +  )-tartrate  has  been  designated  as  l-  or  D-,  depending  on  the  system  used. 
To  avoid  ambiguity  I  shall  indicate  the  isomers  by  the  signs  of  their  rotation. 


PHOSPHATASES  441 

relative  binding  energies  are  shown  in  Table  2-25.  It  was  assumed  that  ( +  )- 
tartrate  is  bound  to  the  enzyme  at  four  points;  the  two  carboxylate  groups 
interact  with  two  enzyme  cationic  groups  and  the  two  hydroxyl  groups 

Table  2-25 
Competitive  Anionic  Inhibitors   or  Prostatic  Phosphatase" 


Inhibitor 


Fluoride  dimer  (HF^ 

( +  )-Tartrate 

DL-Glycerate 

meso-Tartrate 

Arsenate 

D-Malate 

Sulfamate 

Nitrate 

D- Alanine 

L-Leucine 

Diphenylphosphate 

D-Lactate 

L-Serine 

L-Glutamate 

L-Aspartate 

( —  )-Tartrate 

L- Lactate 

L-Malate 

L- Alanine 

Glycine 


Ki 

Relative   —  AF  oi  binding 

{mM) 

(kcal/mole) 

<0.1 

>5.68 

0.13 

5.51 

0.2 

5.25 

0.4 

4.82 

1 

4.25 

1.8" 

3.90 

8 

2.98 

16 

2.55 

27 

2.23 

30 

2.16 

32 

2.12 

50 

1.85 

50 

1.85 

60 

1.74 

64 

1.69 

93 

1.47 

110 

1.36 

280 

0.79 

470 

0.47 

470 

0.47 

"  The  substrate  is  /S-glycerophosphate  {K,^  —  16  mM).  Experiments  with  amino 
acids  at  pH  7.2  and  with  the  rest  at  pH  5.  (From  London  et  al.,  1958.) 

"  Kj  for  D-malate  estimated  from  value  for  DL-malate  since  L-malate  is  a  weak  in- 
hibitor. 

form  hydrogen  bonds  with  the  enzyme  (Fig.  2-14).  On  the  basis  of  this 
model  it  is  possible  to  explain  most  of  the  variations  in  inhibitor  binding. 
(  — )-Tartrate  can  make  contact  at  only  two  points  while  meso-tartrate  can 
make  a  three-point  attachment  in  two  ways.  Only  D-malate  can  attach  at 
three  points  like  (  +  ) -tartrate  but  the  second  hydroxyl  group  is  missing. 
The  most  important  requirement  for  binding  is  a  negative  group  separated 
from  the  nucleophilic  group  by  around  2.9  A.  (-f  )-Tartrate  has  two  such 
units.  In  the  substrates  the  oxygen  atom  of  phosphate  is  the  nucleophilic 
atom.  The  amino  acids  are  rather  poor  inhibitors,  probably  because  at 


442 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


pH  7.2  most  of  the  amino  groups  are  protonated  and  interfere  with  the 
binding.  HFj"  may  bind  across  two  attachment  points  and  its  potency  may 
be  related  to  the  strong  hydrogen  bonds  formed  by  fluorine.  The  nitrate 


Fig.  2-14.  Scheme  for  the  active  site  of 
prostatic  acid  phosphatase  showing  two 
cationic  groups  and  two  hydrogen -bonding 
groups  separated  by  a  seam  in  the  enzyme. 
The  unmarked  circles  represent  various 
enzyme  groups  and  the  stippled  area 
between  them  a  lipophilic  region.  (Modi- 
fied from  London  et  al.,  1958.) 

ion  also  can  bridge  two  of  the  attachment  points,  one  oxygen  atom  being 
negatively  charged  and  the  other  nucleophilic. 

Some  of  the  results  of  Nigam  et  al.  (1959),  shown  in  the  accompanying 
tabulation,  differ  from  those  of  London  et  al.  (1958).  These  inhibitions  were 


Inhibitor  (10  mikf ) 

%  Inhibition 

( -f  )-Tartrate 

100 

Glucarate 

91 

Pyruvate 

80 

Oxalate 

72 

Malonate 

66 

Maleate 

45 

Glutamate 

42 

Glucuronate 

40 

Lactate 

0 

SULFATASES  443 

observed  with  3  mM  /^-glycerophosphate  as  substrate.  The  potency  of  (-(-)- 
tartrate  is  not  evident  in  the  tabulation  since  50%  inhibition  is  produced 
by  0.07  mM.  The  failure  of  lactate  to  inhibit  is  surprising,  especially  as 
pyruvate  is  fairly  potent.  The  susceptibility  to  glucarate  observed  here  has 
not  been  observed  with  Neurospora  phosphatase,  12.5  mM  having  no  effect 
(Kuo  and  Blumenthal,  1961 ),  and  Jeffree  (1957)  found  only  a  moderate  inhi- 
bition {K^=  10  mM)  on  prostatic  phosphatase,  it  being  definitely  less  potent 
than  oxalate  in  contrast  to  the  results  above.  Whether  the  lactone  plays  a  role 
in  this  inhibition  is  not  known.  Two  anionic  polymers  have  been  found  to 
be  more  potent  inhibitors  than  (-f) -tartrate:  polyxenyl  phosphate  inhibits 
the  Neurospora  phosphatase  completely  at  0.16  mM  (Kuo  and  Blumenthal, 
1961),  and  alginate  (556  residues)  inhibits  the  prostatic  enzyme  with  a  K^ 
of  0.0054  mM  (Jeffree,  1957).  These  polymer  inhibitions  are  only  partially 
competitive. 

SULFATASES 

Arylsulfatases  of  type  II  (liver  enzymes  A  and  B  and  molluscan  enzymes) 
are  generally  inhibited  by  sulfate  and  phosphate,  but  type  I  arylsulfatases 
(liver  enzyme  C  and  bacterial  enzymes)  are  resistant.  The  product  inhibi- 
tion by  sulfate  was  first  reported  by  Tanaka  (1938)  and  shown  to  be  com- 
petitive with  the  substrate  nitrocatechol  sulfate  by  Roy  (1953).  The  sus- 
ceptibilities of  ox  liver  arylsulfatases  to  sulfate  vary  markedly:  The  K^  for 
sulfatase  A  is  0.75  mM,  for  sulfatase  B  is  70  mM,  and  sulfatase  C  is  not 
affected  (Roy,  1953,  1954  b,  1956).  The  sulfatase  A  is  inhibited  also  by 
several  organic  sulfates,  although  not  very  potently  (see  accompanying  tab- 
ulation). The  following  do  not  inhibit  at  25-50  mM:  methyl  sulfate,  glu- 
cose-3-suLfate,  glucose-6-sulfate,  and  phenyl  sulfite.  The  fact  that  methyl 
sulfate  does  not  inhibit  and  the  affinity  for  the  enzyme  increases  with  size 
of  the  ring  system  esterified  might  indicate  that  the  sulfate  group  is  not 


Inhibitor 

{mM) 

Relative   —  AF  o{  binding 
(kcal/mole) 

Sulfite 

0.002 

8.07 

Phosphate 

<0.21 

>5.21 

Sulfate 

0.75 

4.42 

2-Phenanthryl  sulfate 

A 

3.40 

Phenyl  phosphate 

.^.3 

3.23 

2-Naphthyl  sulfate         / 

30 

2.15 

1-Naphthyl  sulfate 

40 

1.98 

m-Tolyl  sulfate 

100 

1.42 

Phenyl  sulfate 

300 

0.74 

Benzyl  sulfate 

300 

0.74 

444  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

involved  in  the  binding  of  these  substances.  The  inhibition  by  these  organic 
sulfates  was  stated  to  be  noncompetitive  but  since  some  are  slowly  hydro- 
lyzed  substrates,  the  inhibition  may  be  mixed.  Phosphate  seems  generally 
to  be  a  more  potent  inhibitor  than  sulfate;  this  has  been  observed  on  the 
arylsulfatases  of  rabbit  liver  (Maengwyn-Davies  and  Friedenwald,  1954), 
ox  liver  (Webb  and  Morrow,  1959),  Helix  pomatia  (Dodgson  and  Powell, 
1959),  Charonia  lampas  (Takahashi,  1960  a),  and  beef  and  rabbit  cornea 
(Wortman,  1962).  The  inhibitions,  where  examined,  are  competitive.  Weak 
inhibitions  by  xylenesulfonate  (Maengwyn-Davies  and  Friedenwald,  1954) 
and  benzenesulfonate  (Dodgson  et  al.,  1955;  Dodgson  and  Powell,  1959) 
have  been  observed. 

The  potent  inhibition  by  sulfite  is  interesting  but  the  mechanism  is  not 
yet  understood,  although  it  would  appear  to  be  competitive  (Roy,  1953). 
For  all  the  ox  liver  arylsulfatases,  sulfite  is  bound  3-4  kcal/mole  more  tightly 
than  is  sulfate;  indeed,  sulfatase  C,  although  resistant  to  sulfate,  is  inhibited 
55%  by  0.1  mM  sulfite  (Roy,  1956).  Sulfite  is  also  more  inhibitory  than 
sulfate  on  sulfatases  from  bacteria  and  molluscs  (Dodgson  et  al.,  1955;  Dodg- 
son and  Powell,  1959;  Dodgson,  1959),  although  in  most  cases  the  concen- 
trations used  were  too  high  to  evaluate  the  potency  readily. 

Choline  sulfatase  of  Pseitdomonas  nitroredvcences  is  inhibited  very  strongly 
by  sulfite  (33%  by  0.01  mM  and  83%  by  0.1  mM  when  choline  sulfate  is 
10  mM),  but  is  not  affected  or  somewhat  stimulated  by  sulfate  and  phos- 
phate (Takebe,  1961).  The  steroid  sulfatase  and  glucosulfatase  of  Patella 
vulgata  are  inhibited  by  sulfate  and  even  more  potently  by  phosphate 
(Roy,  1954  a).  Since  the  sulfatases  comprise  so  heterogeneous  a  group  of 
enzymes  and  relatively  few  have  been  adequately  studied,  it  is  difficult  to 
draw  general  conclusions  or  make  valid  correlations,  but  it  is  at  least  evident 
that  analogs  are  occasionally  effective  inhibitors.  It  is  also  likely  that  phos- 
phate must  exert  a  regulatory  action  on  sulfatase  activity  in  vivo. 


ADENOSINETRIPHOSPHATASES 
AND   TRANSPHOSPHORYLASES 

Enzymes  hydrolyzing  ATP  or  transferring  its  terminal  phosphate  to 
various  acceptors  are  frequently  inhibited  by  other  nucleotides.  Competi- 
tive product  inhibition  by  ADP  has  been  noted  for  ATPases  from  several 
sources;  the  inhibition  is  never  marked,  since  ADP  is  usually  bound  some- 
what less  tightly  than  ATP  to  the  enzyme,  but  is  sufficient  to  slow  progres- 
sively the  rate  of  ATPase  reactions.  The  ATP-P^  and  ATP- ADP  exchange 
reactions  catalyzed  by  mitochondria  and  digitonin  particles  are  also  inhib- 
ited by  ADP  (Low  et  al.,  1958;  Cooper  and  Kulka,  1961).  Some  results  with 
ADP  and  other  related  substances  are  shown  in  Table  2-26.  The  inhibitions 
however,  depend  on  the  pH  and  the  concentrations  of  Ca++  and  Mg++,  as 


ADENOSINETRIPHOSPHATASES ,  TRANSPHOSPHORYLASES 


445 


Table  2-26 
Inhibition  of  ATPases  by  Analogs 


ATP 

Analog 

Source               ,     , . 

centration 

Analog 

con- 
centration 

0/ 

T  ,.,.,.                  Reference 
Inhibition 

(mM) 

(milf) 

Liver 

4 

ADP 

2 

23 

Kielley  and  KieUey 

mitochondria 

4 

42 

(1953) 

IDP 

2.2 
4 

0 
16 

ITP 

4 

13 

15 

AMP 

15 

22 

Low  et  al.  (1958) 

ADP 

15 

45 

3.5 

ADP 

3.5 

50 

Cooper  and   Kulka 
(1961) 

Muscle  myosin 

1.07 

ADP 

0.4 
2.4 

7 
12 

Blum  (1955) 

0.21 

ADP 

2.4 

29 

Spinach 

4 

AMP 

4 

37 

VVessels  and 

chloroplasts 

ADP 

4 

29 

Baltscheffsky  (1960) 

Brain 

— 

Adenine 

5 

23 

Gore  (1951) 

Adenosine 

20 

16 

Guanine 

5 

2 

shown  by  Green  and  Mommaerts  (1954).  Addition  of  Ca++  decreases  ADP 
binding  at  pH  6.4  and  increases  it  at  pH  9,  whereas  Mg++  has  no  effect  at 
the  lower  pH  but  decreases  affinity  at  the  higher  pH.  The  K^  for  ADP  is 
around  0.13  roM  in  the  absence  of  Ca++  and  Mg++,  but  around  0.5  mM 
at  pH  6.4  and  40  mM  Ca++.  Kielley  and  Kielley  (1953)  had  shown  with 
liver  mitochondrial  ATPase  that  ADP  does  not  alter  the  optimal  Mg++ 
concentration  for  ATPase  activity,  indicating  that  the  inhibition  is  not  by 
the  binding  of  Mg"*""*",  but  Nanninga  (1958)  reported  that  part  of  the  inhi- 
bition of  myosin  ATPase  by  ADP  is  due  to  chelation  of  Ca++.  In  the  pre- 
sence of  excess  Ca++  this  chelation  can  be  neglected  and  the  true  K^  for 
the  enzyme- ADP  complex  is  found  to  be  4.6  mM  at  pH  7. 

An  interesting  phosphonic  analog  of  ATP,  adenylmethylenediphospho- 
nate: 

0  0  0 

II  II  II 

Adenine-ribose— O— P— O— P— CH— P— O- 

O-  O-  OH 


446  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

was  found  by  Moos  et  al.  (1960)  to  be  unable  to  replace  ATP  in  contracting 
glycerinated  muscle  and  not  to  be  hydrolyzed  by  myosin  ATPase.  However, 
some  inhibition  on  ATPase  is  exerted  (the  affinities  of  the  enzyme  for  ATP 
and  its  analog  appear  to  be  roughly  the  same),  although  this  is  not  com- 
petitive. Mg++  is  able  to  overcome  the  inhibition  at  a  concentration  lower 
than  that  of  the  analog,  indicating  that  Mg++  is  not  simply  complexing 
with  and  removing  free  analog.  The  mechanism  of  the  inhibition  was  re- 
presented by  the  following  reactions: 

EMg 


Mg 

+ 

E 

+ 

+ 

I 

I 

w 

w 

Mgl 

EI 

where  E  is  low-activity  enzyme  and  EMg  is  high-activity  enzyme.  The  inhi- 
bition has  a  dual  basis:  (1)  removal  of  Mg++,  thus  decreasing  the  fraction 
of  the  enzyme  in  the  high-activity  form,  and  (2)  reaction  of  the  low-activity 
enzyme  directly  with  the  analog.  This  situation  may  be  fairly  common  in 
inhibitions  on  enzymes  with  activating  metal  ions. 

A  few  other  nucleotidase  inhibitions  may  be  mentioned.  ITPase  is  inhib- 
ited by  IDP  and  ADP  (Blum,  1955;  Kielley  and  Kielley,  1953).  Indeed, 
ADP  inhibits  ITPase  more  strongly  than  ATPase.  The  ITPase  of  fly  muscle 
is  strongly  inhibited  by  ADP  (K^  =  0.0165  mM)  and  much  less  readily  by 
IDP  (Kj  =  1.59  mM),  the  inhibition  being  competitive  at  low  but  noncom- 
petitive at  higher  concentrations  (Sacktor  and  Cochran,  1957).  GTPase  is 
likewise  inhibited  but  UTPase  is  unaffected  by  either  ADP  or  IDP.  In 
phage-infected  E.  coli  the  hydrolysis  of  deoxycytidine  diphosphate  (deoxy- 
CDP)  is  inhibited  by  deoxyCMP  and  deoxyCTP,  and  the  hydrolysis  of 
deoxyCTP  is  inhibited  by  deoxyCMP  and  deoxyCDP,  in  both  cases  the 
deoxyCMP  being  relatively  less  active  (Zimmerman  and  Romberg,  1961). 

AMP- ATP  transphosphorylase  (myokinase)  from  rabbit  muscle  is  inhi- 
bited by  ADP  {K^  =  0.33  mM)  and  this  is  competitive  with  respect  to 
both  AMP  and  ATP  (Noda,  1958).  The  reverse  reaction  from  2ADP->AMP 
-f  ATP  is  inhibited  by  AMP  {K,  =  0.5  mM)  and  ATP  (A',  =  0.32  mM), 
the  K/s  being  the  same  as  the  K„^s  for  these  substances  (Callaghan  and 
Weber,  1959).  A  much  more  effective  analog  is  adenosine  monosulfate 
{K^  =  0.0186  mM).  Creatine  kinase  is  inhibited  competitively  by  ADP 
(Ki  =  0.27  mM),  AMP  {K^  =  7  mM),  adenosine  (Z,  =  7  mM),  tripoly- 
phosphate  {K,  =  8  mM),  orthophosphate  (A,  =  13  mM),  sulfate  (A,  = 
6  mM),  and  nitrate  {K^  =  22  mM)  (Noda  et  al,  1960).  The  substrate  here 
is  MgATP=  and  it  is  possible  that  the  most  effective  inhibitors  form  Mg 
complexes.  Most  of  the  anions  inhibit  the  forward  reaction  competitively 
with  respect  to  MgATP=  and  the  reverse  reaction  competitively  with  re- 


HYDROXYSTEROID  DEHYDROGENASES 


447 


spect  to  creatine  phosphate  (Nihei  et  al.,  1961).  However,  ADP  competes 
with  MgADP"  in  the  reverse  reaction. 

Since  oxidative  phosphorylation  may  in  some  ways  be  related  to  ATPase 
activity  and  transphosphorylations,  it  is  not  out  of  place  to  discuss  the  ef- 
fects of  various  inorganic  phosphorus  compounds  on  this  process.  The  phos- 
phorylation in  a  rat  liver  mitochondrial  suspension  oxidizing  fumarate,  and 
with  glucose  and  hexokinase  to  trap  the  phosphate,  was  studied  by  Thom- 
son and  Sato  (1960)  (Table  2-27).  Some  of  the  analogs  investigated  reduce 
the  P  :  0  ratio  by  depressing  phosphate  uptake  more  than  oxygen  uptake, 
but  the  only  compound  that  can  be  considered  as  a  true  uncoupler  is  thio- 
phosphate.  In  such  a  complex  system  a  number  of  sites  for  inhibition  are 
evident.  It  is  known  that  anions  can  inhibit  fumarase  and  it  is  possible  that 
other  enzymes  attacking  dicarboxylates  might  be  inhibited.  Hexokinase  is 
inhibited  at  concentrations  interfering  with  oxidative  phosphorylation  only 
by  triphosphate.  Some  of  these  compounds  might  deplete  Mg++  but  it  was 
shown  that  thiophosphate  does  not.  It  is  not  known  if  any  of  these  sub- 
stances can  enter  into  the  phosphorylative  reaction  but  fail  to  form  ATP. 
Further  study  of  thiophosphate  would  seem  warranted  by  these  preliminary 
results. 

HYDROXYSTEROID  DEHYDROGENASES 

The  /?-hydroxysteroid  dehydrogenase  of  Pseudomonas  testosteroni  cata- 
lyzes the  oxidations  of  3/?-  and  17/?-hydroxysteroids  to  their  respective  ke- 
tones with  NAD  as  acceptor.  The  oxidations  of  testosterone  and  17/5-estra- 
diol  are  competitively  inhibited  by  17cif-estradiol  (Talalay  and  Dobson,  1953). 


H,C 


H,C 


OH 


Androstane 


Testosterone 


Androst-l-ene-3, 17-dione 


HO 


Estra-l,  3,  5-triene 


Estrone 


Progesterone 


448  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Table  2-27 
Effects  of  Inorganic  Puosi'HORUs  Compounds  on  Oxidative  Phosphorylation " 


Structu 

re 

Concen- 
tration 
(mA/) 

7, 

Change 

Inhibitor 

O3 

uptake 

Pi 
uptake 

P:0 

Pyrophosphate 

-O-P-O-P- 

o        o 

O' 

1.5 
3 

-  22 

-  20 

-  36 

-  51 

-  16 

-  38 

Triphosphate 

o"       o' 

O—P-O— I'- 
ll         II 
o        o 

o'       o' 

1                        1 

0- 

0 

1 
-P-O 

II 
0 

0.6 
3 

+    8 
-  30 

-  11 

-  48 

-  17 

-  25 

Pyrophosphite 

"O-P-O-P- 

1          1 

H            H 

o' 

1 

-     4 

-     4 

0 

Diphosphite 

O'    O' 

1       1 

0-P-P=0 

II    1 

0     H 
?'    ?' 

3 
15 

-  12 

-  11 

-  4 

-  18 

f  10 
-  16 

Hypophosphate 

1       1 
O-P— P-O 

a  5 

1 

-  25 

-  38 

-  20 

Isohypophos- 
phate 

'O— P— O— P- 

II       II 

o        o 

-H 

3 
15 

t^  10 
-  25 

-     4 

-  15 

Trimetaphos- 
phimate 

O^    /O' 
p 

o^  1          1  /O' 

1 

H 

o'       o' 

0' 

3 

+  12 

+    8 

-     4 

Diimidotri- 
phosphate 

1              1 
"O— P— N— P- 

Tl     1     II 

OHO 

-t 

t 
-P-O 

11 
0 

3 

-     9 

-  24 

-  19 

0.6 

-     4 

-     2 

+    2 

Thiophosphate 

PSO33- 

3 
5 

+  18 
+  26 

0 
-  28 

-  15 

-  42 

15 

-  59 

-  91 

-  78 

"Rat  liver  mitochondria  oxidizing  fumarate  with  phosphate  at  15  mAf.    (From  Thomson  and 
Sato,  1960.) 


HYDROXYSTEROID  DEHYDROGENASES  449 

This  observation  was  extended  to  a  number  of  estra-l,3,5-trienes,  most  of 
which  are  inhibitory  to  the  oxidation  of  testosterone  and  are  not  themselves 
attacked  (Marcus  and  Talalay,  1955).  The  most  potent  inhibitors  are  the 
following  derivatives  of  estratriene:  -3,17a-diol  (a-estradiol),  -3,16«-diol; 
-3,16a,17/5-triol  (estriol);  -3-ol;  -3,17/?-diol-16-one;  and  -3,16/3-diol.  It  ap- 
pears that  the  aromaticity  of  ring  A  combined  with  the  3-OH  group  re- 
sults in  strong  binding.  The  17-ols  are  not  inhibitory  although  there  are 
enzyme  regions  for  the  oxidation  of  either  3-OH  or  17-OH  groups  in  other 
ring  systems.  The  total  ring  system  is  not  necessary  since  dieth/lstilbestrol 
and  hexestrol  are  potent  inhibitors.  The  K/s  for  most  of  the  effective  in- 
hibitors are  around  0.001-0.01  mM  corresponding  to  over-all  interaction 
energies  of  7-8.5  kcal/mole,  implying  a  rather  close  fit  over  the  surface  of 
the  molecules  and  a  summation  of  dispersion  and  polarization  forces.  It  is 
possible  that  the  high  polarizability  of  the  aromatic  ring  A  is  important  in 
augmenting  binding,  this  ring  overlying  some  ionic  group  on  the  enzyme, 
and  the  3-OH  interacts  to  form  a  hydrogen  bond.  The  a,^  specificity  indi- 
cates that  the  steroids  attach  to  the  enzyme  by  their  "rear"  surfaces.  The 
cf-hydroxysteroid  dehydrogenase  is  not  inhibited  so  readily  by  the  estra- 
trienes  as  is  the  (3  enzyme  (Talalay  and  Marcus,  1956).  It  was  stated  that 
the  inhibitions  are  neither  exactly  competitive  nor  noncompetitive,  but  no 
data  or  plots  were  given,  nor  was  the  exact  experimental  procedure  des- 
cribed, so  that  it  is  impossible  evaluate  the  nature  of  the  inhibitions. 

The  most  potent  inhibitor  of  the  Pseudomonas  /?-hydroxysteroid  dehydro- 
genase yet  found  is  the  noncompetitive  2-hydroxymethylene-17a-methylan- 
drostan-17/5-ol-3-one  {K,  =  0.0003  mM),  although  the  competitive  4,4-di- 
methyl-17/?-hydroxyandrost-5-eno(3,2-c)pyrazole  {K^  =  0.0005  mM)  is  al- 
most as  active  (Ferrari  and  Arnold,  1963  a,  b).  The  inhibitions  by  these  and 
simpler  steroids  are  dependent  on  the  pH;  e.g..  diethylstilbestrol  is  approxi- 
mately 20  times  as  effective  at  pH  8.5  than  at  pH  5.5.  Since  it  is  unlikely 
that  the  phenolic  groups  would  ionize  in  this  range  (pK^  for  diethylstilbes- 
trol is  12.2),  this  implies  the  ionization  of  enzyme  groups  at  or  near  the 
active  site.  This  emphasizes  the  importance  of  polarization  of  the  aromatic 
rings  by  anionic  groups  on  the  enzyme  in  determining  the  tightness  of 
binding. 

The  J^-  and  J*-dehydrogenases  which  introduce  unsaturation  into  ring 
A  of  the  3-ketosteroids  at  the  1-  and  4-positions,  respectively,  are  inducible 
enzymes  in  P.  testosteroni,  and  both  are  inhibited  quite  potently  by  estrone 
(Levy  and  Talalay,  1959).  The  J*-3-ketosteroid  reductase  (5a)  of  rat  liver 
microsomes,  which  catalyzes  the  hydrogenation  of  the  4-5  double  bond,  is 
a  NADPH-requiring  enzyme  acting  on  cortisone,  Cortisol,  desoxycorticoste- 
rone,  and  related  steroids  (McGuire  et  al.,  1960).  It  is  competitively  inhi- 
bited by  a  variety  of  less  substituted  steroids,  such  as  androst-l-ene-3,17- 
dione  and  5a-androstane-3,17-dione;  the  5/?-androstane-3,17-dione  is,  how- 


450 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


ever,  inactive,  indicating  very  specific  attachment  to  the  enzyme.  The  K^ 
for  17/5-hydroxyandrosta-l,4-diene-3-one  is  0.146  mM  when  the  substrate 
is  cortisone  (iiC,„  =  0.14  xaM),  and  the  other  inhibitors  presumably  have 
similar  values  for  K^.  It  may  also  be  noted  that  6of-  and  6y5-methyl-ll-keto- 
progesterone  are  inhibitors. 

The  normal  roles  of  these  enzymes  in  bacteria  or  mammals  are  not  as 
yet  well  understood,  although  the  steroid  dehydrogenases  in  Pseudomonas 
enable  this  organism  to  grow  with  the  appropriate  steroids  as  the  sole 
source  of  carbon.  There  are  many  other  enzymes  involved  in  steroid  syn- 
thesis and  catabolism  which  have  not  been  studied  with  respect  to  analog 
inhibition.  When  such  studies  are  made  it  may  well  be  found  that  this  is 
an  important  regulating  mechanism  in  controlling  the  levels  of  the  steroid 
hormones. 

NITRITE  AND  SULFITE  METABOLISM 

Nitrobacter  is  a  soil  autotroph  that  can  obtain  its  energy  from  the  oxida- 
tion of  nitrite  to  nitrate.  Chlorate  at  low  concentrations  (0.01-0.1  mM) 
inhibits  growth  without  affecting  nitrite  oxidation,  whereas  at  higher  con- 
centrations (above  1  mM)  nitrite  oxidation  is  progressively  depressed  (Lees 
and  Simpson,  1955).  The  inhibition  is  completely  reversible  and  it  was  pos- 
tulated that  chlorate  combines  with  a  nitrite-oxidizing  enzyme  at  a  stage 
when  it  is  bound  to  nitrate;  however,  the  inhibition  seems  to  be  unrelated 
to  the  concentrations  of  either  nitrite  or  nitrate,  and  is  not  reversed  by 
increasing  the  nitrite  concentration  (Lees  and  Simpson,  1957).  It  was  then 
postulated  that  chlorate  does  not  inhibit  directly  but  is  first  converted  to 
some  substance,  perhaps  chlorite,  which  inhibits  rapidly  and  irreversibly. 


Nitrite 


o 

0" 

11- 
1     V  \-     ■ 

0/ V 

'o-ci^— 

Nitrate 

Chlorate 

Chlorite 

Inhibitor  (2  mM) 

0/ 

/o 

Inhibition 

CIO3- 

80 

BrOs 

8 

IO3- 

1 

PO3- 

0 

o 


N=C=0 
'N— C=0^ 
'^N=C— O' 

Cyanate 


Related  anions  do  not  inhibit  comparably  with  chlorate  (see  tabulation, 
where  nitrite  is  8mM),  but  cyanate  is  apparently  even  more  inhibitory 


NITRITE    AND    SULFITE    METABOLISM  451 

(Butt  and  Lees,  1960).  However,  the  effect  of  cyanate  is  markedly  depen- 
dent on  the  oxygen  concentration  (see  following  tabulation),  so  that  at 
low  oxygen  levels  inhibition  is  reversed  to  stimulation.   It  was  suggested 


Cyanate 

% 

Change 

(mikf) 

0^  =  2.5% 

O2  =  20% 

0.2 

+30 

-50 

0.3 

+45 

-66 

0.5 

+  18 

-76 

0.7 

-  3 

-82 

1.0 

-46 

-92 

that  a  membrane  transport  system  brings  nitrite  into  the  cells,  and  at  mod- 
erate nitrite  concentrations  and  normal  oxygen  pressures  the  rates  of 
transport  and  oxidation  are  comparable.  At  low  oxygen  concentrations 
there  is  accumulation  of  nitrite  in  the  cells  with  consequent  inhibition  of  the 
nitrite-oxidizing  enzyme  (since  substrate  inhibition  occurs  when  nitrite  is 
above  1  vaM).  Cyanate  may  interfere  with  the  transport  so  that  at  normal 
oxygen  concentrations  the  accumulation  of  nitrite  is  suppressed  and  the 
substrate  inhibition  released.  It  would  be  interesting  to  know  what  effect, 
if  any,  chlorate  has  on  the  nitrite  transport. 

The  oxidation  of  sulfite  by  the  liver  is  inhibited  by  thiosulfate  and  this 
action  is  probably  exerted  on  sulfite  oxidase  (Fridovich  and  Handler,  1954, 
1956).  Thiosulfate  inhibits  both  aerobically  and  anaerobically  (methylene 
blue  reduction).  Although  it  was  originally  stated  that  the  inhibition  is 
competitive,  it  has  been  found  more  recently  (MacLeod  et  al.,  1961)  that 
it  is  not,  at  least  with  respect  to  sulfite.  The  following  reaction  scheme  was 
suggested: 


-S 

-S  +  HSO3 


-S— H 


-S— H 


(+  H2O) 

-s— SO3- >  h-s— H  +  SOr  +  H^ 


and  it  was  proposed  that  thiosulfate  competes  with  the  active  thiosulfonate 
in  the  hydrolytic  step  of  the  sequence.  Methanesulfonate,  ethanesulfonate, 
benzenesulfonate,  and  pyridine-3-sulfonate  are  noninhibitory. 

Wild-type  Neurospora  crassa  can  utilize  sulfate  as  the  sole  source  of  sulfur 
but  certain  mutants  require  more  reduced  forms  for  growth  (Ragland  and 
Liverman,  1958).  Some  strains  can  use  sulfite  and  others  only  thiosulfate. 
Since  sulfate  inhibits  competitively  the  utilization  of  thiosulfate  but  does 
not  interfere  with  the  utilization  of  sulfite,  and  inasmuch  as  previously  it 


452  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

had  been  postulated  that  sulfite  must  pass  through  thiosulfate  to  be  utiliz- 
ed, a  shunt  around  thiosulfate  was  proposed: 


I  I 

Active  sulfate  ->  active  sulfite  ->  active  thiosulfate  -►  growth 

t  t  t 

sulfate  sulfite  thiosulfate 

Sulfate  would  thus  block  the  formation  of  active  thiosulfate  from  exogenous 
thiosulfate,  or  it  could  block  the  utilization  of  active  thiosulfate.  Another 
possibility  is  that  thiosulfate  transport  into  the  cell  is  depressed  by  sulfate, 
in  which  case  there  is  no  necessity  to  assume  a  shunt  as  above. 


SIMPLE    ION    ANTAGONISMS 

Many  examples  of  the  inhibition  of  biological,  metabolic,  or  enzymic 
processes  by  simple  inorganic  ions  are  known  and  in  certain  instances  it  is 
likely  that  the  mechanism  involves  a  competition  between  the  inhibiting 
ion  and  a  necessary  ionic  cofactor.  The  inhibiting  ions  may  be  considered 
as  analogs  of  the  normally  functional  ions.  It  is  impossible  here  to  treat 
this  subject  completely  but  it  may  be  worthwhile  to  mention  briefly  some 
specific  examples  of  competitive  enzyme  inhibition  to  illustrate  the  phe- 
nomenon. MacLeod  and  Snell  (1950)  emphasized  the  possible  importance 
of  such  competitions  in  the  effects  of  certain  ions  on  the  growth  of  bacteria. 
The  growth  of  Lactobacillus  arabinosus  is  suppressed  by  various  ions  and  it 
was  concluded  that  some  of  these  effects  are  due  to  the  fact  that  the  inhib- 
itory ions  are  structural  analogs  of  the  metal  ions  which  are  cofactors  in 
metabolism.  The  following  antagonisms,  among  others,  were  demonstrated: 
K+  reverses  the  inhibitory  effects  of  Na+  and  NH4+;  K+  reverses  the  in- 
hibitory effects  of  Rb+;  Zn++  reverses  the  inhibitory  effects  of  Mn++;  and 
Mg++,  Ca++,  and  Sr++  reverse  the  inhibitory  effects  of  Zn++.  Actually  it  is 
not  certain  that  these  antagonisms  relate  to  metabolic  events,  since  it  is 
possible  that  physiological  ions  play  nonmetabolic  roles  in  bacteria,  and  it 
would  be  interesting  to  investigate  these  antagonisms  on  a  metabohc  level. 
Inasmuch  as  so  many  enzymes  either  involve  ionic  cofactors  or  are  affected 
by  physiological  ions,  it  is  likely  that  such  competitions  are  very  common 
and  important  in  metabolic  regulation. 

The  phosphotransacetylase  of  Clostridium  kluyveri  is  active  only  in  the 
presence  of  K+  or  NH4+  ions.  Inhibition  is  exerted  by  Na+  and  Li+  ions, 
and  this  can  be  overcome  by  increasing  the  K+  or  NH4+  concentration,  so 
that  competition  for  an  enzyme  site  would  seem  likely  (Stadtman,  1955). 
Yeast  acid  phosphatase  requires  Mg++  and  is  inhibited  by  Ca++,  the  inhi- 
bition being  strictly  competitive,  as  shown  by  double  reciprocal  plots  (Tsuboi 
and  Hudson,  1956).  Ba++  does  not  inhibit  at  all,  but  Mn++  and  Zn++  depress 


INHIBITION   BY    MACROIONS  453 

the  activity,  although  less  than  Ca++,  and  this  may  be  competitive  also. 
Ca++  is  a  competitive  inhibitor  of  rabbit  muscle  phosphorylase  6  kinase 
with  respect  to  the  activator  Mg++,  but  is  noncompetitive  with  respect  to 
ATP  {K,„  =  1.9  mM  for  Mg++,  and  K^  =  0.3  vaM  for  Ca++)  (Krebs  et  al, 
1959).  Most  ATPase  are  activated  by  Mg++  and  some  are  inhibited  by  Ca++, 
but  L-myosin  ATPase  is  activated  by  Ca++  and  inhibited  by  Mg++;  in  both 
cases  competition  may  occur.  One  might  speculate  that  the  uncoupling  of 
oxidative  phosphorylation  by  Ca++  may  involve  displacement  of  other  ions 
such  as  Mg++  or  Mn++.  The  question  of  metal  ion  competition  will  be  taken 
up  in  greater  detail  in  the  chapters  on  inhibitions  produced  by  Zn++,  Pb++, 
Cd++,  and  other  metal  cations.  Some  new  ideas  on  the  regulatory  role  of 
simple  cations  in  metabolism,  especially  glycolysis,  may  be  found  in  the 
interesting  discussion  of  Wyatt  (1964).  One  example  of  ion  inhibition  in  an 
enzyme  system  in  which  an  ion  is  the  substrate  will  be  mentioned.  This  is 
the  formation  of  ^-chlorolevulinate  from  /5-ketoadipate  by  an  enzyme,  /?- 
ketoadipate  chlorinase,  from  Caldariomyces  fumago  (Shaw  and  Hager,  1959). 
F~,  Br-,  and  I"  inhibit  this  reaction  around  85%  when  they  are  present  in 
equimolar  concentration  with  Ch  (10  mM).  The  nature  of  the  inhibition 
is  not  known  but  it  is  rather  surprising  that  the  three  inhibiting  ions  are  of 
the  same  degree  of  potency. 


INHIBITION  BY  MACROIONS 

The  inhibition  by  high  molecular  weight  substances  of  enzymes  attacking 
related  high  molecular  weight  substrates  may  be  thought  of  as  analog  inhi- 
bition in  certain  instances,  especially  where  competitive  kinetics  has  been 
demonstrated  (the  type  of  inhibition  has  seldom  been  studied  in  work  on 
polymers).  On  the  other  hand,  these  inhibitions  are  probably  often  nonspe- 
cific in  the  sense  that  the  polymers  interact  with  any  or  all  regions  of  the 
enzyme  surface  rather  than  just  the  active  center.  Spensley  and  Rogers 
(1954)  reviewed  inhibitions  of  this  type  and  suggested  the  terms  macrocat- 
ionic  and  macroanionic  as  applicable  to  effects  exerted  by  positively  charged 
and  negatively  charged  polymers,  respectively.  Early  work  was  mainly  with 
naturally  occurring  macroions,  such  as  heparin  or  protamine,  but  the  pre- 
paration and  use  of  more  homogeneous  and  physically  characterized  syn- 
thetic poljoners  and  copolymers  have  enabled  the  nature  of  the  interactions 
to  be  better  understood.  Some  of  these  inhibitions  may  well  be  physiologi- 
cally important,  for  example,  the  intracellular  effects  of  the  various  types  of 
nucleic  acids  on  enzymes  involved  in  nucleic  acid  metabolism  or  their  role 
in  protein  synthesis.  The  suggestion  by  Jones  and  Gutfreund  (1964),  that 
certain  enzymes  participating  in  metabolic  sequences  may  interact  speci- 
fically with  each  other  to  form  complexes  facilitating  flow  along  the  path- 
way, brings  up  the  possibility  that  macroions  can  interfere  in  this  interaction 


454  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

and  thereby  depress  metabolism  by  a  mechanism  unrelated  to  direct  action 
on  the  individual  enzymes.  General  macroion  inhibition  will  be  briefly  dis- 
cussed in  this  section,  whether  specific  or  nonspecific,  and  that  these  are  all 
instances  of  analog  inhibition  is  not  implied.  Bernfeld  (1963)  has  recently 
reviewed  some  aspects  of  macroanionic  inhibition. 

General    Nature  of  the   Interactions   between    Macroions 

Inasmuch  as  enzymes  are  macroions,  the  interactions  involved,  whether 
specific  or  nonspecific,  will  be  mainly  electrostatic  and  will  depend  primarily 
on  the  total  charges  and  the  distributions  of  charges  on  the  enzymes  and 
the  inhibitory  polymers.  It  is  likely  that  enzymes  at  pH's  removed  from 
their  isoelectric  points  will  interact  to  varying  extents  with  macroions  of 
opposite  net  charge  irrespective  of  whether  the  enzymes  attack  high  mo- 
lecular weight  or  low  molecular  weight  substrates,  but  we  shall  direct  our 
attention  chiefly  to  enzymes  whose  substrates  are  polymeric.  When  inhi- 
bition occurs  under  conditions  in  which  the  enzyme  and  the  macroion  are 
carrying  the  same  net  charge,  it  has  often  been  assumed  that  other  than 
electrostatic  interactions  are  involved,  but  this  is  not  necessary.  Let  us 
picture  an  enzyme  of  isoelectric  point  pH  7  in  a  reaction  medium  of  pH  6.5. 
The  net  charge  on  the  enzyme  will  be  positive,  but  interspersed  with  the 
cationic  groups  on  the  enzyme  surface  there  will  be  many  anionic  carboxy- 
late  groups  in  most  instances.  If  a  macrocationic  substance  can  orient  itself 
on  the  enzyme  surface  so  as  to  react  with  these  anionic  groups,  perhaps 
because  of  complementary  spacing  of  the  oppositely  charged  groups,  inhi- 
bition may  result,  although  it  will  undoubtedly  be  less  than  at  pH's  above 
the  isoelectric  point.  This  is  not  said  to  minimize  the  importance  of  nonelec- 
trostatic  interactions,  which  certainly  must  occur,  particularly  when  the 
inhibitory  macroions  contain  groups  capable  of  forming  hydrogen  bonds  or 
regions  contributing  to  the  binding  through  van  der  Waals'  forces. 

The  question  of  the  specificity  of  macroionic  inhibition  is  a  difficult  one 
and  the  data  available  do  not  allow  us  to  draw  general  conclusions.  It  has 
been  established,  however,  that  a  particular  enzyme  will  be  inhibited  quite 
differently  by  various  macroions  of  the  opposite  net  charge,  and  that  a 
particular  macroion  exerts  very  different  effects  on  a  group  of  enzymes. 
There  is  probably  sufficient  evidence,  to  be  discussed  later,  that  enzymes 
acting  on  macroionic  substrates  are  likely  to  be  inhibited  rather  strongly 
by  other  macroions  in  which  the  charge  distribution  is  similar  to  that  of 
the  substrate,  and  in  such  cases  the  inhibition  may  indeed  be  competitive. 
A  quantitative  treatment  of  the  interactions  must  be  on  a  statistical  basis 
and,  as  far  as  I  know,  this  has  not  been  undertaken,  and  would  indeed  be 
very  difficult  since  there  are  several  factors  about  which  there  is  inadequate 
information. 

The  rate  at  which  macroionic  inhibition  develops  has  not  been  studied 


INHIBITION    BY    MACROIONS  455 

extensively  but  one  can  imagine  some  interesting  phenomena  in  this  con- 
nection. The  initial  binding  of  the  macroion  to  the  enzyme  would  not  be 
expected  to  be  that  for  maximal  interaction  because  the  first  contact  be- 
tween them  would  be  random.  Inasmuch  as  most  macroionic  inhibitions  are 
readily  reversible,  it  is  likely  that  the  polymer  would  move  on  the  enzyme 
surface  until  near  maximal  or  maximal  interaction  occurs.  The  inhibition 
for  this  reason  might  increase  with  time,  the  rate  depending  on  several 
factors,  such  as  the  binding  affinity  of  the  individual  group  interactions 
and  the  flexibility  of  the  polymer,  and  progressive  developments  of  inhi- 
bition have  been  experimentally  observed.  Inasmuch  as  many  configura- 
tions of  the  macroion  on  the  enzyme  surface  may  be  characterized  by  inter- 
action energies  very  near  the  maximal,  it  is  likely  that  even  at  equilibrium 
each  enzyme  molecule  will  not  be  inhibited  to  the  same  degree,  especially 
for  those  enzymes  acting  on  small  substrates,  since  the  inhibition  will  usual- 
ly be  due  to  a  steric  interference  by  a  polymer  chain  passing  over  or  near 
an  active  site. 

Factors  Determining  the  Degree  of  Inhibition 

The  most  important  factors  relating  to  the  inhibitory  macroion  would  be 
(1)  the  molecular  weight  or  polymer  length,  (2)  the  density  of  ionic  groups 
on  the  polymer,  or  the  repeat  distance  between  them,  (3)  the  over-all  con- 
figuration of  the  polymer,  i.e.,  linear,  branched,  or  globular,  and  (4)  the 
flexibility  of  the  polymer.  The  last  factor  is  perhaps  very  important  but 
has  been  generally  ignored.  If  one  assumes  an  approximately  globular  en- 
zyme, the  net  binding  energy  and  the  degree  of  inhibition  may  well  depend 
on  the  ability  of  the  polymer  to  conform  to  the  enzyme  surface,  specifically 
to  wrap  around  it  so  that  interactions  between  many  ionic  groups  can  take 
place.  Some  of  the  ionic  polysaccharides  must  not  be  too  flexible  and  this 
may  limit  the  effects  they  have  on  certain  enzjTnes,  while  the  synthetic 
macroions  vary  in  flexibility  over  a  wide  range.  Entropy  factors  must  be 
very  significant  in  the  binding  of  macroions,  and  would  to  a  great  extent 
depend  on  the  deviation  of  the  bound  polymer  from  its  statistical  configu- 
ration in  solution.  A  polysaccharide  or  polypeptide  macroion  of  molecular 
weight  10,000  would  contain  roughly  35  units  and  the  total  extended  length 
would  be  200-250  A  if  linear.  Such  a  macroion  might  be  able  to  encircle 
an  average  enzyme  1  or  2  times  or,  if  it  is  randomly  distributed  over  the 
enzyme  surface,  would  cover  very  roughly  about  10-15%  of  the  enzyme. 
Many  macroions  used  to  inhibit  enzymes  are,  of  course,  much  larger,  often 
being  of  molecular  weights  of  100,000  or  over. 

Two  characteristics  of  the  media  used  in  inhibition  studies  are  particu- 
larly important,  namely,  the  pH  and  the  ionic  strength,  since  the  interac- 
tions between  enzyme  and  macroion  are  mainly  of  the  ion- ion  type.  The 
pH  will  determine  the  net  charge  on  the  enzyme  and  occasionally  the  ioni- 


456  2,  ANALOGS  or  enzyme  reaction  components 

zation  state  of  groups  with  which  the  inhibitor  reacts,  and  in  all  cases  in 
which  this  has  been  studied  a  marked  dependence  on  the  pH  has  been 
demonstrated.  An  increase  in  the  ionic  strength  should  reduce  such  inhibi- 
tions because  of  the  competition  of  the  small  ions  for  the  enzyme  and  ma- 
croion  groups,  and  this  has  been  repeatedly  confirmed  experimentally  (data 
for  the  inhibition  of  trypsin  by  polyglutamate  are  given  in  Table  1-15-6). 
Hydration  of  the  ionic  groups  must  also  be  a  significant  factor  in  reducing 
the  inhibitions  from  what  might  be  expected  on  the  basis  of  interactions 
in  a  vacuum,  so  that  anything  which  modified  the  extent  of  hydration  of 
either  enzyme  or  macroion  might  secondarily  affect  the  inhibition.  A  final 
factor  which  can  markedly  reduce  such  inhibitions  is  the  presence  of  ma- 
croionic  impurities  in  the  preparation  if  one  is  not  working  with  pure  en- 
zymes. It  has  been  shown  many  times  that  inhibitions  by  macroions  can 
be  prevented  or  actually  reversed  by  other  macroions  of  opposite  charge  to 
the  inhibitor.  Such  results  are  not  particularly  significant  since  they  imply 
only  that  the  inhibitor  can  also  bind  to  nonenzyme  macroions,  a  fact  which 
can  be  better  demonstrated  with  other  techniques,  but  they  emphasize  the 
possible  importance  of  such  impurities  in  the  studies  on  macroionic  inhi- 
bitions. 

Trypsin  and  Chymotrypsin 

The  isoelectric  point  of  trypsin  is  close  to  pH  11  and  that  of  casein  is 
between  4  and  4.5;  thus  the  hydrolysis  of  casein  involves  the  interaction 
of  a  macrocationic  enzyme  with  a  macroanionic  substrate  at  pH  values  near 
neutrality.  Since  heparin,  a  strongly  negatively  charged  sulfated  polysac- 
charide, was  known  to  form  complexes  with  positively  charged  proteins, 
Horwitt  (1940)  examined  its  action  on  trypsin  and  found  a  rather  potent 
inhibition  at  pH  7.3.  Inhibition  does  not  occur  unless  the  enzyme  is  incubat- 
ed with  heparin  before  the  addition  of  the  casein,  possibly  indicating  a 
competitive  type  of  interaction.  Acidification  to  pH  3  leads  to  a  dissociation 
of  the  trypsin-heparin  complex  with  restoration  of  full  activity.  The  pH^pt 
for  trypsin  is  possibly  shifted  from  8.4  to  lower  values  by  heparin  (Glazko 
and  Ferguson,  1940);  it  is  not  known  if  this  means  that  enzyme  combined 
with  heparin  can  act  on  casein  —  it  is  difficult  enough  to  understand  the 
pHopt  of  proteolytic  enzymes  in  the  absence  of  inhibitors.  The  distance 
between  sulfate  groups  in  heparin  is  10.2  A,  which  is  approximately  equiv- 
alent to  3  peptide  residues  in  proteins,  so  it  was  suggested  by  Kornguth 
and  Stahmann  (1960)  that  heparin  may  bridge  the  active  site  by  combining 
with  cationic  groups  on  either  side.  The  active  site  appears  to  be  covered, 
since  the  hydrolysis  of  benzoylarginamide  by  trypsin  is  inhibited.  Poly-or- 
L-glutamate  and  polycysteate  also  inhibit  trypsin,  but  poly-y-D-glutamate 
does  not,  and  this  is  probably  correlated  with  the  different  distances  between 
C00~  groups  in  these  macroanions.  Poly-D-lysine  inhibits  the  tryptic  hy- 
drolysis of  poly-L-lysine,  equimolar  concentrations  giving  complete  inhi- 


INHIBITION    BY    MACROIONS  457 

bition,  showing  the  importance  of  the  configuration  of  the  polypeptide 
chains  (Tsuyuki  et  al.,  1956).  Polyacrylate  at  concentrations  around  0.25- 
0.5  mg/ml  inhibits  trypsin  at  high  pH  and  accelerates  catalysis  at  low  pH 
(Morawetz  and  Sage,  1955).  Denatured  hemoglobin,  the  substrate,  has  an 
isoelectric  point  around  7.8,  so  that  above  pH  7.8  the  polyacrylate  can 
combine  only  with  the  positively  charged  trypsin  (the  trypsin-polyacrylate 
complex  has  some  activity),  whereas  at  lower  pH's  polyacrylate  also  binds 
to  the  hemoglobin.  Since  the  hemoglobin-polyacrylate  complex  is  more  sus- 
ceptible to  trypsin  and  since  hemoglobin  is  much  in  excess  of  trypsin  (thus 
binding  most  of  the  polyacrylate),  the  rate  is  stimulated  at  lower  pH  values. 
The  inhibitions  by  polyglutamate  and  polyacrylate  are  reduced  by  increas- 
ing the  ionic  strength,  as  expected  (Table  1-15-6).  It  has  been  shown  that 
copolymers  of  glutamate  with  other  amino  acids  (e.g.,  tyrosine,  phenylala- 
nine, or  leucine)  are  more  effective  inhibitors  than  glutamate  polymers, 
but  the  copolymer  of  glutamate  and  alanine  is  less  inhibitory  (Rigbi  and 
Sela,  1964).  Ornithine  polymers  or  copolymers  with  ornithine  are  not  inhi- 
bitory and  will  reactivate  trypsin  inhibited  by  glutamate  polymers.  This 
is  one  instance  in  which  the  inhibition  produced  by  glutamate  polymers  or 
copolymers  is  progressive  and  depends  on  the  incubation  time.  From  the 
different  degrees  of  inhibition  brought  about  by  the  various  copolymers 
and  the  effects  of  the  ionic  strength,  it  was  concluded  that  forces  other 
than  electrostatic  are  involved  in  the  binding.  It  is  interesting  that  trypsin 
seems  to  be  particularly  susceptible  to  macroanions,  inasmuch  as  neither 
chymotrypsin  nor  papain  is  inhibited  by  heparin,  although  both  enzymes 
are  negatively  charged  at  physiological  pH. 

The  hydrolysis  of  acetyltyrosine  ethyl  ester  and  methylhippurate  by  chy- 
motrypsin is  inhibited  35-50%  by  various  proteins  (seralbumin,  ovalbumin, 
and  /?-lactoglobulin )  at  concentrations  equivalent  to  the  enzyme  and  in  the 
absence  of  salt  (Hofstee,  1960).  Addition  of  salts,  particularly  multiply 
charged  ions,  reduces  or  abolishes  the  inhibition.  These  proteins  are  as  po- 
tent inhibitors  as  the  naturally  occurring  chymotrypsin  inhibitors,  but  differ 
in  not  being  so  sensitive  to  salt  concentration.  Carboxymethylcellulose  and 
nucleic  acids  (both  DNA  and  RNA)  also  inhibit  chymotrypsin  (Hofstee, 
1961).  The  complexes  are  dissociated  by  100  mM  KCl.  The  inhibition  is 
not  complete  at  maximal  binding  of  nucleic  acid,  indicating  that  the  active 
center  is  not  directly  involved  in  the  interaction.  Methylhippurate  was  the 
substrate  and  if  protein  substrates  had  been  used  it  is  likely  that  the  active 
center  would  not  have  been  accessible. 

Pepsin 

The  hydrolysis  of  hemoglobin  by  pepsin  is  rapidly  inhibited  by  poly-L- 
lysine  and  this  is  readily  reversible  by  adding  heparin,  which  complexes 
with  the  inhibitor  (Katchalski  etal.,  1954).  Cationic  polyornithine  and  poly- 


458 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


p-aminophenylalanine  act  similarly,  but  anionic  polyalanine,  poly  aspartate, 
and  polyglutamate  do  not  inhibit,  indicating  purely  electrostatic  binding. 
The  inhibition  disappears  at  high  poly-L-lysine  concentrations  (Dellert  and 
Stahmann,  1955).  Changes  in  the  optical  transmittance  also  occur  (Fig.  2- 
15);  that  is,  low  concentrations  of  inhibitor  complex  with  the  enzyme  to 
decrease  the  solubility,  but  as  the  inhibitor  concentration  rises  the  com- 


FiG.  2-15.  Inhibition  of  pepsin  by  polylysine  of  mean  molecular  weight  2100 
at  pH  4.7.  Transmittance  determined  at  400  m//.   (Data  from   Dellert  and 

Stahmann,   1955.) 

plexes  become  more  soluble.  However,  the  correlation  between  inhibition 
and  transmittance  is  such  as  to  suggest  that  the  depression  of  enzyme  ac- 
tivity is  by  no  means  directly  related  to  the  aggregate  size  of  the  complex. 
Some  have  believed  that  pepsin  plays  a  role  in  the  genesis  or  maintenance 
of  gastric  ulcers  and  hence  have  looked  for  inhibitors  that  might  be  affective 
clinically.  Strange  to  say,  they  have  invariably  used  macroanions  such  as 
heparin,  chondroitin  sulfate,  polyhydromannuronic  sulfate,  and  various  pol- 
ymers formed  by  condensation  of  aldehydes  with  hydroquinonesulfonate 
(Levey  and  Sheinfeld,  1954;  Marini  and  Levey,  1955;  Heymann  et  al.,  1959). 
The  isoelectric  point  of  pepsin  is  around  2.8  so  the  relative  effectiveness  of 
macroanions  and  macrocations  might  depend  on  the  experimental  or  phys- 
iological pH.  Although  the  hydrolysis  of  casein  is  inhibited  to  some  extent 
by  these  polymers,  it  is  possible  that  this  is  due  in  part  or  wholly  to  the 
formation  of  complexes  with  the  casein.  It  has  been  claimed  that  chon- 
droitin sulfate  and  the  polymeric  sulfonates  reduce  the  number  of  ulcers 
in  Shay  rats,  but  it  is  doubtful  if  this  is  related  to  pepsin  inhibition,  even 
if  it  occurs,  and  there  are  other  explanations  (for  example,  inhibition  of 
lysozyme,  which  has  also  been  implicated  in  ulceration). 


INHIBITION    BY    MACROIONS  459 


Lysozyme 


It  is  not  surprising  that  this  mucolj'lic  enzyme  is  inhibited  by  a  variety 
of  macroanions  since  its  isoelectric  point  is  above  10.5.  Heparin  is  a  fairly 
potent  inhibitor  of  lysozyme  (Kaiser,  1953),  and  it  has  been  stated  that 
this  is  competitive  with  substrate  (which  was  a  dried  preparation  of  M. 
lysodeikticus)  (Kerby  and  Eadie,  1953).  Inhibition  is  also  exerted  by  hyalur- 
onate,  polysaccharide  of  Pneumococcus,  polyglutamate,  DNA,  and  RNA 
(Skarnes  and  Watson,  1955).  Various  synthetic  polymeric  sulfonates  also 
inhibit  to  different  degrees  (Heymann  et  al.,  1959).  The  most  potent  inhi- 
bitor yet  found  for  lysozyme  is  poly  glucose  sulfate  (molecular  weight  around 
20,000),  although  oxidized  polyglucose  (containing  carboxylate  groups) 
is  likewise  very  active;  tetraglucose  sulfate  is  without  activity  (Mora  and 
Young,  1959).  These  inhibitions  are  generally  reversed  by  increasing  salt 
concentration  or  by  the  addition  of  a  macrocation,  such  as  protamine,  to 
bind  the  inhibitor.  Copolymers  of  glutamate,  tyrosine,  phenylalanine,  and 
leucine  are  potent  inhibitors  of  lysozyme,  and  the  inhibitions  can  be  com- 
pletely reversed  by  polylysine  (Sela  and  Steiner,  1963).  The  greater  inhi- 
bitory activity  of  the  copolymers  compared  to  the  homopolymer  of  gluta- 
mate is  attributed  to  nonionic  bonds;  although  this  is  the  most  likely  ex- 
planation, one  must  recognize  that  the  charge  distribution  is  quite  different 
in  the  copolymer  relative  to  the  homopolymer. 

Hyaluronidase 

It  was  noted  by  Pantlitschko  and  Kaiser  (1951)  that  hyaluronidase  is 
not  significantly  inhibited  by  low  molecular  weight  substances  or  by  high 
molecular  weight  substances  unless  they  are  esterified  with  sulfate  or  are 
otherwise  anionic,  and,  furthermore,  that  inhibitory  macroanions  must  be 
filiform  and  not  globular.  Heparin  and  artificially  sulfurated  polysaccharides 
are  inhibitory;  sulfurated  hyaluronate,  which  can  with  some  justification 
be  thought  of  as  a  true  analog  of  hyaluronate,  inhibits  well.  Hyaluronate  is 
a  high  molecular  weight  polymer  of  i\'-acetylhyalobiuronate  units  and  hence 
contains  free  C00~  groups;  however,  sulfuration  essentially  doubles  the  neg- 
ative charge  on  the  molecules,  and  prevents  degradation  by  the  enzyme. 
Nitrated  and  acetylated  hyaluronates  are  also  inhibitory.  A  few  macro- 
anionic  inhibitors  can  be  mentioned  but  require  no  discussion:  chitin  sul- 
fates, polymers  formed  from  formaldehyde  and  various  phenolic  sulfonates 
(e.g.,  hydroquinone,  catechol,  and  resorcinol),  polymers  formed  from  form- 
aldehyde and  various  hydroxybenzoates,  polyesters  of  phosphate  with  phe- 
nols and  aniline,  polystyrenesulfonate,  sulfated  pectate,  polymethacrylate, 
amylopectin  sulfate,  and  heparin  (Rogers  and  Spensley,  1954;  Bernfeld  et 
al,  1961). 

Alburn  and  Whitley  (1954)  suggested  that  the  inhibition  of  hyaluronidase 


460  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

by  heparin  is  competitive  and  this  was  studied  in  detail  by  Houck  (1957  a), 
who  found  competitive  behavior  from  l/v-l/(S)  plots  for  both  heparin  and 
chondroitin  sulfate  B.  The  inhibitor  constants  and  their  variation  with 


Ki  {mM) 


Temperature  K^  (mlf) 


Heparin  Chondroitin  sulfate  B 


22° 

0.70 

6.17 

27° 

0.74 

6.45 

32" 

0.78 

6.68 

37° 

0.81 

7.00 

6.58 
6.85 
7.10 
7.41 


temperature  are  tabulated,  and  from  these  values  it  was  possible  to  cal- 
culate the  important  thermodynamic  quantities  shown  in  the  following 
tabulation: 


-  AF<^  -  AH°  -  zl<S° 

(kcal/mole)      (kcal/mole)         (e.  u.) 


Heparin  3.07  5.0  6.8 

Chondroitin  sulfate  B  3.0  4.9  6.2 


The  rather  weak  binding  might  indicate  that  only  a  fraction  of  the  anionic 
groups  on  heparin  or  chondroitin  sulfate  B  interact  at  close  range  with 
enzyme  cation  groups.  It  is  difficult  to  predict  the  entropy  changes  in  the 
interactions  of  such  complex  molecules  because  several  factors  may  be  in- 
volved, e.g.,  the  restriction  of  polymer  configuration,  possible  changes  in 
enzyme  structure,  and  release  of  water  of  hj^dration.  The  rather  small 
changes  in  AS  observed  are  probably  the  result  of  the  balancing  of  larger 
changes  in  different  directions.  The  importance  of  ion-ion  interactions  is 
shown  by  the  marked  reduction  in  the  inhibitions  when  the  ionic  strength 
rises  above  0.3. 

The  differences  in  inhibitory  activity  between  low  molecular  weight  sub- 
stances and  polymers  made  from  them  are  well  illustrated  by  Hahn  and 
Fekete  (1953).  Various  phenolic  compounds  inhibit  testicular  hyaluron- 
idase  weakly,  but  upon  polymerizing  these  with  formaldehyde  it  is  possible 
to  obtain  potent  inhibitors.  Their  results  are  expressed  in  terms  of  the  po- 
tency relative  to  resorcinol.  y-Resorcylate  has  an  activity  of  1.5  while  its 
polymer  has  an  activity  of  980.  The  most  active  inhibitor  is  the  polymer 

COO-    COO-    COO-    coo- 

— C— R— C— R— C— R— C— R— C— 


INHIBITION    BY    MACROIONS  461 

of  gentisate  with  values  around  2000-2500  (Hahn  and  Frank,  1953).  Poly- 
esters of  phosphate  with  phloretin  (or  other  polyphenols)  are  very  inhibi- 
tory to  hyaluronidase,  0.005  mg/ml  completely  abolishing  enzyme  activity 
(Diczfalusy  et  al.,  1953).  Phloroglucinol  phosphate  polymer  may  be  even 
more  potent,  0.00013  mg/ml  inhibiting  80%  (Ferno  et  al,  1953).  Such  poly- 
mers may  be  represented  as: 

o-  o-  o- 

— 0— P— 0— R— 0— P— O— R— 0— P— O— 
OH  OH  OH 

The  extent  of  the  inhibition  may  depend  primarily  on  the  distance  between 
anionic  groups  and  the  ability  of  the  polymer  to  assume  the  appropriate 
configurations  on  the  enzyme  surface. 

Ribonuclease 

Pancreatic  ribonuclease  is  strongly  and  competitively  inhibited  by  hep- 
arin (Zollner  and  Fellig,  1952,  1953)  but  the  results  by  different  investigators 
vary  quite  widely,  due  perhaps  to  different  experimental  conditions  (espe- 
cially pH  and  ionic  strength),  different  preparations  of  heparin,  and  differ- 
ent techniques  for  measuring  the  enzyme  activity.  The  competitive  nature 
of  the  inhibition  (at  least  the  reduction  in  inhibition  upon  increasing  RNA 
concentration)  has  been  confirmed  by  Roth  (1953)  and  Houck  (1957  b). 
Increase  in  ionic  strength  reduces  the  inhibition  as  expected  (Houck,  1957  b; 
Heymann  et  al,  1958;  Fellig  and  Wiley,  1959;  Lorenz  et  al,  1960),  although 
Houck  found  some  deviation  from  this  at  very  high  NaCl  concentrations. 
The  results  of  Lorenz  et  al  (1960)  are  quite  typical  (see  accompanying 
tabulation): 


NaCl  (m 

31) 

0/ 

/o 

Inhibition 

0 

97 

30 

50 

50 

33 

100 

0 

Lowering  the  pH  from  around  7.5-8.0  to  5.0  progressively  augments  the 
inhibition  (Zollner  and  Fellig,  1953;  Roth,  1953  b),  which  is  perhaps  a  re- 
flection of  the  increasing  positive  charge  on  the  ribonuclease  (isoelectric 
point  is  9.5).  The  ribonucleases  of  rat  kidney  and  liver  (Roth,  1953  b)  and 
rat  and  guinea  pig  serum  (Rabinovitch  and  Dohi,  1957)  are  also  inhibited 


462  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

by  heparin.  It  has  been  known  for  many  years  that  heparin  depresses  cell 
division  and  it  has  been  said  to  prevent  gelation  of  the  mitotic  apparatus. 
Paff  et  al.  (1952)  found  that  heparin  inhibits  mitosis  in  cultured  chick  heart 
fibroblasts  and  after  24  hr  there  is  a  marked  accumulation  of  granular  ri- 
bonucleoprotein  in  the  cells.  It  was  postulated  that  heparin  might  interfere 
with  the  metabolism  of  nucleoproteins  and  thereby  block  mitosis. 

The  inhibition  of  ribonuclease  by  DNA  was  first  clearly  shown  by  Houck 
(1957  b)  and  this  could  be  thought  of  more  reasonably  as  a  true  analog 
inhibition.  Likewise,  deoxyribonuclease  can  be  inhibited  by  RNA,  the  K^ 
being  around  0.00001  vaM  for  the  endonuclease  of  E.  coli  (Lehman  et  al., 
1962  a,  b).  Although  inhibition  occurs  with  RNA  from  various  sources,  the 
most  potent  inhibitor  is  the  amino  acid  acceptor  RNA  from  E.  coli,  the 
inhibition  being  competitive.  The  other  DNA-cleaving  enzymes  tested  are 
not  inhibited.  The  potency  and  specificity  of  this  inhibition  cannot  but 
stimulate  thoughts  on  the  possible  regulatory  relationships  intracellularly. 

Other  macroanions  have  variable  effects  on  ribonuclease.  ZoUner  and 
Fellig  (1952)  reported  no  inhibition  by  chondroitin  sulfate,  hyaluronate, 
and  alginate,  but  Houck  (1957  b)  found  chondroitin  sulfate  A  and  hyalur- 
onate to  inhibit  equivalently  with  heparin.  Synthetic  polyglucose  sulfate  is 
a  competitive  inhibitor,  the  effect  decreasing  with  increase  in  the  pH,  the 
net  charge  on  the  ribonuclease  being  reduced  (Mora,  1962).  Vandendriessche 
(1956)  studied  the  inhibitions  by  sulfonated  pectin,  poly-j9,'/)-dioxydibenzyl 
phosphate,  and  poly-L-aspartate,  and  found  them  to  be  much  weaker  than 
heparin,  while  Fellig  and  Wiley  (1959)  claimed  that  the  sulfation  of  a  va- 
riety of  polysaccharides  (e.g.,  cellulose,  amylose,  amylopectin,  dextran,  pec- 
tate,  and  nitrochitin)  produces  inhibitors  often  more  potent  than  heparin 
(although  in  this  work  the  inhibition  by  heparin  was  unaccountably  weak), 
Ribonuclease  is  also  inhibited  by  copolymers  of  glutamate  and  tyrosine  (or 
phenylalanine),  which  are  more  affective  than  poly  aspartate  or  poly  gluta- 
mate (Sela,  1962).  Interactions  between  the  benzene  rings  of  the  aromatic 
amino  acids  and  certain  groups  on  the  enzyme  were  believed  to  occur  in 
addition  to  the  electrostatic  forces.  Possibly  the  most  potent  inhibitors 
were  discovered  by  Heymann  et  al.  (1958)  in  a  survey  of  66  macroanions 
of  synthetic  origin,  some  inhibiting  around  50%  at  0.001  mg/ml.  It  was 
noted  that  the  inhibitory  activity  is  markedly  reduced  in  the  presence  of 
proteins,  a  point  worth  considering  in  the  use  of  such  substances  in  cellular 
preparations.  A  number  of  these  polymers  exhibit  antiviral  activity  against 
influenzal  and  vaccinial  infections  in  eggs.  Sulfate  groups  seem  to  be  par- 
ticularly able  to  confer  inhibitory  activity  on  polymers  and  in  the  sulfated 
polysaccharides  the  carboxylate  groups  may  be  relatively  unimportant, 
since  Dickman  (1958)  showed  that  sulfation  of  pectate,  pectate  methyl 
ester,  and  pectic  amide  gives  inhibitors  roughly  equiactive.  It  should  finally 
be  noted  that  ribonucleases  of  different  origins  may  not  be  equally  suscep- 


INHIBITION    BY    MACROIONS  463 

tible  to  macroanions.  A  striking  difference  was  demonstrated  by  Nishitnura 
(1960),  bovine  ribonuclease  being  readily  inhibited  by  polyvinyl  sulfate  to 
which  Bacillus  subtilis  ribonucleases  are  completely  resistant. 

Lipoprotein    Lipase 

Macroions  inhibit  lipoprotein  lipase  but  affect  other  lipases  little  or  not 
at  all.  The  enzyme  from  chicken  fat  is  inhibited  by  macrocations  in  a  purely 
noncompetitive  fashion  and  the  inhibition  is  rapidly  reversible  either  by 
dialysis  or  the  addition  of  a  macroanion  (Korn,  1962).  Poly-L-lysines  of 
increasing  chain  length  are  progressively  more  effective,  which  is  one  of  the 
few  observations  relating  polymer  size  to  inhibition.  The  degree  of  inhibition 
by  various  copolymers  of  tyrosine  and  lysine  depends  on  the  content  of 
lysine,  indicating  purely  electrostatic  binding.  The  enzyme  is  also  inhibited 
by  macroanions,  which  may  act  noncompetitively  or  competitively  (e.g., 
heparin  and  polyglucose  sulfate).  The  lipoprotein  from  mouse  heart  is  inhi- 
bited by  many  polysaccharide  sulfates  as  long  as  there  is  at  least  0.6  sulfate 
group  per  repeat  unit  (Bernfeld  and  Kelley,  1963).  The  potency  of  the  inhi- 
bition is  independent  of  the  configuration  of  the  polysaccharide,  whereas 
in  the  case  of  the  chicken  enzyme  the  highly  branched  polymers  are  less 
effective  than  the  more  linear  ones. 

Polynucleotide  Phosphorylase 

This  enzyme,  usually  obtained  from  Azotobacter  vinelandii  or  Micrococcus 

lysodeikticus ,  catalyzes  the  synthesis  of  polynucleotides,  such  as  polyade- 

nylate: 

ADP  +  (A.AIP)„  ^  P,  +  (AMP)„^i 

The  synthesis  of  a  particular  polynucleotide  may  be  inhibited  by  another 
polynucleotide;  thus  polyuridylate  inhibits  the  synthesis  of  poly  adenylate, 
and  polyuridylate,  poly  adenylate,  and  RNA  inhibit  the  synthesis  of  poly- 
cytidylate  (Mii  and  Ochoa,  1957).  The  formation  of  polyadenylate  has  been 
shown  to  be  inhibited  by  variously  degraded  yeast  RNA  and  polyadenylate 
(Hendley  and  Beers,  1959,  1961;  Beers,  1961).  Increase  in  substrate  con- 
centration reduces  the  inhibition  but  not  in  a  strictly  competitive  fashion. 
Competition  with  primers  or  activators  was  considered  to  be  the  most  likely 
mechanism.  If  the  polynucleotides  are  too  extensively  depolymerized,  the 
inhibitory  activity  falls,  indicating  a  certain  minimal  chain  length  for  op- 
timal inhibition.  It  was  also  established  that  phosphate  groups  on  or  adja- 
cent to  the  3'-position  of  the  terminal  ribose  units  are  necessary.  It  is  im- 
portant to  remember  in  such  systems  that  interactions  between  different 
polynucleotide  chains,  perhaps  through  hydrogen  bonding,  can  occur  (Warn- 
er, 1957);  the  possible  role  of  such  interactions  in  the  inhibitions  observed  is 
not  yet  completely  understood. 


464  2.   ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Amylases 

Both  a-  and  /^-amylases  from  barley  are  inhibited  by  heparin  at  concen- 
trations around  0.1  mg/ml  (Myrback  and  Persson,  1953  a,  b).  These  inhi- 
bitions show  a  striking  pH  dependence.  At  pH  5  or  above  there  is  no  inhi- 
bition whereas  below  pH  5  the  inhibition  increases  rapidly  and  is  very  mark- 
ed at  pH  4.5.  The  isoelectric  points  of  amylases  usually  lie  between  pH  5 
and  6,  so  it  is  likely  that  the  combination  of  heparin  with  the  enzymes 
below  pH  5  is  due  to  the  positive  charge  arising  in  this  range.  The  electro- 
static nature  of  the  binding  is  indicated  by  the  protection  afforded  by  high 
concentrations  of  NaCl  during  the  incubation  of  the  enzyme  and  the  heparin. 
Once  the  inhibition  is  established,  raising  the  NaCl  concentration  will  not 
reactivate.  The  presence  of  substrate  also  protects  the  enzyme,  pointing  to 
a  basically  competitive  mechanism.  Human  salivary  amylase  is  also  inhib- 
ited by  heparin  but  the  critical  pH  is  at  least  one  unit  higher,  possibly 
because  of  a  higher  isoelectric  point  than  the  barley  enzymes  (Astrup  and 
Thorsell,  1954). 

Enzymes  Acting  on  Substrates  Which  Are  Not  Macroions 

Macroions  interact  with  proteins  generally  when  the  conditions  are  favor- 
able (e.g.,  when  the  total  charges  on  protein  and  macroion  are  opposite, 
although  this  is  not  a  necessary  condition,  and  when  the  ionic  strength  is 
low),  so  it  is  not  surprising  that  many  enzymes  whose  substrates  are  small 
molecules  are  inhibited.  The  combination  is  probably  seldom  at  the  active 
center,  but  more  often  a  bridging  or  covering  of  the  active  center  by  larger 
molecules  bound  at  many  points  and  in  no  specific  orientation.  This  type 
of  inhibition  is,  of  course,  independent  of  structural  relations  with  the  sub- 
strate, but  must  always  be  considered  in  the  use  of  macroionic  inhibitors, 
A  few  examples  only  will  be  mentioned. 

Prostatic  acid  phosphatase  is  potently  inhibited  by  polyphloretin-phos- 
phate  in  a  noncompetitive  fashion  (Diczfalusy  et  al.,  1953;  Beling  and  Dicz- 
falusy,  1959).  K^  is  given  as  1.55  //g/ml.  Polyestradiol  phosphate  is  even 
more  inhibitory  {K^  =  0.55  //g/ml).  The  inhibitions  increase  as  the  pH  is 
lowered  below  the  pH^pt.  Alginate  of  556  residues  and  molecular  weight 
of  92,000  is  also  strongly  inhibitory  (^,  =  0.0054  mM);  the  mechanism  is 
partially  competition  with  substrate  and  partially  interference  with  stabi- 
lizing or  protective  substances  present  (Jeffree,  1957).  The  variation  of  the 
inhibition  with  chain  length  is  complex:  chains  of  10-100  residues  inhibit 
less,  but  below  10  the  inhibition  rises  (Jeffree,  1956).  A  third  potent  inhi- 
bitor is  polyxenyl  phosphate,  a  polydisperse,  small  molecular  weight  poly- 
mer of  branched  chains  and  random  coils,  66%  inhibition  being  given 
by  0.001  mM  (Hummel  et  al.,  1958).  Polyhydroquinone  is  almost  as  potent, 
polyethylenesulfonate  and  sulfonated  polystyrene  inhibit  moderately,  and 


INHIBITIONS    BY    NUCLEOTIDES  465 

polyacrylate  and  chondroitin  sulfate  are  relatively  inactive.  The  inhibition 
by  polyxenylphosphate  is  noncompetitive,  partially  reversed  by  raising  the 
NaCl  concentration,  and  maximal  at  pH  4.6,  decreasing  on  either  side. 

/?-Fructofuranosidase  of  yeast  is  inhibited  by  heparin  and  chitin  disiilfate 
at  low  pH's  (Astrup  and  Thorsell,  1954).  The  glucuronidases  from  several 
rat  tissues  are  inhibited  by  heparin  and  hyaluronate  (Becker  and  Frieden- 
wald,  1949).  Fumarase  is  inhibited  92%  by  heparin  at  a  concentration  of 
0.2  mg/ml  in  a  pH  range  of  5.5-6.0,  whereas  nucleic  acid  and  chondroitin 
sulfate  inhibit  only  26%  and  11%,  respectively,  at  2  mg/ml  (Fischer  and 
Herrmann,  1937).  These  examples  indicate  that  inhibitions  of  this  sort  are 
widespread.  There  has  been  only  one  investigation  of  the  effects  of  ma- 
croions  on  a  complex  metabolic  sequence,  the  study  of  Dische  and  Ash  well 
(1955)  on  the  actions  of  ribonucleate  and  some  smaller  anions,  such  as  sul- 
fate, on  anaerobic  glycolysis  in  pigeon  hemolysates.  RNA  inhibits  lactate 
formation  48%  at  1  mg/ml  and  the  formation  of  phosphoglj^ceraldehyde 
30%  at  3  mg/ml.  There  would  thus  appear  to  be  at  least  two  sites  of 
action,  the  major  effect  being  on  the  transformation  of  3-phosphoglyceral- 
dehyde  to  lactate. 

INHIBITIONS    BY    NUCLEOTIDES 
AND  RELATED  SUBSTANCES 

Enzymes  acting  on  pyrimidines,  purines,  nucleosides,  nucleotides,  or  poly- 
nucleotides are  frequently  inhibited  by  analogs  of  these  substrates.  Some- 
times the  inhibitors  are  normally  occurring  substances  and  it  is  here  that 
some  of  the  most  clear-cut  and  important  examples  of  feedback  control 
and  metabolic  regulation  have  been  demonstrated.  In  other  cases  the  inhi- 
bitors are  synthetically  derived  abnormal  analogs,  which  are  frequently 
quite  depressant  to  rapidly  growing  cells  where  nucleotide  metabolism  is 
active  and  have  for  this  reason  been  studied  with  regard  to  carcinostasis. 
Many  instances  of  inhibition  have  been  reported,  some  of  which  are  sum- 
marized in  Table  2-28,  but  thorough  quantitative  work  and  studies  of  the 
mechanisms  are  rather  uncommon. 

Most  of  the  inhibitions  in  Table  2-28  appear  to  be  competitive  and  prob- 
ably many  of  those  in  which  the  kinetics  were  not  studied  are  competitive. 
Although  the  inhibitory  activity  of  most  of  these  analogs  is  low  or  moderate, 
a  few  analogs,  particularly  the  abnormal  aza  and  fluoro  derivatives,  are 
quite  potent.  It  seems  that  all  parts  of  the  nucleotide  structure  can  con- 
tribute to  the  binding.  Where  only  the  nature  of  the  purine  or  pyrimidine 
component  is  varied,  the  inhibitions  may  be  very  different,  indicating  that 
the  ring-substituted  groups  can  be  important.  The  pentose  structure  is  also 
a  determinant  since  different  activities  are  observed  in  nucleotides  contain- 
ing ribose  or  deoxyribose.  Finally,  the  number  of  phosphate  units  in  the 


466 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


m 

a 

^ 

o 

O 

Iz; 

^oi 

-s! 

o"  :S 

H 
en 
« 

c 
1— 1 

CO 

a 

c    g   — 

H 
H 
< 

§    £    S 

A 


o  -^^ 


S 


W 


Cm 


m 


Cm 


a 
1^ 


^ 

(35 

1— 1 

<A 

s:! 

A 

o 

^ 

TJ 

'p 

> 

.SP 

eS 

Q 

^^ 

bO 

CO 

>— ^ 

TS 

o 

(M 

pq 

Oi 

05 

i-H 

.2 

> 

73 

CD 

"3 

SI 

5 

S 

05 

30      05      ■>#         O      r-H      00      >— I 

O    Tf    T)<      O    CO    Ci    — < 


CO    CO    CO    CO 
(M(NC<I      COCOCOCO      Oi 


>, 

c 

fl^    '2 

© 

0) 

'S 

.5    § 

<» 

3 

3 

4)     7 

o 

3 

2 

c 

X 

eS 
O 

a 
o 

Ph 

a) 

& 
O 

-1-2 

ft 

o 

-4-^        r^ 

c 

r- 

cS 

o3 

tH 

k< 

0)        C 

c3 

-^ 

§  Q 

S 
< 

< 

< 

« 

^ 

^ 

(N    ci 

<N 

!z;  ^ 

GO 

CO 

<>l 

<© 

<i3 

o 

t^ 


s 

^ 

c 

c3 

4) 

c 

LO 

0) 

C 

■g 

— H 

c« 

'2 

>> 

73 

(D 

■43 

T3 

T3 

P5 


C     cS 


i^ 


c<i   00  CO   -H 


ic  >c  »n  »o 
dodo 


0 

2 

>. 

(U 

fi 

■^ 

c 

OQ 

83 

« 

73 

73 

_>, 

<I 

<1 

INHIBITIONS    BY    NUCLEOTIDES 


467 


05 


s 

(D 


.!>  s  «  2 

^  o  >  o 


TJl 


:^  I 


o  o 


O    t^    O    fO 


o  o  o  o  o 


"3 

.s  s 

0^ 

-a  -a 

PM    Plh 

o 

K 

<  •< 

<  <; 

Pm" 

Pm 

PM 

S 

S 

S 

I— 1 

I— 1 

h- 1 

o  p^ 

0 

Ph 

o  f^ 

-S  H 

-S 

H 

-§  H 

^  O 

O 

J  a 

ki 

kl 

ki 

g    'TJ 

a 

-o 

g  -o 

a  c 

ft 

c 

ft    c 

3    ^ 

—<< 

eS 

^  =« 

o    « 


.-     O  P^   Ph 

H      3      3^Qgg«jJO 


Pm 


PM 
H 

S  P^ 

;s  ■§  p^ 

'5c  'tic  W 


O 


O   to 


t3   ^ 


« 

•^ 

, ^ 

-fi 

5i 

m 

w 

<xl 

«3 

>. 

5. 

!^ 


fe)  EtJ 


O 


>^ 


-2 


'3 

§1 
®    r 


S « 


468 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


P3 


^ 


c 


a>    o  ^ 


O  -^ 


,0 


5i     OJ 


m 


OO-^OQ0Q000C0Q0G0>C"S      t- 


a) 

o    c    c 
?    o  !§ 

^  ^^ 

10  o  o 


I 


P-i  Cu 

^  ^  fe  H 

o  Q  H  o 

T3  O  O  X) 


■■5    O 

o 


Q  o 


c 

Ah   !3   & 
O  ^  t3 


13 


a 


^ 


t3 


K) 


P3 


INHIBITIONS    BY    NUCLEOTIDES  469 


(D  o 

§  o 

c3    ^  - 


0^  CO  ^^ 

WW  S 


o  ©  ^  o 


ultOOTftO^^IOOCJQOfMtOlOiMlMOCOO-"* 


-H      Tj< 


TS 

%^ 

o   ">> 

-1^       r" 

eS     a 

c^ 

^  - 

o 

t3 

I*  ^^ 


CD    ffO 


Ci-^XC-lOOOCiC-H-^O-^OTfOTtiO-^O-^O        o^ 


-  '^  -M 

><     O    ^  S 

®'^>i  CL|  CLiP-iCLiiOj  CUi  CU  n_, 


I  g  'a 


c4  eS 

si  o.i 


470 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


tf 


^ 


r^ 


^3 

'si 
c 


<»    o  ^ 


O    -^ 


03 

3 

t3 

^ 

CB 

M 

73 

(H 

c 

iS 

0) 

c3 

^_. 

C 

:g 

3 

c6 

O 

■+-» 

05 

o 

W 

a 

1—1 

><     GO 


t^ 


O    <M    •*    O    <M    -t 


o 


fe) 


c 

O 

3 

cS 

(U 

-ki 

a 

a 

o 

N 

^ 

-ti 

» 

(i 

>> 

y, 
o 

02 

C 

t3 

^ 

P 

S 

^^ 

a 

— 

c3 

'2 

a 

73 

^ 

cS 

"^ 

oS 

a 

a 

ID 

c 

'o 

CI 

t3 

CI 

.2 

<» 

« 

0 

'C 

> 

'C 

> 

P>iH 

piH 

5    s     5   s 


ts^ 


:0    t^    CO    ^    ^ 


o  o 


p— t    fo    CO    CO    CO    CO 

S  S  d  d>  d>  S 


pL,  Ph  Cm  S:! 

O  ^  §  9 
<1  <1  S  P 


1 1 

1      d 

CO     -1^ 

o    ® 

-     cS 

u     bD 

^    ^ 

O     O 

2  ^ 

O      O 

b  -^ 

1^ 

-C     a; 

5^ 

INHIBITIONS    BY    NUCLEOTIDES 


471 


H 


X!  :2 


a 

'c 

a 

u 

03 

» 

bC 

M 

a 

t3 

03 

,^ 

o 

CO 

M 

aj 

t-( 

eS 

c3 

t-i 

a) 

U) 

H 

O 

M 

cS 

W 

cS 

S 

P5 


IC    O    O    ^     "O 


X    X 


IC    -H    (M    LO    (M    iM 


CD 

CD     05 


fN    (M    (M         (M    IM 


Pm 

Q 

<: 

+" 

^ 

X 

^-^ 

^ 

X 

aT 

Ph" 

o3 

H 

S 

< 

c6 

T3 

O 

C 

P^  Ph 

<3  M 


•2  -c 


c3      tlS 


CL|    P-! 

<5  O 


.s  « 

c   S 

a;    cs 

P^ 

o 

5  § 

0) 

73  -r, 

+3        H 

H 

"3  "s 

0)     c 

'c 

t3 

T3 

<0    .S 

^  ft 

O 

O 

CO 

oq    C<J 

c3 

<1 

fe;  fe; 

Ph 


P.  a 


o 


o 


c 

(1 


W 


fe) 


as 

(C 

i 

aj 

c 

ii 

C 

03 

c3 

a 

<D 

o 

CJ 

eS 

■^ 

c 

3 

s 

3 

T3 
a; 

5 

02 

O 

(-C 

aj    o 

O     >> 

C    J3 


a;   73 


'So   2 

c    ft 


472 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


« 


^ 


C     rt     fl 


w 


o 


o  o  o  o  o  o  o 


c 

c 

c 

« 

"^ 

fO 

to 

i 

o 

^ 

a 

O 

2      03       ^ 


C   •=   'o    ;3 


^  ^  t^  H  hiq  O  U) 


'-' 

0) 

'ti 

o 

^ 

Q 

t^ 

< 

PH 


>H 


Ah 

Ph 

»o 

lO 

OJ 

i 

0) 

c 

C 

g 

'2 

'2 

'2 

« 

S 

'C 

3 

3 

'S 

.^ 

s 

cS 

cS 

T3 

SI 

N 

N 

>-^ 

"u 

*? 

<1 

< 

a 

p 

«o 

CO 

CO 

Ah 

PM 

lO 

in 

V 

V 

c 

£3 

'43 

o 

'•5 

O 

d 

2 

O 

"^ 

o 

>H 


a 
>, 

tsi 
C 


0) 

o    0) 


INHIBITIONS    BY    NUCLEOTIDES  473 


Ah 


J2  '^  o  g 

pq^  ffi  pq  P5"  <1"  tf 


rfl 

c3 

C 

r1 

ISl 

g 

W 

-C 

fO 

TO 

c 

<1) 

fc< 

o< 

<D 

ft 

T! 

01 

C 

W 

< 

o   o   ^ 

O    CO    l> 

odd 


■^CSOCOfMOOOOOO-^OO 


■^    t^    00    t^    QO 

^H    S<l    o    'M    CD 

o  o  o  o  o 

OOOOO05TtiC0-<*l00C<5  CO 

OOodo^0^r-^(MO-^0^^— <-H 


c    c    ^ 

i§     ^2     ^              ^  fH 

■S    -S    -o     C   2:3  ft                H           §    ^                                                               n, 

p.             PhP,222:2^  H               ft         c  <        d,        p^        p,        PhO^S 

P              DPodoJo  ft             H         <ft        <         <         <         <1Ht3 
6 

^                            <^  i  S  ft  ft 


ft 


V      ft      b  CO  >. 


^-  I  gill-  "^  I?? 


is 

3 


^ 

s 

a; 

ni 

r/l 

3 

^ 

n 

§ 

TO  -S  5       °0 

P5  t^  P50~  g"  fi 


!r    ja  O   -Q  ft 


ft  "C        s= 


^    '^         ^ 


474 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


tf 


^ 


§  g  a 


M  — ' 


i4 


Ph  ^  CM 

o  p^  o 

03  'Tj  to 

O  ■<  O 


fe 


■^  £ 


c  ^ 

S    Oi     00 
TJ    ^    C5 


o  o 
d  d 


73 

c 

o3 


3 

■ft 

O 

c3 

C 

'3 

c 
o 

M) 

o 

« 

ID 

s 

<D 

s 

CO    Ci 

Oi 

o 

O       03 

ee 

s 

c^ 

X!       C 

S) 

a 

6 

t 

OQ 

o   i2 

-S3 

o 

.-C 

<» 

A     o 

H 

ft 

CO 

ft 

CO 

H 

ft  -Ci 
to     •-' 

O 

s 

O 

o 

O      P 

^ 

ci5 

.22 

t      ^ 

>•   lO     g    CD 
S    ""  j» 


P-i  Ph  Pli  Ph 

Q  §  H  P 

O  Q  Q  O 

"C  t;  "C  "O 


c«  Ph  - 

1^  Ph  5::' 

-S  £  Q 

O  < 


OS 

U5 

_^_^ 

O) 

^^ 

l-H 

CO 

^ 

T3      r-* 

f** 

c  ^ 

e 

=«          tH 

u 

.2   -2 

fl    2 

eS 

S  -fs 

T3 

'0     n 

O 

fi     o 

-^ 

^  w 

>c» 

t> 

W5 

I-H 

d 

o  .cr  00  00  o 

C    ^    M    00 


2    ft   ?^ 


lO 


^•9^ 


p    to 


P-I  p-l 
P  P 


S  G  S;  rj  P^  JJ  _iH 
-S  ii  'O  ^  H  •<  ^ 
<jj         ■<         •<   CO   CO 


I 

c 


o   C 


B^ 


Si  s 
'^^  ■" 


INHIBITIONS    BY    NUCLEOTIDES 


475 


O    O    O    O    <M 


M      i-H      .^      fH 

o  CO  ;3  CD 

O     05    "g     § 


^ 

-o 

—- 

OS 

cS 

<D 

c 

P 

0) 

o 

C 

t^ 

to 

3 

N 

d 

05     lO 


O 


« 


M 


rt    ^    —    -M    -N    <M 


o  H 
c 


<:  <  t^  ^r;  K 


<1  -< 


^ 


O  CO 


r  Ph 


Ph   S 


<;  <  s 


o    . 

Ti    Ph 

Ph 


-a 

0) 

G 

-ki 

ci 

c3 

«7l 

ft 

o 

Ph 
H 

03 

o 

50 

Ph 

<^ 

>> 
Ph 

S 

PM 

0  <J 

o 

o  C 

1  '§ 


S   Ph 


P4 


^9 


Ph 


I^ 


Pm 


c6 

ft 

O 

-a 
O 

;-! 

?>> 


e6 

< 

_C 

© 

o 

^ 

4^ 

•72      (D 

O 

CD 

fi 

>1 

■g 

>>  .s 

.2 

3 

'TS 

2 

r^   -^ 

a3 

H 

H 

476 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


P5 


^ 


■2^ 


2  lo 

Cd      (—1 


r— I         © 

s  s 


00C<lt^0iCt^'<*C50       O  00^ 


r-      (M      O 

o  o  o 
o  o  o 


O      -H 

O      O      -H 

o  o  o 


■*ooo   ^ooooo 


PM 


p^ 


3 

o 

T3 

o 

p 

-0 

P^ 

Ph 

f^ 

^ 

Pm 

Ph 

LO 

lO 

>o 

lO 

lO 

o 

fe] 


=3      S 


Pi:; 


ce 

CD 

V^ 

ii 

^ 

O 

'g 

■+3 

f-< 

& 

^ 

>. 

H 

M 

P5 


c 

c3 

(H 

t-M 

Pm 

o 
o 

P 

't: 

P 

P^ 

02 


^      03 


PM 


.s  "^ 


X 


^ 


Ph    fli 


o 


P^     o     o     C       "S    -^    "g 
B    'bX)  "bi)  ^      t^     3     M 


><^ 


.3     ft 


o 

c 

/IS 

_o 

cc 

'-3 

03 

r5 

0) 

g 

3 

■-3 

_fl 

a) 

'-3 

g 

CO 

o 

-73 

tc 

C 

a> 

cS 

1^  ^D 


CO 

r^ 

-Jf 

M 

^ 

a 

rH 

-t^ 

o 

OD 

o   ^ 


o    c 


INHIBITIONS    BY    NUCLEOTIDES  477 

nucleotide  is  a  factor  in  the  binding.  However,  the  binding  energy  does 
not  always  increase  with  increase  in  phosphate  residues  or  total  negative 
charge;  the  potencies  of  the  inhibition  of  pig  kidney  phosphodiesterase  are 
AMP  >  ADP  >  ATP.  Little  is  known  about  the  topography  of  enzyme 
sites  for  nucleotides  but  it  is  clear  that  multiple  binding  sites  must  be  in- 
volved. The  relative  binding  energies  for  competitive  inhibitors  of  yeast 
pyridoxal  kinase  (see  tabulation)  (Hurwitz,  1953)  show  that  here  the  purine 


Inhibitor 

Relative 

—  AF  oi  binding 
(kcal/mole) 

Adenine 

4.60 

Adenosine 

4.55 

AMP 

4.51 

ADP 

5.17 

ITP 

3.77 

Pyrophosphate 

3.90 

component  is  of  primary  importance,  addition  of  ribose  or  phosphate  res- 
idues having  little  effect.  Yet  ITP  inhibits  while  inosine  does  not,  and 
pyrophosphate  is  bound  fairly  tightly  to  the  enzyme.  The  phosphate  resi- 
dues in  ADP  must  not  be  oriented  for  interaction  as  optimally  as  free 
pyrophosphate. 

Azaguanine  and  Azauracil 

The  8-azapurines  are  usually  inhibitory  to  growth  and  this  has  been  gen- 
erally attributed  to  the  incorporation  of  these  analogs  to  form  abnormal 
polynucleotides  which  are  nonfunctional  or  inhibitory.  However,  it  has  more 
recently  been  found  that  these  analogs  or  their  immediate  metabolic  prod- 
ucts are  potent  inhibitors  of  certain  enzymes  involved  in  purine  metab- 
olism, and  it  is  possible  that  such  actions  contribute  to  the  growth  de- 
pression. 8- Azaguanine  has  been  studied  the  most  thoroughly  and  has  been 
shown  to  inhibit  the  growth  of  many  bacteria,  fungi,  algae,  viruses,  tissue 
culture  cells,  chick  embryos,  epithelium,  and  tumors.  It  is  usually  antag- 
onized by  guanine  or  guanidylate,  and  occasionally  by  other  purines,  nu- 
cleosides, and  nucleotides.  The  intention  is  not  to  discuss  these  azapurines 
in  detail  since  it  is  a  very  large  subject  but  to  mention  only  a  few  observa- 
tions bearing  on  enzyme  inhibition. 

Adenosine  deaminase  is  inhibited  reversibly  by  8-azaguanine  {K,  =  0.28 
mM)  and  the  inhibition  was  stated  to  be  noncompetitive,  although  the 
l/f-l/(S)  plot  appears  to  indicate  uncompetitive  or  coupling  inhibition 
(Feigelson  and  Davidson,  1956  b).  Xanthine  oxidase  is  also  strongly  inhi- 
bited (Table  2-2).  Unfortunately,  very  few  enzymes  operative  in  purine 


478  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

metabolism  have  been  examined  with  respect  to  8-azaguanine  inhibition, 
or  to  inhibition  by  nucleosides  and  nucleotides  of  8-azaguanine,  but  this 
type  of  mechanism  for  the  growth  inhibition  must  be  borne  in  mind.  8- 
AzaGTP  is  formed  from  8-azaguanine  and  can  serve  as  a  substrate  for 
adenylosuccinate  synthetase;  it  also  inhibits  competitively  with  respect  to 
GTP  (Cohen  and  Parks,  1963).  Here  one  sees  the  interesting  situation  in 
which  8-azaGTP  stimulates  the  rate  when  GTP  is  low  and  inhibits  the 
rate  when  GTP  is  high.  It  was  pointed  out  that  in  such  cases  one  might 
obtain  selective  inhibition  of  an  enzyme  in  tissues  having  a  relatively  high 
substrate  concentration.  This  behavior  is  characterized  in  the  double  re- 
ciprocal plot  by  the  curve  for  the  inhibited  reaction  crossing  the  curve  (or 
straight  line)  for  the  uninhibited  reaction  and  being  nonlinear.  8-Azaguanine 
suppresses  the  induction  of  liver  glucose-6-phosphatase,  fructose-l,6-diphos- 
phatase,  and  tryptophan  pyrrolase  in  the  rat  (Kvam  and  Parks,  1960)  and 
the  formation  of  catalase  in  yeast  (Bhuvaneswaran  et  al.,  1961).  These  ef- 
fects on  protein  synthesis  may  relate  to  an  interference  with  nucleic  acid 
metabolism,  but  whether  it  is  a  general  depression  or  a  more  specific  block 
is  not  known.  Studies  on  8-azaguanine  in  animals  are  complicated  by  the 
rapid  deamination  to  the  relatively  noninhibitory  8-azaxanthine  so  that 
only  a  fraction  of  the  quantity  fed  is  available  for  either  inhibition  or  in- 
corporation (Mandel,  1955).  Thus  the  usefulness  of  8-azaguanine  in  tumor- 
istasis  is  limited. 

6-Azauracil  is  also  generally  growth-inhibiting  and  tumoristatic.  It  is 
possible  that  6-azauridine-5'-P  (6-azaUMP)  is  the  true  inhibitor,  since  it 
has  been  shown  that  orotidylate  decarboxylase  is  strongly  inhibited  by 
6-azaUMP,  leading  to  the  accumulation  of  orotidylate  (Handschumacher 
and  Pasternak,  1958;  Pasternak  and  Handschumacher,  1959).  6-Azauridine 
is  metabolized  to  6-azaUMP  and  inhibition  of  the  decarboxylase  after  ad- 
ministration of  6-azauracil  was  demonstrated,  so  it  is  likely  that  the  block 
in  pyrimidine  metabolism  is  at  this  point  and  that  this  is  an  important 
mechanism  in  the  tumoristatic  action.  The  inhibition  of  the  decarboxylase 
is  characterized  by  a  K,  of  0.00075  mM  (Handschumacher,  1960).  It  is 
interesting  to  speculate  that  a  similar  mechanism  might  be  involved  in  the 
action  of  8-azaguanine. 

Fluoropyrimidines  and   Feedback   Inhibitions   in   Pyrimidine   Pathways 

The  fluoropyrimidines  are  among  the  most  potent  inhibitors  of  nucleic 
acid  biosynthesis  yet  discovered  but  the  sites  of  action  have  not  been  com- 
pletely elucidated.  The  5-halogen  analogs  of  orotate  inhibit  the  conversion 
of  orotate  to  the  uridine  phosphates,  the  most  active  being  the  fluoro  com- 
pound (Stone  and  Potter,  1957).  It  was  suggested  that  some  of  the  action 
could  be  due  to  nucleotide  analogs  formed  from  these,  and  it  was  shown 
that  5-fluoroorotate  is  converted  to  5-FUMP  in  yeast  (Dahl  et  al.,  1959). 


INHIBITIONS    BY    NUCLEOTIDES 


479 


However,  5-fluoroorotate  blocks  earlier  in  the  sequence,  an  inhibition  of 
dihydroorotase,  which  forms  dihydroorotate  by  cyclization  of  carbamylas- 
partate,  having  been  observed  (Smith  and  Sullivan,  1960).  Orotate  also 
inhibits  but  more  weakly,  this  being  an  example  of  negative  feedback. 
Orotidylate  decarboxylase  which  forms  UMP  is  inhibited  by  UMP,  although 
not  by  uridine,  UDP,  or  UTP,  and  it  is  possible  that  5-FUMP  would  also 
inhibit  at  this  locus  (Blair  and  Potter,  1961).  The  conversion  of  dUMP  to 
TMP  by  thymidylate  synthetase,  which  is  phage-induced  in  E.  coli,  is  very 
potently  inhibited  by  5-F-dUMP,  K^  being  around  0.00005  mM,  and  follow- 
ing a  competitive  phase  there  is  irreversible  reaction  with  the  enzyme  (Ma- 
thews and  Cohen,  1963).  This  illustrates  the  principle  that  analogs  often 
simulate  feedback  inhibition  if  they  are  structurally  similar  to  the  com- 
pound normally  exerting  the  inhibition.  Some  of  the  inhibitions  observed 
in  pyrimidine  nucleotide  metabolism  are  shown  in  the  following  scheme 
modified  from  Smith  and  SuUivan  (1960). 

A  +  CP 


(FCMP,  CTP) 


CA 


(O,  FO) 


DHO 


DNA- 


(FdUMP) 
TMP  -« — K dUMP 


)(  (FO) 


OMP 


X   (UMP) 


UMP 


-UDP 


(FUMP, 
FUDP) 

X  * 


UTP 


CTP 


]<.    (FUMP,  FUDP) 
RNA 

(A  =  aspartate,  CP  =  carbamyl-P,  CA  =  carbamylaspartate,  DHO  =  di- 
hydroorotate, 0  =  orotate,  OMP  =  orotidine-5'-P,  and  fluorinated  analogs 
are  indicated  by  an  inital  F). 

The  effects  of  5-fluorouTacil  on  protein  synthesis  in  E.  coli  and  B.  mega- 
teriimi  are  very  interesting  because  total  protein  synthesis  is  not  altered 
significantly  but  the  proteins  formed  have  abnormal  amino  acid  compo- 
sitions (Gros  and  Naono,  1961).  For  example,  the  proteins  contain  less 
proline  and  tyrosine  but  more  arginine.  The  alkaline  phosphatase  has  nor- 
mal catalytic  activity  but  is  less  thermostable,  whereas  a-galactosidase 


480  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

seems  to  be  synthesized,  as  shown  serologically,  but  is  catalytically  inactive. 
An  RNA  fraction  into  which  5-fluorouracil  is  rapidly  incorporated  was  de- 
tected and  it  is  possible  that  this  is  responsible  for  the  changes  in  protein 
synthesis.  The  halopyrimidines  are  very  potent  growth  inhibitors  but,  al- 
though much  is  known  of  their  fate  and  actions  (Brockman  and  Anderson, 
1963),  the  over-all  effects  produced  are  usually  so  complex  that  the  primary 
sites  of  block  have  not  yet  been  determined.  These  analogs  can  be  metabol- 
ized into  such  a  variety  of  abnormal  substances  which  can  inhibit  at  many 
different  sites  that  it  is  likely  no  single  mechanism  for  the  growth  inhibition 
will  be  found.  More  investigations  of  the  changes  in  the  steady-state  con- 
centrations of  the  intermediates  in  these  pathways,  as  they  are  affected  by 
the  analogs,  would  be  valuable  in  determining  the  important  loci  attacked. 
A  few  examples  of  enzyme  inhibitions  which  may  be  involved  in  feed- 
back regulation  of  pyrimidine  and  purine  metabolism  will  be  cited  because 
of  the  importance  of  this  type  of  inhibition  in  the  control  of  nucleic  acid 
and  protein  synthesis.*  Gots  and  Gollub  (1959)  described  the  suppression  of 
the  formation  of  purine  precursors  in  E.  coli  whenever  a  purine  which  sup- 
ports growth  is  added,  and  the  accumulation  of  5-amino-4-imidazolecar- 
boxamide  is  suppressed  in  certain  mutants.  Various  purine  analogs  (e.g.,  6- 
mercaptopurine  and  2,6-diaminopurine)  act  like  the  normal  feedback  inhibi- 
tors, only  more  potently.  These  analogs  can  thus  act  on  at  least  two  sites 
in  the  biosynthetic  sequence,  the  formation  of  purine  precursors  and  the 
eventual  utilization  of  the  purines,  their  therapeutic  usefulness  possibly  be- 
ing related  to  this  type  of  sequential  inhibition.  It  may  be  noted,  however, 
that  Rubin  et  al.  (1964)  have  recently  shown  that  sequential  inhibition  in 
pyrimidine  biosynthesis  is  not  synergistic.  Combinations  of  5-azaorotate, 
which  competitively  inhibits  the  conversion  of  orotate  to  orotidylate,  and 
6-azauridine,  which  after  its  metabolism  to  6-azauridylate  competitively  in- 
hibits the  conversion  of  orotidylate  to  uridylate,  do  not  produce  greater 
inhibitions  in  either  isolated  enzyme  systems  or  leucocytes  than  are  seen 
with  the  individual  analogs.  Almost  every  step  in  the  nucleotide  synthesis 
has  been  shown  to  be  inhibited  by  more  distal  intermediates.  The  PRPP- 
amidotransf erase,  which  catalyzes  the  first  irreversible  and  specific  step  in 
purine  synthesis,  utilizing  glutamine  as  the  amino  donor,  is  inhibited  by 
AMP,  ADP,  ATP,  GMP,  GTP,  other  nucleotides,  and  some  analogs  (Wyn- 
gaarden  and  Ashton,  1959);  adenylosuccinate  synthetase  (Wyngaarden  and 
Greenland,  1963),  aspartate  transcarbamylase  (Bresnick,  1963;  Gerhart  and 
Pardee,  1962),  phosphoribosylformylglycineamidine  synthetase  (Henderson, 

*  Although  an  intermediate  or  product  in  a  metabolic  sequence  is  shown  to  be 
an  inhibitor  of  an  enzyme  catalyzing  a  previous  step,  it  is  perhaps  not  justified  to 
call  it  a  feedback  inhibitor,  which  implies  that  inhibition  occurs  during  the  in  vivo 
operation  of  the  pathway.  For  various  reasons  the  substance  may  not  play  a  role  in 
regulating  metabolism,  even  though  it  is  a  reasonably  potent  inhibitor. 


INHIBITION    BY    NUCLEOTIDES  481 

1962),  and  other  enzymes  are  inhibited  in  similar  manner  but  each  exhibits 
a  unique  pattern  (Table  2-28);  while  TTP  inhibits  several  steps  in  its  for- 
mation, including  CDP  -^  dCDP,  deoxyuridine  -^  dUMP,  and  deoxythy- 
midine  ->  dTMP  (Ives  et  al.,  1963).  All  of  these  inhibitions  and  many  more 
constitute  possible  feedback  situations,  but  in  the  cell  probably  only  a  few 
are  important,  since  the  concentrations  of  some  intermediates  may  never 
rise  sufficiently  to  exert  an  effect,  and  compartmentalization  may  limit  the 
actions  of  these  inhibitors.  We  have  mentioned  that  certain  enzymes  ap- 
pear to  contain  sites  specially  evolved  for  feedback  inhibition  (Gerhart  and 
Pardee,  1962,  1964),  the  best  documented  case  being  aspartate  transcarba- 
mylase,  which  is  inhibited  particularly  well  by  CMP,  CDP,  and  CTP.  This 
enzyme  is  normally  a  tetramer  and  it  may  be  that  these  inhibitors  alter  the 
subunit  interactions  since  the  monomer  is  not  inhibited.  Another  interesting 
example  of  this  phenomenon  is  the  inhibition  of  xanthosine-5'-P  aminase 
by  psicofuranine  (9-D-psicofuranosyl-6-aminopurine),  which  occurs  in  two 
steps,  a  reversible  pyrophosphate-dependent  reaction  with  the  enzyme  and 
an  irreversible  xanthosine-5'-P-dependent  reaction  (Udaka  and  Moyed, 
1963).  The  first  step  can  be  observed  in  a  psicofuranine-resistant  bacterial 
strain  and  here  the  inhibition  is  noncompetitive.  It  would  appear  that  the 
inhibitor  is  bound  to  a  different  site  than  that  at  which  the  substrate  reacts 
and  this  second  site  could  have  regulatory  function. 

Some  interesting  results  have  been  obtained  in  the  analysis  of  the  inhi- 
bitions produced  by  the  metabolites  of  6-mercaptopurine,  a  few  of  which 
will  be  mentioned  briefly.  One  product  into  which  6-mercaptopurine  is  con- 
verted is  6-thio-IMP,  a  potent  competitive  inhibitor  of  IMP  dehydrogenase 
(which  is  involved  in  the  formation  of  GMP  in  certain  cells)  {K^^  =  0.0036 
mM)  (Atkinson  et  al.,  1963).  The  inhibition  proceeds  rather  slowly,  requir- 
ing 10-20  min  for  completion,  and  the  enzyme  is  then  inactivated  (Hamp- 
ton, 1963).  Evidence  was  presented  that  reaction  occurs  with  an  SH  group 
on  the  enzyme,  a  stable  disulfide  bond  being  formed  with  the  6-thio-IMP. 
This  is  one  example  where  an  analog  turns  out  to  be  an  SH  reagent.  On  the 
other  hand,  adenylosuccinate  lyase  is  inhibited  by  6-thio-IMP  only  if  a 
metal  ion  is  present  and  it  was  postulated  that  the  metal  ion  forms  a  bridge 
between  the  SH  groups  (E-S-Me-S-IMP)  (Bridger  and  Cohen,  1963).  These 
inhibitions  create  new  possibilities  by  which  analogs  can  inactivate  enzymes. 
Another  product  derived  from  6-mercaptopurine  is  6-mercaptopurine  ribo- 
side-5'-diphosphate,  which  inhibits  polynucleotide  phosphorylase  quite  po- 
tently (50%  inhibition  by  around  0.03  mM),  rapidly,  and  competitively 
(Carbon,  1962).  The  role  this  enzyme  plays  in  vivo  or  the  significance  of 
such  inhibition  is  not  known. 

These  few  remarks  on  the  effects  of  nucleotides  and  related  substances 
are  made  only  to  suggest  certain  interesting  aspects  of  enzyme  inhibition 
which  broaden  our  concepts  of  how  analogs  may  act;  adequate  coverage  of 
this  subject,  young  as  it  is,  would  require  a  volume  of  this  size  or  more. 


482 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


INHIBITIONS  BY  COENZYME  ANALOGS 

This  field  in  which  interest  was  stimulated  by  the  demonstration  of  the 
mechanism  of  sulfonamide  action  is  a  large  one  because  of  the  great  amount 
of  work  done  on  the  growth  inhibition  of  microorganisms  by  these  analogs, 
so  that  here  the  presentation  will  be  restricted  to  those  aspects  directly  re- 
lated to  enzyme  inhibition  and  specific  metabolic  disturbance.  The  competi- 
tion between  a  coenzyme  analog  and  a  coenzyme  for  combination  with  the 
apoenzyme  for  which  the  coenzyme  is  essential  is  basically  of  the  same 
nature  as  the  examples  of  substrate  analogs  discussed  previously.  However, 
there  are  usually  additional  complexities  due  primarily  to  the  greater  num- 
ber of  sites  for  antagonism.  Figure  2-16  indicates  some  of  the  reactions 


(1) 

PRECURSORS  iSLC     -i^^ 

X 

PHOSPHATASE   [(4)  (3)  ATP  + 

KINASE 

\     /         (5, 

CP  +  E  = 

E- 

■CP 

E-CP 
A  ^  B 

Fig.  2-16.  General  scheme  for  the  formation  and  possible  reactions 
of  coenzymes  in  cells.  C  =  a  unit  of  the  active  coenzyme  (e.g., 
nicotinamide,  adenine,  or  thiamine),  CP  =  the  active  coenzyme, 
E-CP  =  the  active  enzyme-coenzyme  complex  for  the  reaction 
A  ->  B,  and  X  =  any  derived  substance  from  the  unit  C,  which 
may  be  inactive,  or  active  to  some  degree  after  phosphorylation, 
or  capable  of  interfering  with  the  formation  or  action  of  the  coen- 
zyme. Reaction  1  may  be  a  simple  diffusion  into  the  cell  or  be 
mediated  by  facilitated  diffusion  or  active  transport;  reaction  (2) 
occurs  in  cells  which  synthesize  the  coenzyme  from  precursors;  re- 
action (3)  is  usually  a  phosphorylation;  reaction  (4)  is  a  dephos- 
phorylation;  and  reaction  (5)  represents  the  complexing  of  the 
coenzyme  with  the  apoenzyme. 


involved  in  coenzyme  formation,  destruction,  and  function.  Inasmuch  as 
the  catalytically  active  forms  of  most  coenzymes  are  formed  within  cells 
from  precursors,  these  reactions  and  the  membrane  processes  responsible 
for  entrance  of  the  precursors  must  be  considered  as  possible  loci  for  analog 
interference.  Furthermore,  in  many  instances  the  analogs  are  metabolized 
along  the  same  pathways  as  the  coenzymes  to  form  inhibitory  products. 
Certain  coenzymes  are  active  in  phosphorylated  forms  and  the  reaction 
immediately  forming  the  active  coenzyme  is  often  a  phosphorylation  in- 


INHIBITIONS   BY   COENZYME   ANALOGS  483 

volving  ATP  and  a  kinase.  The  analogs  are  occasionally  phosphorylated 
and  exert  their  major  effects  in  this  form.  The  inability  of  most  phosphor- 
ylated  substances  to  enter  cells  readily  makes  it  necessary  to  use  the  analog 
of  the  coenzyme  precursor  if  inhibition  in  cell  preparations  is  to  be  obtained. 
Thus  the  initial  analog  or  any  of  its  metabolic  products  may  interfere  in  a 
number  of  reactions  involving  the  coenzyme,  and  it  is  this  that  militates 
against  facile  interpretations  from  superficially  simple  results.  It  should  also 
be  evident  that  when  the  reaction  catalyzed  by  the  coenzyme-dependent 
enzyme  (A  ^  B  in  Fig.  2-16)  is  determined,  the  kinetics  of  inhibition  by 
an  analog  of  the  coenzyme  precursor  will  generally  not  be  simple  and,  al- 
though the  fundamental  block  may  be  strictly  competitive,  the  quantita- 
tive relationship  between  the  analog  and  the  precursor  will  not  necessarily 
be  competitive. 

The  direct  effect  of  a  coenzyme  analog  on  the  enzyme  reaction  requiring 
the  cooperation  of  the  coenzyme  will  depend  on  the  tightness  with  which 
the  coenzyme  is  bound  to  the  enzyme.  Some  coenzymes  are  so  tightly  bound 
that  they  remain  on  the  enzyme  through  numerous  isolation  procedures, 
and  in  such  cases  the  addition  of  an  analog,  even  though  it  has  a  high 
affinity  for  the  enzyme,  may  not  be  able  to  replace  the  natural  coenzyme 
rapidly  enough  to  induce  inhibition.  It  must  be  remembered  that  the  analog 
does  not  actively  displace  the  coenzyme  (i.e.,  it  does  not  force  it  from  the 
enzyme)  but  only  binds  to  the  free  enzyme;  if  essentially  all  of  the  enzyme 
is  combined  with  coenzyme,  there  is  little  opportunity  for  the  analog  to  act. 
For  this  reason  experiments  on  coenzyme  analogs  are  frequently  done  with 
reconstituted  enzymes.  In  such  cases  the  enzyme  and  coenzyme  are  disso- 
ciated by  some  means  and  the  effect  of  the  analog  on  the  reconstitution  of 
the  active  enzyme  is  investigated,  this  allowing  the  analog  to  act  on  the 
free  enzyme  and  to  demonstrate  competitive  behavior.  This  technique  is 
not,  of  course,  so  applicable  to  cellular  systems. 

It  has  been  frequently  stated  that  coenzyme  analogs  are  specific  inhi- 
bitors. This  is  true  in  one  sense  inasmuch  as  these  analogs  or  their  meta- 
bolic derivatives  appear  to  interfere  only  with  those  enzymes  or  reactions 
involving  the  corresponding  normal  coenzymes,  in  most  instances.  On  the 
other  hand,  the  coenzymes  often  participate  in  several  different  types  of 
metabolic  activity  so  that  the  metabolic  disturbances  produced  by  the 
analogs  may  not  be  specific  with  respect  to  a  single  reaction.  For  example, 
analogs  of  pyridoxal  seem  to  interfere  specifically  with  pyridoxal  metab- 
olism or  the  functions  of  pyridoxal  phosphate,  but  pyridoxal  phosphate 
plays  a  role  in  many  reactions  of  amino  acids  —  racemization,  transamina- 
tion, oxidative  deamination,  decarboxylation,  hydrolytic  cleavage  —  as  well 
as  being  an  important  component  of  other  enzyme  systems,  such  as  muscle 
phosphor ylase,  so  that  a  deficiency  of  pyridoxal  phosphate  can  induce  wide- 
spread disturbances.  In  addition  to  this,  a  generalized  depression  of  amino 


484  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

acid  metabolism  can  secondarily  bring  about  changes  in  systems  not  in- 
volving pyridoxal  phosphate  through  reduction  in  the  concentrations  of 
amino  donors  or  suppression  of  enzyme  synthesis.  Whether  the  coenzyme 
analogs  can  be  considered  as  specific  or  not  will  depend  on  the  complexity 
and  general  metabolic  activity  of  the  preparation  being  studied. 

Most  coenzymes  are  derivatives  of  vitamins  and  it  has  usually  been  anti- 
cipated that  analogs  would  induce  vitamin-deficiency  states.  This  has  been 
demonstrated  in  some  cases;  that  is,  effective  analogs  have  been  found  to 
produce  a  pattern  of  symptoms  roughly  similar  to  those  seen  in  deficiency 
of  the  corresponding  vitamin.  Nevertheless,  it  should  be  clearly  understood 
that  the  situations  are  basically  different.  A  dietary  restriction  of  a  vitamin 
leading  to  a  generalized  depletion  in  the  tissues  would  not  necessarily  bring 
about  functional  changes  identical  to  those  caused  by  an  analog,  which 
could  be  much  more  effective  in  interfering  with  certain  functions  of  the 
coenzyme  than  simple  depletion  and  possibly  leave  other  functions  untouch- 
ed. All  of  the  various  enzymes  binding  a  particular  coenzyme  do  not  have 
the  same  affinities  for  an  analog.  Even  though  the  analog  primarily  inter- 
fered with  the  transport  of  the  vitamin  into  the  cell,  or  blocked  its  further 
metabolism  to  the  active  coenzyme,  it  is  not  justifiable  to  conclude  that  a 
state  of  generalized  depletion  will  result,  because  these  effects  will  pre- 
sumably not  be  exerted  equally  on  all  tissues.  The  differential  penetration 
of  the  analog  into  the  various  tissues  will  perhaps  be  one  important  factor 
in  determining  the  response.  Contrary  to  vitamin  depletion,  analogs  often 
cause  a  rise  in  the  renal  excretion  of  coenzyme  or  its  metabolites,  due  to 
the  displacement  of  the  normal  coenzyme  by  the  analog  in  the  tissues  and 
its  release  from  the  cells.  The  analog  might  also  alter  the  formation  of  the 
coenzyme  from  its  precursors,  or  inhibit  the  metabolism  of  the  active  co- 
enzyme, or  in  some  manner  change  the  renal  excretion  or  resorption  of  the 
coenzyme  or  its  precursors,  so  that  a  variety  of  effects  on  over-all  excretion 
is  possible.  If  it  is  desired  to  demonstrate  metabolic  or  functional  defects 
due  to  an  analog  in  a  short  period  of  time,  it  is  usually  necessary  to  restrict 
the  intake  or  reduce  the  medium  concentration  of  the  coenzyme  or  its  pre- 
cursor, since  the  relationship  between  the  analog  and  the  coenzyme  is  usual- 
ly competitive  or  pseudocompetitive,  but  in  such  cases  one  must  use  the 
coenzyme-depleted  preparation  as  a  control  to  characterize  the  effects  of 
the  analog. 

ANALOGS  OF   NICOTINAMIDE 
AND   THE    PYRIDINE    NUCLEOTIDES 

The  importance  of  nicotinate  and  nicotinamide  is  as  precursors  of  the 
coenzymes  NAD  and  NADP,  and  they  do  not,  as  far  as  is  known,  act  di- 
rectly in  any  metabolic  system,  nor  do  they  usually  occur  in  significant 
concentrations  in  living  cells.  Some  of  the  reactions  involved  in  NAD  syn- 


ANALOGS   OF   NICOTINAMIDE  485 

thesis  and  breakdown  are  shown  in  the  accompanying  diagram.  The  major 
route  of  NAD  formation,  at  least  in  mammalian  tissues,  is  probably  through 
reactions  (l)-(3)  since  the  alternative  pathway  (8)-(10)  is  kinetically  and 
thermodynamically  unfavorable.  Analogs  of  nicotinate  can  thus  either  di- 
rectly inhibit  any  of  these  reactions  or  enter  into  the  reactions  to  form 
abnormal  intermediates,  and  perhaps  analogs  of  NAD  or  NADP,  which 
are  inhibitory.  (See  reactions  on  page  486). 

Inhibition   of  NAD   Nucleosidase   (NADase)   by   Nicotinamide 
and    Related    Compounds 

It  will  be  convenient  to  discuss  first  the  direct  inhibitions  by  simple 
pyridine  derivatives  and  then  proceed  to  those  substances  incorporated  into 
NAD  analogs.  There  is  a  constant  turnover  of  NAD  in  tissues  and  at  least 
a  fraction  of  the  degradative  process  is  attributable  to  NADase,  and  in 
tissue  extracts  or  homogenates  the  splitting  of  NAD  may  be  an  important 
factor  determining  the  dehydrogenase  activity.  Thus  inhibitors  of  NADase 
might  be  expected  under  certain  circumstances  to  protect  the  coenzyme. 
Furthermore,  it  will  be  evident  later  that  the  mechanisms  of  NADase  inhi- 
bition are  involved  in  the  formation  of  abnormal  NAD  analogs.  Mann  and 
Quastel  (1941)  were  the  first  to  observe  an  inhibition  of  NAD  breakdown 
by  nicotinamide.  They  worked  with  brain  suspensions  and  determined  NAD 
by  adding  lactate  dehydrogenase  and  lactate.  Nicotinamide  at  25  mM  was 
found  to  prevent  the  breakdown  of  NAD  almost  completely,  and  addition 
of  nicotinamide  increases  the  respiration  of  various  systems  oxidizing  lactate 
by  preventing  the  destruction  of  NAD.  Nicotinate,  on  the  other  hand,  is 
completely  inactive.  Many  investigators  have  subsequently  used  nicotin- 
amide to  preserve  NAD  in  various  preparations,  often  in  very  high  concen- 
trations and  without  regard  for  the  other  possible  inhibitions  it  might  exert. 
Handler  and  Klein  (1942)  soon  showed  that  NADP  splitting  is  also  inhib- 
ited by  nicotinamide. 

Mcllwain  and  Rodnight  (1949)  pointed  out  that  the  indiscriminate  use 
of  high  nicotinamide  concentrations  to  protect  NAD  in  metabolic  studies 
is  unnecessary,  since  almost  complete  inhibition  of  NADase  is  seen  at  con- 
centrations from  2  to  10  mM  (actually  they  showed  that  2.67  mM  inhibits 
73%).  The  problem  of  the  proper  concentration  of  nicotinamide  to  use  is  a 
difficult  one  because  the  NADases  of  various  tissues  and  organisms  show 
marked  differences  in  susceptibility  to  inhibition.  The  early  work  was  all 
done  on  brain  NADase,  which  is  quite  sensitive,  and  it  has  been  found  that 
some  other  NADases  are  also  sensitive,  e.g.,  from  beef  spleen  (Zatman  et 
al.,  1953).  However,  the  enzymes  from  rabbit  erythrocytes  (Alivisatos  and 
Denstedt,  1952;  Rubinstein  et  al.,  1956;  Malkin  and  Denstedt,  1956),  mouse 
mammary  gland  and  tumor  (Branster  and  Morton,  1956),  and  lupine  seed- 
lings (Hasse  and  Schleyer,  1961)  are  only  moderately  sensitive  to  nicotin- 


486 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


x: 
a. 

o 

o    <D 

>.   d 

a  c 
2   b£ 


« 


a> 


2;     c    g 


^    is 


t-i       ^       o 


e   s 

ni       a! 


M  ^  Q  i!  Q 

rt  -^  _  _  _ 

..    >.  o  ^  eg 

"     C  rt  ^  ^ 

K   rt  —  --  -- 


2  xi 
o  ^ 


9-K       -S. 


.S      3 


CM     CO     in     oj 


ANALOGS   OF   NICOTINAMIDE  487 

amide  {K^s  usually  between  20  and  100  mM).  The  NADase  from  Neuro- 
spora  crassa  is  quite  resistant  to  nicotinamide  (Kaplan  et  al.,  1951). 

The  inhibition  by  nicotinamide  is  competitive  with  respect  to  NAD  for 
the  weakly  inhibited  NADases  of  rabbit  erythrocytes  (Alivisatos  et  al.,  1956; 
Hofmann,  1955),  lupine  seedlings  (Hasse  and  Schleyer,  1961),  and  Neuro- 
spora.  However,  the  inhibition  of  the  sensitive  mammalian  NADases  from 
brain  and  spleen  is  noncompetitive  and  the  elucidation  of  the  mechanism 
by  Zatman  et  al.  (1953)  has  provided  important  information  on  NAD  me- 
tabolism and  its  inhibition  by  a  variety  of  agents.  The  binding  of  nicotin- 
amide to  the  enzyme  is,  however,  readily  reversible  upon  dilution  or  dialysis. 
The  following  reaction  mechanism  was  suggested  as  a  working  hypothesis: 

+NRPPRA  +  enzyme  ^  enzyme  -  +RPPRA  +  N 

l+H.O 
enzyme  +  RPPRA  +  H+ 

An  intermediate  enzyme  complex  which  is  subsequently  hydrolyzed  is  as- 
sumed. The  hydrolysis  is  irreversible  and  N  +  RPPRA  will  not  form  NAD. 
The  inhibition  by  nicotinamide  is  thus  a  competition  with  water  for  the 
enzyme-+RPPRA  complex.  This  complex  is  not  a  Michaelis-Menten  ES 
complex  but  a  covalent-linked  compound  in  which  the  energy  of  the  ni- 
cotinamide— riboside  bond  is  conserved.  The  free  energy  for  the  hydrolysis 
of  this  bond  is  — 8.2  kcal/mole,  and  its  conservation  in  the  complex  is  very 
important  for  the  exchange  reactions  catalyzed  by  this  enzyme.  If  this 
mechanism  is  valid,  one  should  observe  exchange  between  free  nicotinamide 
and  the  nicotinamide  in  NAD,  and  this  was  demonstrated  by  using  nicotin- 
amide-C^*.  These  NADases  might  be  considered  as  transglycosidases  and 
able  to  transfer  the  RPPRA  group  to  compounds  structurally  related  to 
nicotinamide  to  form  NAD  analogs  (Zatman  et  al.,  1954  a).  The  NADases 
which  are  weakly  inhibited  do  not  operate  by  such  a  mechanism  and  do 
not  catalyze  exchange  reactions. 

Another  enzyme  which  is  nicotinamide-sensitive  and  catalyzes  a  similar 
exchange  reaction  is  the  nicotinamide  riboside  phosphorylase  of  human 
erythrocytes  (Grossman  and  Kaplan,  1958  a,  b).  l/w-l/(S)  plots  showed  the 
inhibition  to  be  uncompetitive,  which  is  usually  interpreted  as  a  combina- 
tion of  the  inhibitor  with  the  ES  complex,  but  in  this  case  is  perhaps  due 
to  the  transfer  nature  of  the  reaction.  In  the  scheme: 

ERN  v=^  E  +  RN 

'> 

ER  '^ 

ERP  — >  E  +  RP 

where  nicotinamide  riboside  is  the  substrate  and  — d(Rl^)ldt  =  /-(ERP),  it 
is  seen  that  nicotinamide  will  slow  the  reaction  by  shifting  the  equilibrium 


488 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


in  favor  of  ERN,  and  that  actually  competition  with  phosphate  rather  than 
with  nicotinamide  riboside  might  be  expected.  A  further  complication  is 
the  finding  that  the  exchange  reaction  and  the  sensitivity  to  nicotinamide 
depend  on  a  cofactor,  which  was  isolated  and  shown  to  be  either  ergothio- 
neine  or  a  closely  related  compound.  It  is  possible  that  ergothioneine  acts 
as  a  ribosyl  acceptor  and  this  would  modify  the  kinetics  of  the  nicotin- 
amide inhibition.  It  is  interesting  that  ergothioneine  will  make  the  Neuro- 
spora  NADase  inhibitable  by  nicotinamide. 

A  variety  of  substances  related  to  nicotinamide  or  other  portions  of  the 
NAD  molecule  are  inhibitory  to  NADases  (Table  2-29).  There  is  a  good 
deal  of  variation  in  susceptibility  between  the  different  enzymes.  In  this 
connection  it  may  be  mentioned  that  Handler  and  Klein  (1942)  found  that 
rabbit  brain  NADase  is  readily  inhibited  by  5-10  mM  nicotinamide  but 
not  inhibited  at  all  by  160  mM  picolinate,  quinolinate,  benzamide,  a-amino- 
nicotinate,  trigonelline,  adenine,  adenosine,  or  pyridine.  These  inhibitions 
probably  involve  different  mechanisms.  Some  are  not  competitive  and  the 
inhibitors  probably  participate  in  the  transfer  reaction  as  does  nicotinamide, 
while  others  are  competitive  and  the  inhibitors  are  bound  reversibly  to  sites 
at  the  active  center,  thereby  preventing  the  binding  of  NAD.  The  inhibition 


Nicotine 


€f 


Nikethamide 


CON' 


.C2H5 


COO 


^N 


CH3 

Trigonelline 


,CH, 


4(5)-3'-Pyridyl- 
glyoxaline 


CONHNH, 


CONHNH-CH 


Isoniazid 


CH, 


of  beef  spleen  NADase  by  isoniazid  exhibits  unique  kinetics  inasmuch  as 
an  increase  in  the  NAD  concentration  actually  increases  the  inhibition,  this 
suggesting  that  some  interaction  between  NAD  and  isoniazid  occurs,  the  an- 
alog formed  being  the  active  inhibitor.  Isonicotinamide  inhibits  similarly  to 
isoniazid.  3-Substituted  pyridines  generally  inhibit  somewhat  more  strong- 
ly than  the  4-substituted  compounds.  One  of  the  most  potent  inhibitors  is 


ANALOGS   OF   NICOTINAMIDE  489 

4(5)-3'-pyridylglyoxaline  but  the  mechanism  is  unknown;  it  is  quite  pos- 
sible that  NADases  other  than  from  brain  may  not  be  so  potently  inhibited 
by  this  substance  since  nicotine  does  not  inhibit  the  beef  spleen  enzjTne. 
Another  surprisingly  potent  inhibitor  is  theobromine,  which  is  bound  to 
rabbit  erythrocyte  NADase  around  1  kcal/mole  more  tightly  than  the  other 
purines  tested,  the  inhibition  being  competitive.  Malkin  and  Denstedt 
(1956)  concluded  from  the  inhibition  data  that  NAD  is  attached  to  the 
enzyme  surface  at  the  quaternary  nitrogen  and  the  pyrophosphate  group. 
It  is  rather  strange  that  adenine  is  a  reasonably  effective  inhibitor,  whereas 
adenosine  or  the  adenine  nucleotides  are  much  weaker  or  completely  with- 
out action,  since,  if  adenine  were  bound  in  the  same  position  as  it  is  when 
part  of  the  NAD  molecule,  one  might  expect  ribose  and  phosphate  groups 
to  augment  the  binding.  Thus  the  inhibition  by  adenine  and  other  purines 
may  involve  interaction  with  the  enzyme  surface  in  a  manner  unrelated  to 
the  normal  binding  of  the  purine  component  of  NAD.  This  is  further  borne 
out  by  the  studies  on  multiple  inhibition  by  Hofmann  and  Rapoport  (1957) 
(see  accompanying  tabulation),  inasmuch  as  adenine  does  not  add  to  the 
inhibitions  produced  by  inhibitors  presumably  interacting  with  enzyme 
groups  binding  NAD. 


Inhibitors 

(I)/(S) 

%  Inhibition 

Nicotinamide 

50 

64 

Adenine 

9 

20 

Both 

65 

NMN 

10 

44 

Adenine 

9 

20 

Both 

43 

NADP 

4.5 

40 

Adenine 

9 

20 

Both 

39 

3-Acetylpyridine  and   the   Formation   of  Analogs  of  NAD 

In  a  search  for  pjTidine  derivatives  which  might  have  vitamin  activity 
against  black  tongue  in  dogs,  Woolley  et  al.  (1938)  observed  that  3-acetyl- 
pyridine  is  not  only  ineffective  but  kills  nicotinamide-deficient  animals  in 
1  day,  normal  dogs  being  unaffected.  3-Acetylpyridine  rapidly  produces 
signs  of  nicotinate  deficiency  in  mice  and  at  the  LDjq  (around  3  mg  per  day) 
the  animals  succumb  in  3  to  4  days  (Woolley,  1945  b).  The  effects  produced 
are:  rapid  respiration,  motor  incoordination  followed  by  complete  paralysis, 
emaciation,  and  inflamed  skin  and  tongue.  The  mice  can  be  completely 


490 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


P5 


."^ 


c 


O 


o   o    o   o 


iz;        A 


^ 


S       =^ 


'C 

i=l 

« 

e 

cS 

fl 

cS 

S 

pS 

(D 

« 

"p* 

m 

-iil 

■=? 

c8 

o 

w 

w 

s 

O  -in 


■ft 


d    d 


Ph 


m 

0) 

i 

a 

73 

■» 

m 

C 

g 

o 

CI 

O 

C 

t3 

CM 

'-J3 
O 
o 

■-3 

8 

<: 

<1 

<1 

< 

S 

5? 

Ph 


pq 


ANALOGS   OF   NICOTINAMIDE  491 


o  ^- 

la  o 

I— I  o> 


lOOOOO  OO^C^lO^MCXioOOOOO-^O 


ooooo  o  ooooo  oo 


m  cS  3  ID 

V3  »  ID  ©  "2  e3  05 

.2  -O  73  '-^  .2  "43  '-S             -S      O     '«     ^                                  Q 

§  "^  1  .2  §  .1  I           :^    -^    '^    g                              § 


c 
c 

o 

n 

bl) 

;>, 

<» 

;3 

T) 

Ti 

© 

^ 

s 

eS 

Ul 

PO 

>> 

^ 

« 

Ph 

C! 

o 

fO 

^ 
^ 

H 

t-l 

492 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


P5 


^ 


Q    o 


■*      lO      »C     GO     OJ 
S<1     (M      (M     C^     (M 


o    o    o    o    o 
'^    ^    iri    00    >o 

^      r-H      ^      ^ 


o    o    o    o    o    o    o 


CO     CD     CO     CO     CO     CD     ^ 

<6  S  S  <6  S  S  o 


c 

c 

o 

'd 

sc 

S 
2 

o 

1 

ft 
o 

cS 

1 

CD 

S 

o 

o 

0) 

IS 

4) 

a 
o 

i 
o 

(P 

aj 

IB 

C 

>. 

c 

(» 

« 

5ti 

^ 

•n 

ji 

o 

j3 

ci3 

^ 

T3 

03 

H 

< 

H 

W 

H 

g 

X 

'SJ 

-< 

o 

Ph 


cS      «     .S 


P5 


<1    < 


00 
O 

P. 

o 

>. 


ANALOGS   OF   NICOTINAMIDE  493 


o 
o 


35 

Ti 

l-H 

c 

... 

cS 

C 

C 

<U 

^.^ 

ert 

-k^ 

fO 

S 

c 

OS 

-^  ■*       o   o   o   o 

0«     tJ4     CO     oo 


d  d    d    d    d       d       d 


03 

'S 

2 

j3 

03 

^ 

0) 

O 

o 

o 

TS 

o 

r: 

C 

N 

fi 

4^ 
O 

o 

2 
'S 
o 

o 

^ 

1— 1 

^ 

S 

^: 

cS      .S      ^      .S 

^     ^    ■ "     ^                         .S  -5  —  W 

'S  C  "o  CL|      p 

.S     5    .2    .2     o    .2     «    S               -§  ^  2  P    <1 


>> 

a 

aj 

l» 

^ 

40 

Si 

03 

Si 

3 

^  ,c        03    a 

=1  1       ,2  5 


494  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

protected  by  providing  nicotinate  or  nicotinamide  in  the  diet.  On  the  other 
hand,  yeast  and  most  bacteria  seem  to  be  quite  resistant  to  3-acetylpyri- 
dine,  although  the  growth  of  Lactobacillus  casei  in  nicotinate-free  medium 
is  inhibited  around  50%  at  16.5  mM,  a  depression  that  can  be  reversed  by 
nicotinate  but  not  by  nicotinamide.  Chick  embryos  are  killed  by  450-600  //g 
3-acetylpyridine  injected  into  the  eggs  and  sublethal  doses  cause  distur- 
bances in  embryogenesis  (Ackermann  and  Taylor,  1948),  These  effects  can 
be  completely  reversed  by  nicotinamide;  even  6000  jugjegg  of  the  analog 
can  be  counteracted  by  380  //g  of  nicotinamide,  indicating  a  competitive 
relationship.  Changes  in  the  heart,  characteristic  of  nicotinate  deficiency, 
are  produced  by  perfusion  of  1.6-8  mM  3-acetylpyridine  through  the  iso- 
lated rabbit  heart,  dysrhythmias  and  a-v  block  occurring  within  30  min 
(Braun,  1949).  Subsequent  perfusion  with  nicotinamide  reverses  these  ef- 
fects but  the  concentration  must  be  around  100  times  that  of  the  analog. 
These  early  observations  all  point  to  the  interference  by  3-acetylpyridine 
in  the  metabolism  or  function  of  nicotinate  or  nicotinamide.  If  it  is  assumed 
that  the  primary  role  of  these  metabolites  is  the  formation  of  the  NAD  and 
NADP  coenzymes,  the  following  possible  mechanisms  for  inhibition  by  3- 
acetylpyridine  might  be  imagined.  (1)  Inhibition  of  some  step  in  the  syn- 
thesis of  NAD  [especially  reactions  (1)  to  (3)  in  the  scheme  on  page  486], 
(2)  inhibition  of  the  interconversion  of  nicotinate  and  nicotinamide,  (3)  en- 
trance into  one  of  the  pathways  of  nicotinate  metabolism  to  form  inhibitory 
intermediates,  (4)  formation  of  an  NAD  analog,  either  through  the  normal 
pathway  or  by  the  exchange  reaction  catalyzed  by  NADase,  (5)  inhibition 
of  NADases  and  related  enzymes,  and  (6)  direct  interference  with  NAD  or 
NADP  to  inhibit  dehydrogenase  activity. 

It  will  be  well  to  consider  certain  aspects  of  the  metabolism  of  3-acetyl- 
pyridine before  taking  up  the  problem  of  how  this  analog  induces  its  inhi- 
bitory effects.  3-Acetylpyridine  at  doses  around  0.5  g/day  increases  the  urin- 
ary excretion  of  iV-methylnicotinamide  in  both  normal  and  nicotinate-de- 
ficient  dogs  (Gaebler  and  Beher,  1951).  The  iV-methylnicotinamide  could 
arise  either  from  a  disturbance  of  nicotinamide  metabolism,  since  iV-meth- 
ylation  is  an  important  reaction  in  the  elimination  of  nicotinamide,  or 
directly  from  the  3-acetylpyridine.  The  oxidation  of  3-acetylpyridine  to 
nicotinate  might  be  anticipated  because  benzoate  is  formed  from  acetophe- 
none  in  the  tissues,  the  entire  sequence  being 

3-acetylpyridine   ->■   nicotinate   ->   nicotinamide   ->   iV-methylnicotinamide. 

It  was  found  that  the  latter  explanation  is  correct  by  determining  labeled 
iV-methylnicotinamide  formed  from  labeled  3-acetylpyridine  (Beher  et  al., 
1952).  It  was  pointed  out  that  in  the  course  of  its  oxidation  and  methylation 
the  analog  might  also  interfere  with  nicotinate  metabolism.  Since  iV-meth- 
ylnicotinamide  accounts  for  only  about  10%  of  the  administered  3-acetyl- 


ANALOGS   OF   NICOTINAMIDE  495 

pyridine,  other  excretory  products  were  investigated  and  increased  urinary 
nicotinate  and  various  glucuronides  were  found  (Beher  and  Anthony,  1953). 
No  urinary  3-acetylpyridine  could  be  detected.  An  interesting  suggestion 
that  the  oxidation  of  3-acetylpyridine  may  involve  NAD(P)  enzymes  was 
made;  this  might  mean  that  in  nicotinate-deficient  animals,  where  NAD(P) 
levels  are  low,  the  oxidation  of  3-acetylpyridine  would  be  impaired  and  the 
analog  would  be  more  toxic.  3-Acetylpyridine  presents  the  strange  situation 
wherein  the  analog  is  detoxified  to  the  normal  metabolite,  and  this  would 
presumably  tend  to  counteract  the  inhibitory  effects.  In  low  dosage  (25- 
60  mg/day),  3-acetylpyridine  can  protect  against  black  tongue  in  dogs  but 
at  higher  dosage  it  can  create  a  nicotinate  deficiency  (McDaniel  et  al.,  1955). 
Animals  may  have  a  limited  ability  to  oxidize  3-acetylpyridine;  small 
amounts  are  mainly  oxidized  and  little  3-acetylpyridine  is  left  to  inhibit, 
whereas  the  larger  doses  exceed  the  metabolic  capacity  of  the  system.  This 
is  indicated  by  the  results  of  Guggenheim  and  Diamant  (1958),  who  deter- 
mined the  excretion  of  iV-methylnicotinamide  in  rats  given  comparable  doses 
of  nicotinamide  and  3-acetylpyridine  (see  tabulation).  Beyond  a  dose  of  50 


Dose 

iV-Methylnicotinamide  excretion 

from: 

(mg/kg) 

Nicotinamide 

3  -  Acetylpyridine 

Ratio  " 

0 

23 

23 

10 

65 

40 

2.5 

20 

182 

105 

1.9 

50 

225 

120 

2.1 

100 

700 

218 

3.5 

200 

1950 

376 

5.5 

"  Ratio  calculated  after  substracting  endogenous  excretion. 

mg/kg  there  seems  to  be  relatively  less  oxidation  of  the  analog.  Adminis- 
tration of  3-acetylpyridine-C^*H3  to  rats  leads  to  30%  of  the  activity  ex- 
pired as  CO2  and  44%  eliminated  in  the  urine  during  24  hr  (Beher  et  al., 
1959);  since  the  total  dosage  was  probably  around  100  mg/kg,  smaller  doses 
might  be  even  more  efficiently  oxidized.  3- Acetylpyridine  is  also  partially 
metabolized  to  NAD  and  the  3-acetylpyridine  analog  of  NAD,  as  will  be 
discussed  shortly.  Finally,  nicotinamide  mononucleotide  excretion  is  aug- 
mented by  3-acetylpyridine  and  it  is  possible  that  this  mainly  originates 
directly  from  the  analog  (McDaniel  et  al.,  1955).  The  metabolism  of  3-acetyl- 
pyridine and  the  compounds  derived  from  it  thus  depend  on  the  species, 
the  dose,  and  whether  the  animals  are  normal  or  nicotinate-deficient. 

We  shall  now  examine  the  effects  of  3-acetylpyridine  on  the  tissue  levels 
of  NAD,  the  formation  of  NAD  analogs,  and  the  enzymic  activities  of  these 


496  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

analogs.  Most  of  this  work  has  been  done  by  Kaplan  and  his  associates  at 
Johns  Hopkins,  and  a  summary  of  their  most  important  results  will  be 
given.  It  was  first  demonstrated  that  the  incubation  of  NAD,  brain  NADase, 
and  isonicotinyl  hydrazide  (isoniazid,  IHN)  leads  to  the  formation  of  the 
INH  analog  of  NAD,  which  was  isolated  in  good  yield,  and  it  was  postulated 
that  the  antitubercular  activity  of  isoniazid  may  be  related  to  the  appear- 
ance of  this  nonfunctional  or  inhibitory  analog  (Zatman  et  al.,  1954  b). 
It  was  soon  shown  that  a  variety  of  pyridine  derivatives  can  exchange  with 
nicotinamide  in  the  presence  of  certain  NADases  to  form  NAD  analogs; 
these  include  isonicotinamide,  iproniazid,  ethylnicotinate,  and  S-acetylpyri- 
dine  (N.  0.  Kaplan  c^oi.,  1954).  The  formation  of  3-AcPyr-NAD*  in  tissue 
homogenates  and  whole  animals  is  inhibited  by  nicotinamide.  The  exchange 
reaction  and  hydrolysis  may  be  represented  as: 


NRPPRA  +  E 

^ 

E-NRPPRA 

E-RPPRA  -> 

E  +  RPPRA 

ti-'^ 

XRPPRA  +  E 

:i± 

E-XRPPRA 

where  NRPPRA  is  NAD,  XRPPRA  is  the  NAD  analog,  E-RPPRA  is  the 
relatively  stable  ribosyl  enzyme  complex,  and  X  is  the  pyridine  derivative 
exchangeable  with  nicotinamide.  The  over-all  exchange  reaction  would  be: 

NRPPRA  +  X  ^  XRPPRA  +  N 

Injection  of  3-acetylpyridine  leads  to  a  rise  in  total  pyridine  nucleotides  in 
most  tissues;  in  the  liver  this  is  NAD  and  none  of  the  analog  is  demon- 
strable, due  presumably  to  the  oxidation  of  3-acetylpyridine  to  nicotinate, 
whereas  in  brain,  spleen,  and  tumors  3-AcPyr-NAD  appears.  In  tumors  the 
NAD  content  actually  decreases  as  3-AcPyr-NAD  increases.  The  equilibrium 
between  NAD  and  any  of  its  analogs,  and  the  ratio  of  their  concentrations 
in  a  particular  tissue,  will  depend  on  (1)  the  AF  between  NAD  and  the 
analog,  (2)  the  concentrations  of  N  and  X,  (3)  the  rate  of  transformation 
of  X  to  nicotinate,  if  it  occurs,  and  (4)  the  relative  bindings  of  NAD  and 
its  analogs  to  the  dehydrogenases.  The  time  courses  for  the  formation  of 
3-AcPyr-NAD  from  3-acetylpyridine  and  the  toxic  reactions  led  to  the  sug- 
gestion that  the  toxic  and  lethal  actions  are  related  to  the  NAD  analog; 
whether  the  toxicity  depends  on  a  reduction  of  NAD  or  a  rise  in  3-AcPyr- 
NAD  was  undecided. 

In  order  to  determine  the  nature  of  the  effects  of  3-acetylpyridine  on 
tissue  metabolism,  it  will  be  necessary  to  consider  the  ability  of  3-AcPyr- 

*  The  analogs  of  NAD  will  be  designated  by  prefixes  of  this  type,  following  Kaplan, 
since  this  is  convenient  if  not  exactly  accurate. 


ANALOGS   OF   NICOTINAMIDE  497 

NAD  to  replace  NAD  as  the  coenzyme  for  the  various  dehydrogenases. 
3-AcPyr-NAD  can  function  in  most  NAD-dependent  dehydrogenase  reac- 
tions. In  some  cases  it  can  be  reduced  more  rapidly  than  NAD  (horse  liver 
alcohol  dehydrogenase,  beef  liver  glutamic  dehydrogenase,  Lactobacillvs  d- 
and  L-lactate  dehydrogenases)  and  in  other  cases  proceeds  more  slowly 
(yeast  alcohol  dehydrogenase,  beef  heart  lactate  dehydrogenase,  yeast  gly- 
ceraldehyde-3-P  dehydrogenase),  while  in  a  few  instances  the  rates  are  ap- 
proximately equivalent  (rabbit  muscle  lactate  dehydrogenase)  (N.  0.  Kap- 
lan et  al,  1956;  van  Eys  et  al,  1958;  N.  0.  Kaplan,  1959;  Stockell,  1959). 
3-AcPyr-NADP  is  reduced  about  one  fifth  as  fast  as  NADP  in  the  pig 
heart  isocitrate  dehydrogenase  system  and  is  inactive  in  erythrocyte  glu- 
cose-6-P  dehydrogenase  (N.  0.  Kaplan  et  al.,  1956;  Marks  et  al.,  1961).  The 
relative  rates  do  not  necessarily  reflect  the  relative  bindings  to  the  dehy- 
drogenases. In  those  cas^s  where  coenzyme  activity  is  low  but  binding  is 
appreciable,  the  NAD  or  NADP  analogs  can  inhibit  the  dehydrogenases; 
thus  3-AcPyr-NAD  inhibits  glucose-6-P  dehydrogenase  quite  strongly  {K,= 
0.03  mM)  and  this  is  competitive.  NAD  analogs  other  than  3-AcPyr-NAD 
are  usually  less  active  and  tend  to  be  more  inhibitory.  Thionicotinamide- 
NAD,  nicotinyl-hydroxamate-NAD,  and  nicotinyl-hydrazide-NAD  com- 
petitively inhibit  lactate  and  alcohol  dehydrogenases,  whereas  3-benzoyl- 
pyridine-NAD  inhibits  beef  heart  lactate  dehydrogenase  uncompetitively 
(Anderson  and  Kaplan,  1959).  The  introduction  of  3-acetylpyridine,  or 
other  pyridine  analogs,  can  thus  produce  several  effects  on  tissue  dehydro- 
genase activity,  and  in  the  general  case  will  bring  about  an  imbalance  of 
the  normal  relative  substrate  oxidations,  due  to  altering  the  rates  of  the 
various  dehydrogenases  in  different  ways.  Unfortunately  there  has  not  yet 
been  sufficient  study  of  the  oxidative  abilities  of  tissues  isolated  from  ani- 
mals treated  with  3-acetylpyridine.  However,  it  is  probably  safe  to  assume 
that  at  least  a  major  cause  of  the  toxic  effects  is  the  inhibition  of  certain 
dehydrogenases  by  the  3-AcPyr-NAD  formed. 

The  various  NAD  analogs  have  been  very  useful  in  demonstrating  differ- 
ences between  dehydrogenases  from  different  tissues  or  species.  For  exam- 
ple, beef  heart  and  rabbit  muscle  lactate  dehydrogenases  react  better  with 
NAD  than  with  3-AcPyr-NAD,  but  the  lactate  dehydrogenases  from  lob- 
ster heart  and  thorax  muscle  react  better  with  the  analog  (N.  0.  Kaplan, 
1959).  Kaplan  et  al.  (1960)  have  pointed  out  that  the  molecular  heterogeneity 
of  enzyme  active  centers  has  phylogenetic  significance.  It  is  possible  to 
classify  animals  with  respect  to  the  affinities  of  their  dehydrogenases  for 
the  coenzymes  or  their  analogs,  and  it  is  hoped  that  further  investigation 
along  these  lines  will  elucidade  some  of  the  evohitionary  problems  relative 
to  the  changes  in  the  active  center  configurations. 

We  must  now  examine  the  evidence  for  other  sites  of  action  for  3-acetyl- 
pyridine and  related  analogs,  Mcllwain  (1950)  reported  that  3-acetylpyri- 


498 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


dine  inhibits  spinal  cord  NADase  65%  at  11  laM.  Such  inhibition  could  be 
due  to  (1)  direct  competition  with  NAD,  (2)  reaction  with  E-RPPRA  to 
form  a  relatively  stable  complex,  thereby  depleting  free  enzyme,  or  (3)  in- 
hibition by  a  3-AcPyr-NAD  analog  formed.  In  the  case  of  the  INH-NAD 
analog,  the  inhibition  seems  to  be  mainly  of  the  third  type  (Zatman  et  al., 
1954  b).  Nicotinamide  deaminase  is  inhibited  quite  well  by  3-acetylpyri- 
dine;  the  inhibition  is  competitive  and  50%  at  (I)/(S)  =  20  (Grossowicz 
and  Halpern,  1956  b).  Yeast  alcohol  dehydrogenase  is  inhibited  directly 
by  substituted  pyridines  (van  Eys,  1956;  van  Eys  and  Kaplan,  1957  a). 
The  inhibitions  are  related  to  the  pK„'s  of  the  analogs  (Table  2-30)  and  a 

Table  2-30 
Inhibition  of  Yeast  Alcohol  Dehydrogenase  by  Substituted  Pyridines" 


Substituted 
pyridine 

Concentration   for   50%   inhibition    (mil/) 

pA'a 

Total  base 

Pyridinium  ion 

iV -methyl 
derivatives 

4-CH3 

6.11 

20 

0.013 

3-CH3 

5.82 

40 

0.013 

5.8 

None 

5.27 

70 

0.0066 

5.5 

3-CONH2 

3.40 

230 

0.00029 

5.0 

3-COCH3 

3.39 

— 

— 

4.0 

3-CHO 

3.37 

— 

— 

4.2 

3-COOC2H5 

2.24 

300 

0.000026 

3.6 

3-CN 

1.45 

600 

0.0000085 

3.2 

3-SO3- 

2.9 

13 

0.0000051 

— 

"  From  van  Eys  and  Kaplan  (1957  a). 


straight  line  is  obtained  by  plotting  pK^  against  p^^.  The  pyridinium  ions 
are  presumably  the  active  inhibitors.  The  iV-methyl  derivatives  are  rela- 
tively weak  inhibitors.  The  pyridine  N  must  be  important  for  the  binding, 
its  properties  being  altered  by  the  substituents  (the  stronger  the  electro- 
negativity of  the  substituent,  the  greater  the  inhibition).  It  is  thus  evident 
from  these  data  that  the  pyridine  analogs  can  inhibit  various  enzymes  di- 
rectly; it  is  likely  that  these  effects  are  not  as  important  as  those  arising 
from  the  corresponding  NAD  analogs  in  whole  animals. 

If  the  major  actions  of  3-acetylpyridine  are  mediated  through  3-AcPyr- 
NAD,  the  susceptibility  of  microorganisms  or  animals  to  3-acetylpyridine 


ANALOGS   OF   NICOTINAMIDE  499 

will  depend  primarily  on  the  exchange  activity  of  the  NADases  present, 
and  perhaps  secondarily  on  the  ability  to  oxidize  the  3-acetylpyridine  to 
nicotinate. 

Some  of  the  effects  of  3-acetylpyridine  on  tissue  functions  and  whole 
animals  were  mentioned  at  the  beginning  of  this  section,  and  some  of  the 
more  recently  studied  actions  will  now  be  discussed.  The  LDjq  for  the  in- 
traperitoneal route  is  300-350  mg/kg  in  mice  and  80  mg/kg  in  rats  (Cogge- 
shall  and  MacLean,  1958).  Hicks  (1955)  found  that  the  administration  of 
3-acetylpyridine  to  rats  and  mice  at  doses  around  the  LD50  produces  ne- 
crosis of  adrenal  medulla,  of  certain  neurons  in  the  supraoptic  nucleus  of 
the  hypothalamus,  and  of  the  pyramidal  layer  of  the  hippocampus.  No 
effects  on  the  cerebral  cortex  were  observed,  contrary  to  the  action  of  most 
metabolic  inhibitors.  These  effects  are  not  seen  in  nicotinate  deficiency, 
but  the  picture  may  represent  a  more  accelerated  and  acute  deficiency; 
it  is  possible  that  the  regions  affected  are  more  dependent  on  the  pyridine 
nucleotides,  but  differential  penetration  might  also  be  a  factor.  Coggeshall 
and  MacLean  (1958)  found  that  single  LD50  doses  to  rats  lead  to  weakness 
of  the  extremities,  inspiratory  rhonchi,  urinary  incontinence,  and  other 
symptoms,  but  gross  pathological  examination  of  the  organs  showed  no- 
thing remarkably  abnormal.  Surviving  mice  show  motor  incoordination  and 
a  slight  to  complete  loss  of  neurons  in  the  hippocampal  areas  CA3  and  CA4; 
some  damage  to  other  hippocampal  areas  and  the  dentate  gyrus  may  occur, 
but  no  changes  in  other  brain  areas  were  detected.  It  was  concluded  that 
the  hippocampus  must  be  metabolically  different  from  the  rest  of  the  brain. 

Rats  given  100  mg/kg  of  3-acetylpyridine  develop  ataxia,  hyperkinesis, 
and  convulsions  and  it  was  found  that  5-10%  of  the  total  brain  pyridine 
dinucleotides  is  3-AcPyr-NAD  (Brunnemann  et  al.,  1962).  The  maximal 
levels  of  the  abnormal  analogs  occur  at  6-8  hr  and  various  regions  of  the 
brain  differ  in  the  fraction  incorporated,  the  highest  levels  being  found  in 
the  hippocampus  (Herken  and  Neuhoff,  1963;  Willing  et  al.,  i964).  The  ad- 
ministration of  4-acetylpyridine  does  not  lead  to  incorporation  or  to  toxic 
symptoms.  This  Berlin  group  of  workers  favors  the  concept  that  the  cen- 
tral neurological  effects  of  3-acetylp>Tidine  are  due  primarily  to  interference 
in  electron  transport  as  a  result  of  the  inhibitions  produced  by  the  3- AcPyr- 
NAD(P)  formed  in  the  brain. 

Hollander  and  I  have  studied  the  effects  of  the  acetylpyridines  on  the 
isolated  rat  atrium  and,  although  the  work  is  not  yet  complete,  the  basic 
actions  are  clear.*  3-Acetylpyridine  at  1  mM  increases  the  contractile  ten- 
sion of  the  atria  10%  and  simultaneously  the  resting  and  action  potential 
magnitudes  are  increased  4-5%.  The  action  potential  duration  and  con- 

*  We  have  found  that  most  commercial  samples  of  the  acetylpyridines  are  quite 
impure  and  redistillation  under  reduced  pressure  is  necessary  to  obtain  reliable  pre- 
parations. 


500  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

duction  rate  are  either  not  affected  or  slightly  decreased.  The  effects  are 
somewhat  greater  at  5  niM,  contractile  tension  increasing  25%.  2-Acetyl- 
pyridine  and  4-acetylpyridine  produce  very  much  the  same  effects,  and 
nicotinamide  itself  quite  potently  stimulates  atrial  contractility,  although 
in  this  case  the  resting  and  action  potentials  tend  to  decrease.  In  view  of 
these  effects  by  substances  not  giving  rise  to  NAD  analogs,  the  acute  ac- 
tion of  3-acetylpyridine  on  the  atria  must  be  attributed  to  some  other  me- 
chanism. The  lack  of  depressant  activity  at  these  concentrations  is  also 
interesting,  since  metabolic  disturbances  invariably  produce  certain  char- 
acteristic changes.  The  contractile  stimulation  by  nicotinamide  increases 
with  concentration  and  at  100  mM  is  around  100%.  The  mechanism  is  not 
understood  but  seems  to  be  unrelated  to  membrane  potential  changes. 

Inhibition   of  Dehydrogenases   by   Nicotinamide  and    Related   Compounds 

The  inhibition  of  NADases  and  inhibitions  dependent  on  the  NADase- 
catalyzed  exchange  reactions  have  been  discussed.  We  now  turn  to  the 
inhibitions  of  NADP-dependent  dehydrogenases  by  nicotinamide  and  other 
substituted  pyridines.  The  groups  and  interactions  involved  in  the  binding 
of  the  pyridine  coenzymes  to  the  dehydrogenases  have  been  discussed  by 
Shifrin  and  Kaplan  (1960).  Sulfhydryl  groups  are  often  important  and  have 
frequently  been  thought  to  react  with  the  pyridine  N  of  the  coenzymes, 
while  in  some  dehydrogenases  Zn++  is  involved  and  perhaps  reacts  with 
the  phosphate  or  adenine  residues.  It  is  apparent  that  the  coenzymes  are 
bound  in  different  ways  to  different  dehydrogenases  and  this  will  determine 
to  some  extent  the  ability  of  analogs  to  inhibit.  The  nature  of  the  very 
tightly  bound  intramitochondrial  NAD  in  unknown,  but  possibly  it  is  much 
more  difficult  to  inhibit  intact  mitochondrial  dehydrogenases  than  the  iso- 
lated and  often  reconstituted  enzymes  usually  studied. 

The  first  report  of  an  inhibition  of  metabolism  by  nicotinamide  was  made 
by  Baker  et  al.  (1938)  in  connection  with  a  study  of  the  action  of  nicotine 
on  cerebral  respiration.  However,  the  inhibition  of  brain  slice  respiration  is 
slight,  30  mM  depressing  the  oxidation  of  glucose  8%  and  lactate  15%. 
A  greater  differential  effect  on  glucose  and  lactate  oxidation  is  exerted  by 
nicotinate,  the  inhibitions  at  30  mM  being  9%  and  57%,  respectively.  Von 
Euler  (1942)  made  the  initial  investigation  of  dehydrogenase  inhibition  and 
found  inhibitions  of  both  beef  liver  glucose  dehydrogenase  and  heart  lactate 
dehydrogenase  (see  accompanying  tabulation).  The  generally  weak  inhi- 
bitory activity  and  the  greater  potencies  of  nicotinate  and  pyridine-3-sulf- 
onate,  compared  to  nicotinamide  and  pyridine-3-sulfonamide,  may  indicate 
that  these  substances  do  not  actually  combine  with  the  pyridine-binding 
site  on  the  enzymes,  and  it  is  even  doubtful  if  they  may  be  classed  as  NAD 
or  NADP  analogs.  However,  the  inhibition  by  the  pyridine-3-sulfonate  ap- 
pears to  be  competitive.  Brink  (1953  b)  continued  von  Euler's  work  in 


ANALOGS   OF   NICOTINAMIDE 


501 


Inhibitor 


Relative  "KT  " 


Lactate  DH 


Glucose  DH 


Salicylate 

p-Aminobenzoate 

Salicylamide 

Pyridine-3-sulfonate 

Nicotinate 

Benzenesulfonate 

m  -  Aminobenzoate 

Benzoate 

Pyridine-3-sulfonamide 

Nicotinamide 

Benzamide 

Trigonelline 

Adenosine 

Adenosine-3'-P 

ATP 


7.7 

— 

14.1 

8.6 

14.2 

— 

22.1 

27.6 

32.6 

12.0 

43.4 

34.1 

51.7 

— 

54.2 

75.2 

60.0 

135 

105 

41.5 

— 

46.3 

inhibition 

No  inhibition 

7.2 



12.6 

— 

21.0 



"  The  relative  "K/'  values  were  calculated  from  the  inhibitions  at  varying  con- 
centrations in  order  that  the  relative  potencies  of  the  inhibitors  could  be  more  readily 
compared;  they  are  not  absolute  values. 

Stockholm  but  determined  the  inhibitor  constants  by  plotting,  so  that  the 
values  in  the  accompanying  tabulation  for  beef  liver  glucose  dehydrogenase 


Inhibitor 


Ki  Relative  —  ZiF  of  binding 

(mM)  (kcal/mole) 


4-Pyridoxate 

Pyridoxal 

3-Hydroxypyridine 

Nicotinate 

2-Methylnicotinat« 

Nicotinamide 

Isonicotinyl  hydrazide 

Pyridine 

Isonicotinate 

Trigonelline 

ATP 

Adenosine-3'-P 

Adenosine 

Adenine 

Phosphate 


0.3 

4.98 

0.8 

4.39 

9.2 

2.89 

21.5 

2.36 

25 

2.26 

23 

2.32 

200 

0.99 

300 

0.74 

inhibition 

inhibition 

1.75 

3.92 

3.5 

3.48 

9.25 

2.89 

12 

2.73 

810 

0.13 

502 


2.  ANALOGS  OF  ENZYME  EEACTION  COMPONENTS 


are  more  reliable  than  in  the  tabulation  above.  The  K^^^  for  NAD  is  0.00428 
mM  (relative  —AF  would  be  7.63  kcal/mole)  so  that  none  of  these  inhi- 
bitors are  bound  nearly  so  tightly.  Substitution  in  the  3-position  of  the  pyri- 


coo 


CONH, 


Nicotin- 
amide 


SOpNH, 


Pyridine-3- 
sulfonamide 


CONHNH, 


Isonicotinyl 
hydrazide 


OH 


CHO 


COO 


COO 


3 -Hydroxy - 
pyridine 


Salicylate 


CHoOH 


CH,OH 


dine  ring  is  necessary  for  significant  inhibition,  but  the  nature  of  this  group 
can  vary  considerably  and  certainly  no  marked  electrostatic  interaction  is 
involved.  The  pyridine  N  would  not  seem  to  be  of  much  importance  in  the 
binding,  since  benzamide  is  about  as  inhibitory  as  nicotinamide,  and  ben- 
zenesulfonate  almost  as  potent  as  pyridine-3-sulfonate,  and  yet  A"-methyl- 
ation  (to  form  trigonelline)  abolishes  the  inhibition.  The  extra  2.1  kcal/mole 
provided  by  the  3-hydroxy  group  suggests  the  possibility  of  hydrogen  bond- 
ing to  the  enzyme  from  this  position,  but  dipolar  and  dispersion  interactions 
could  also  account  for  this.  It  may  be  noticed  that  the  results  with  alcohol 
dehydrogenase  (Table  2-30)  are  in  certain  respects  different  than  those 
with  glucose  dehydrogenase.  In  this  connection  one  must  remember  that 
the  ionization  constants  of  these  analogs  should  be  considered,  and  it  may 
be  that  the  major  effect  of  the  substituent  groups  is  by  modification  of  the 
p^a  of  the  pyridine  N.  Until  these  problems  have  been  treated  quantita- 
tively, it  is  impossible  to  evaluate  accurately  the  relationship  between 
structure  and  inhibitory  activity.  In  any  event,  it  is  clear  that  the  major 
binding  energy  of  NAD  is  contributed  by  the  adenine  nucleotide  portion 
of  the  molecule,  so  that  pyridine  derivatives  might  not  be  expected  to  be 
potent  inhibitors  of  dehydrogenases.  The  relatively  strong  inhibitions  pro- 
duced by  pyridoxal  and  4-pyridoxate  may  be  significant  for  cellular  meta- 
bolic regulation  and  further  study  of  the  effects  of  these  substances  on 
various  dehydrogenases  is  probably  warranted. 

Nicotinamide  has  been  frequently  used  in  homogenates  to  inhibit  the 
splitting  of  NAD  by  NADases,  as  discussed  above,  and  often  at  concentra- 


ANALOGS   OF   NICOTINAMIDE  503 

tions  sufficiently  high  to  interfere  with  dehydrogenases.  Feigelson  et  al. 
(1951)  investigated  this  problem  in  liver  homogenates  and  noted  first  that 
nicotinamide  reduces  endogenous  respiration,  an  effect  reversed  by  NAD. 
At  50-100  mM,  nicotinamide  stimulates  the  endogenous  respiration  some- 
what, perhaps  due  to  protection  of  NAD,  but  at  higher  concentrations  in- 
hibits quite  potently.  Malate  dehydrogenase  was  partially  purified  and  ni- 
cotinamide inhibited  competitively  with  respect  to  NAD  with  K^  —  0.00367 
mM,  and  K,  =  113  raM,  corresponding  to  about  5  kcal/mole  tighter  bind- 
ing for  the  NAD.  Care  must  thus  be  used  in  the  choice  of  nicotinamide  con- 
centration when  NADase  inhibition  is  desired.  Results  with  lactate  and 
glucose-6-P  dehydrogenases  from  rabbit  erythrocytes  are  verj'  similar  (Ali- 
visatos  and  Denstedt,  1952).  Nicotinamide  inhibits  competitively  with  K^ 
around  100  mM.  It  was  also  shown  that  incubation  of  the  apoenzyme  with 
nicotinamide  in  the  absence  of  NADP  leads  to  progressive  irreversible  inac- 
tivation  of  the  dehydrogenases;  this  may  be  related  to  the  possible  location 
of  binding  sites  for  the  coenzymes  on  adjacent  helices  of  the  apoenzyme, 
separation  of  these  sites  occurring  unless  they  are  held  together  by  the 
coenzymes.  Nicotinamide  also  inhibits  6-phosphogluconate  dehydrogenase 
competitively  (Dickens  and  Glock,  1951).  The  NADH  oxidases  from  pigeon 
liver  microsomes  and  mitochondria  are  inhibited  40%  and  23%  by  20  mM 
and  80%  and  78%  by  200  mM  nicotinamide,  respectively  (Jacobson  and 
Kaplan,  1957  a). 

A  unique  nicotinamide  derivative,  A'-piperidinomethylnicotinamide, 
which  is  claimed  to  be  a  specific  dehydrogenase  inhibitor  has  been  reported 
by  Matkovics  et  al.  (1961).  Inhibition  of  methylene  blue  reduction  in  liver 
homogenates  in  the  presence  of  various  substrates  was  studied.  Unfortu- 
nately the  inhibitor  concentrations  are  not  given  (only  the  milligrams  added) 


CONH-CH.— N 


N-Piperidinomethylnicotinamide 

and  no  control  experiments  on  endogenous  activity  are  included.  Inhibition 
of  the  oxidation  of  glucose,  malate,  lactate,  and  glutamate  was  observed. 
However,  succinate  oxidation  is  also  inhibited,  indicating  that  this  sub- 
stance is  not  specific  for  the  NAD(P)  dehydrogenases.  Furthermore,  no 
evidence  for  competition  with  the  coenzymes  was  provided.  Much  more 
work  must  be  done  before  this  inhibitor  can  be  accepted  as  having  specific 
anticoenzyme  activity. 


504  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Pyridine-3-sulfonate  and   Pyridine-3-sulfonamide 

These  analogs  might  be  expected  to  inhibit  nicotinate  and  nicotinamide 
metaboHsm.  Both  inhibit  the  growth  of  various  bacteria  and  the  inhibitions 
can  be  overcome  by  nicotinamide  (Mcllwain,  1940).  Reference  to  the  tab- 
ulation on  page  501  shows  that  lactate  dehydrogenase  is  inhibited  more 
by  these  analogs  than  the  corresponding  nicotinic  compounds,  whereas  glu- 
cose dehydrogenase  behaves  in  the  opposite  fashion  (von  Euler,  1942). 
Feeding  pyridine-3-sulfonate  at  5%  in  the  diet  to  mice  produces  no 
signs  of  nicotinate  deficiency,  but  the  mouse  does  not  require  exogenous 
nicotinate  (Woolley  and  White,  1943  a).  However,  nicotinate-deficient  dogs 
are  made  worse  by  administration  of  the  analog  (Woolley  et  al.,  1938), 
although  Gaebler  and  Beher  (1951)  observed  no  effect  of  0.5-2  g/day  of 
pyridine-3-sulfonate  on  the  excretion  of  A^-methylnicotinamide,  erythrocyte 
coenzyme  level,  or  general  health  of  either  normal  or  nicotinate-deficient 
dogs.  Hicks  (1955)  found  hippocampal  necrosis  in  only  one  animal  given 
pyridine-3-sulfonate,  so  that  it  is  presumably  not  as  effective  as  3-acetyl- 
pyridine.  Brain  NADase  is  not  inhibited  by  pyridine-3-sulfonamide  (Mcll- 
wain, 1950),  and  the  sulfonate  does  not  significantly  inhibit  either  beef 
spleen  NADase  (Zatman  et  al.,  1954  a)  or  nicotinamide  deaminase  (Grosso- 
wicz  and  Halpern,  1956  b).  The  most  potently  inhibited  enzyme  examined 
seems  to  be  yeast  alcohol  dehydrogenase,  the  sulfonate  being  the  most  po- 
tent inhibitor  of  all  the  substituted  pyridines  (Table  2-30).  There  is  no 
evidence  that  NAD  analogs  can  be  formed  from  these  substances.  The 
respiration  of  resting  Mycobacterium  phlei,  either  without  substrate  or  with 
glycerol,  is  not  inhibited  by  1  mM  pyridine-3-sulfonate,  but  proliferating 
bacterial  respiration  is  inhibited  52-85%  (almost  completely  by  10  niM) 
(Miiller  et  al.,  1960).  These  scattered  observations  do  not  arouse  much  in- 
terest in  these  analogs,  but  perhaps  the  proper  systems  have  not  been 
studied. 

6-Aminonicotinamide 

This  analog  has  been  called  the  most  potent  nicotinamide  antagonist 
available  (Johnson  and  McCoU,  1955).  The  acute  LD50  in  mice  is  35  mg/kg, 
although  2  mg/kg/day  leads  to  50%  mortality  by  the  eleventh  day  (John- 
son and  McColl,  1956).  Simultaneous  administration  of  50  mg/kg  nicotina- 
mide raises  the  LD50  8-fold.  It  is  very  toxic  to  rabbits,  producing  loss  of 
motor  control  and  paralysis,  and  in  rats  it  produces  these  and  other  signs 
of  nicotinate  deficiency  (Halliday  et  al.,  1957).  The  endogenous  respiration 
of  liver  homogenates  from  treated  mice  is  depressed  70%  and  lactate  oxi- 
dation is  depressed  49%;  addition  of  NAD  counteracts  these  depressions. 
However,  no  effect  is  observed  when  the  analog  is  added  directly  to  liver 
slices.  The  possibility  of  the  formation  of  a  NAD  analog  was  entertained 
and  such  an  analog  was  soon  isolated  following  incubation  of  NAD,  6-ami- 


ANALOGS   OF   NICOTINAMIDE  505 

nonicotinamide,  and  NADase  (Johnson  and  McCoU,  1956).  The  NAD  ana- 
log was  also  detected  spectroscopically  in  the  livers  and  kidneys  of  treated 
mice.  This  NAD  analog  is  completely  inactive  with  yeast  alcohol  dehydro- 
genase. If  the  analyses  of  Shapiro  et  at.  (1957)  actually  represent  true  NAD, 
the  small  decreases  observed  following  administration  of  6-aminonicotina- 
mide  at  30  mg/kg  for  3  days  (14%  in  liver,  17%  in  adenocarcinoma,  and 
none  in  brain)  would  point  to  the  NAD  analog  as  being  inhibitory  to  dehy- 
drogenases. This  explanation  was  accepted  by  Friedland  et  al.  (1958)  on 
the  basis  of  decreases  in  tissue  ADP  and  ATP,  as  well  as  oxidative  inhi- 
bition. Although  it  has  generally  been  assumed  that  the  central  effects  of 
6-aminonicotinamide  are  due  to  the  formation  of  an  abnormal  NAD(P)  ana- 
log and  to  reduction  in  normal  NAD(P),  Redetzki  and  Alvarez- O'Bourke 
(1962)  found  that  the  NAD  level  in  the  brain  is  only  slightly  depressed, 
despite  the  rather  marked  decrease  of  liver  NAD,  and  obtained  no  evidence 
for  the  occurrence  of  an  abnormal  analog.  The  6-aminonicotinamide  analog 
of  NAD  inhibits  creatine  kinase  and  pyruvate  kinase  noncompetitively, 
about  40%  depression  occurring  at  1  mM  (von  Bruchhausen  1964).  It  is 
unlikely  that  these  actions  can  be  important  in  vivo  unless  these  enzymes 
are  much  more  sensitive  in  intact  cells. 

Administration  of  the  analog  to  adenocarcinoma-bearing  mice  leads  to 
inhibition  of  certain  enzymes  determined  in  homogenates  of  the  tumor: 
lactate  dehydrogenase  is  not  affected,  glyceraldehyde-3-P  dehydrogenase  is 
inhibited  44%,  the  conversion  of  /?-hydroxybut>Tate  to  acetoacetate  is  in- 
hibited 69%,  and  a-ketoglutarate  oxidase  is  inhibited  83%  (Dietrich  et  at., 
1958).  It  was  believed  that  the  NAD  analog  is  quite  tightly  bound  to  the 
apoenzymes  and  prevents  the  combination  with  NAD. 

6-Aminonicotinamide  exerts  a  depressing  action  against  the  growth  of 
certain  lymphosarcomas  and  adenocarcinomas,  and  this  is  reversed  by  nic- 
otinamide (Halliday  et  al.,  1957).  Tumor  regression  occurs  at  3-4  mg/kg/ 
day  but  some  weight  loss  also  occurs;  at  lower  doses  the  weight  loss  can 
be  minimized  with  some  reduction  in  carcinostatic  activity,  but  combined 
at  these  lower  doses  with  8-azaguanine  it  is  reasonably  effective  (Shapiro 
et  al.,  1957).  It  was  considered  to  represent  a  new  class  of  potentially  useful 
carcinostatic  agents.  It  is  interesting  that  6-aminonicotinate  is  one-seventh 
to  one-fifteenth  as  toxic  as  the  amide,  suggesting  either  that  penetration 
of  the  acid  is  Kmiting  or  that  conversion  to  the  amide  is  slow. 

Inhibition   of  NAD(P)   Enzymes   by  Various  Nucleotides 
and   Related   Substances 

The  study  of  these  inhibitions  has  three  major  purposes:  (1)  to  obtain 
information  on  the  nature  of  the  active  centers  and  the  binding  groups  of 
the  coenzymes,  (2)  to  understand  better  the  mutual  relationships  between 
these  naturally  occurring  substances  and  the  possible  regulatory  effects 


506  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

exerted  in  cellular  metabolism,  and  (3)  to  find  useful  inhibitors  that  may 
specifically  inhibit  particular  reactions  in  complex  systems.  Some  of  the 
results  on  different  types  of  enzyme  involving  NAD  or  NADP  are  summa- 
rized in  Table  2-31.  It  is  unfortunate  that  in  very  few  instances  have  the 
types  of  inhibition  been  determined  and  it  is  seldom  possible  to  calculate 
accurately  the  K/s  or  even  relative  K/s,  from  which  interesting  binding 
energy  information  might  be  obtained. 

One  may  first  ask:  Does  the  inhibitory  activity  generally  increase  as  ri- 
bose  and  phosphate  groups  are  added?  The  answer  is  roughly  in  the  ajBfir- 
mative  for  NAD  kinase,  NADH  pyrophosphatase,  NADH  oxidase,  NAD  : 
NADP  transhydrogenase,  alcohol  dehydrogenase,  and  malate  dehydroge- 
nase, but  in  a  few  enzymes  there  appears  to  be  no  definite  trend,  while  in 
some  the  addition  of  a  group  may  reduce  the  binding.  The  addition  of  a 
phosphate  to  adenosine  to  form  5'-adenylate  (AR  — »■  5'-ARP)  leads  to  only 
0.1  kcal/mole  extra  binding  to  the  alcohol  dehydrogenase  and  0.4  kcal/mole 
to  the  NAD  kinase,  but  an  increased  binding  of  over  2.4  kcal/mole  for  the 
NADH  pyrophosphatase.  The  further  addition  of  a  phosphate  to  form  ADP 
increases  the  binding  approximately  0.7  kcal/mole  for  the  NAD  kinase, 
1.1  kcal/mole  for  alcohol  dehydrogenase,  and  0.6  kcal/mole  for  the  NAD  : 
NADP  transhydrogenase,  whereas  the  binding  to  NADH  oxidase  or  NADH 
pyrophosphatase  is  unchanged  or  slightly  reduced.  Addition  of  another  phos- 
phate to  form  ATP  leads  to  increased  binding  only  for  the  liver  NADH 
oxidase.  Addition  of  a  ribose  to  ADP  to  form  ARPPR  has  no  effect  for 
NAD  kinase  but  increases  the  binding  around  0.7  kcal/mole  with  NADH 
pyrophosphatase.  Final  addition  of  nicotinamide  to  ARPPR  to  form  NAD 
increases  the  binding  around  1.9  kcal/mole  for  NAD  kinase  and  NADPH- 
glutathione  reductase,  whereas  a  reduction  of  1.9  kcal/mole  in  the  binding 
to  NADH  pyrophosphatase  is  observed.  Addition  of  nicotinamide  to  2'-P- 
ARPPR  to  form  NADP  leads  to  a  2.3  kcal/mole  increase  in  binding  for  the 
NADPH-glutathione  reductase  and  to  very  little  change  for  the  NADP- 
cytochrome  c  reductase.  The  marked  variation  in  behavior  between  en- 
zymes and  the  uncertainty  in  the  accuracy  of  the  energy  values  make  it  im- 
possible to  draw  definite  conclusions  or  formulate  rules  for  these  inhibitions. 
It  appears  that  all  the  components  of  the  NAD  and  NADP  molecules  can 
participate  in  the  binding,  although  not  all  of  them  need  function  for  a 
particular  enzyme.  The  rather  marked  inhibition  occasionally  exerted  by 
NADH  on  NAD  reactions,  or  by  NAD(P)  on  NAD(P)H  reactions,  indicates 
not  only  the  specificity  of  these  enzymes  but  points  to  a  somewhat  different 
orientation  of  the  oxidized  and  reduced  forms  on  the  enzymes. 

A  more  interesting  correlation  emerges  when  one  considers  the  variation 
of  inhibitory  potency  with  the  position  of  phosphate  groups  on  the  adenyl 
ribose.  In  NAD  the  5-position  is  phosphorylated  and  enzymes  involving 
NAD  are  more  readily  inhibited  by  5'-AMP  than  by  2'-  or  3'-AMP  (NAD 


ANALOGS   OF   NICOTINAMIDE 


507 


kinase  and  malate  dehydrogenase).  However,  NADP  is  additionaUy  phos- 
phorylated  in  the  2-position  and,  as  has  been  especially  emphasized  by 
Neufeld  et  al.  (1955),  enzymes  reacting  with  NADP  are  frequently  inhibited 
by  2'-AMP.  In  addition  to  the  enzymes  listed  in  Table  2-31  (NADPH  dia- 
phorase,  NADP-cytochrome  c  reductase,  glucose-6-P  dehydrogenase,  and 
isocitrate  dehydrogenase),  they  found  phosphogluconate  dehydrogenase  and 
NADP-activated  oxalacetate  decarboxylase  to  be  inhibited  by  2'-AMP  more 
than  by  the  other  AMP's.  It  was  suggested  that  2'-AMP  may  be  a  useful 
inhibitor  to  distinguish  between  NAD  and  NADP  enzymes.  It  must  be 
admitted  that  an  insufficient  number  of  NAD  enzymes  have  been  examined. 
It  is  interesting  that  the  NAD  :  NADP  transhydrogenase  is  inhibited  more 
potently  by  2'-AMP  than  by  3'-  and  5'-AMP. 


100 


Fig.  2-17.   Inhibitions  of  glutamate  semialdehyde  reductase  by  various  nucleosides 
and  nucleotides.  (From  Smith  and  Greenberg,  1957.) 


The  importance  of  phosphate  groups  for  the  binding  of  inhibitors  of  this 
type  is  seen  strikingly  in  the  study  on  glutamic  semialdehyde  reductase  by 
Smith  and  Greenberg  (1957)  (Fig.  2-17).  The  inhibitions  by  AMP,  ADP, 
and  ATP  are  competitive,  but  NADP  inhibits  noncompetitively.  Although 
the  addition  of  the  first  phosphate  to  form  AMP  has  little  effect  (not  more 
than  0.1  kcal/mole),  the  addition  of  each  of  the  next  two  phosphates  to 


508 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


03 


o 

o 

ij 

< 

Is 

< 

« 

Ph 

Q 

t»> 

a 

<: 

H 

IB 

Q 
H 

^i 

O 

"— ' 

H 

iJ 
O 

c 

t3 

_o 

^ 

'-13 

o°  -2 

;£ 

o 

KH 

s 

<! 

> 

^5 

>H 

"  s 

M 

M 

en 

ft 

<: 

A 

;?; 

t-l 

o 

-IJ 

P5 

O 

'2 

Q 

'a 

-< 

'"' 

^ 

o 

g 

> 

IJ 

o 

o 

^ 

% 

cS 

1— 1 

GO 

> 

N 

Is 

w 

fe 

o 

is 

4) 

o 

O 

M 

tH 

H 

M 

O 

S 

;s 

(35 


e6 


> 


o 


CO 


e  5 
o 


o  o 


Oi      C<l     -*     <35 

o    o  >— I  fJ 


O 


o  o  o    o    o  o  o 


bo 
o 


fc  CO  CO  "^  CO 


<1 

tf 

PM 

CM 

05 

;2; 

(Xl 

fM 

-,    fM   Oh 

Om 

Oh 

Oh 

P5 

^   ^   ^ 
S   1=   ."^ 

P5 

03 

03 

:?  P5 

*? 

< 

<t1 

lO    <1 

w 

<1    lO    lO 

c^ 

fO 

to 

P    Q 


P  P 


(N    O 
C<1    o 

o  o 


o 


CO    o 

^H      l-H 

d  d 


■*    O    CO    o 


Oh    Oh 
03  03 


_  ^ 

12;  03 

Oh 

03  Oh 

03 

Oh  Oh 

03 

Pm  Oh 
03  03 

lO 

-^ 

<  < 

fO 


m 


P5 


e8 

eS 

(5 

^ 

Ph    01 

-       «> 

•      M 

to     bc 

CO      o 

s-S 

o     >> 

eS     >*> 

o  j:: 

-P    Ji 

s    <» 

S     0) 

o-^ 

o-^ 

ANALOGS   or   NICOTINAMIDE  509 


r-  »  (-1 

3  ^  O 


CO    (M 

d  d 


eu 

CLh 

« 

K 

<< 

<: 

3 


o  o  o  o 


Pm 

CLi 

CM 

PLH 

»5 

Pm 

(1^ 

05 

P5  X 

<r: 

<  <  < 

■<!     5^1 

P5P5     ^  ^^  ^   X  A  X  ^%   c^  ^;  ^  ^  ^  ^ 


PM        £  Pm 


S  MM  ig  M 

PQ  PL)  PM  Pq  CLi 


2-   00 

Q 

o  p: 

<1 

<^  H 

^ 

;?;  < 

510 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


cr; 


Ui 


a 

a 

? 

ce 

^ 

c 

,^ 

]S 

03 

^ 

< 

s 

P3 

& 


O  O  Q  O  O 


O 


O 


^ 


r^ 


O    lO    CD    O 

O    O    O 

O    <M    LO    >C 

»C    lO    lO 

^^ 


CO     CO     GO     O     rt     -* 


c~  t^  t-  i> 


ic     t^  cc  CO   CO  CO 

O— 1^       -HCOCOCOCO    — 


Pm  Pli  PL| 
^  (l^  P^ 
Pi    Pi    tf 


^     ^     ^     ^     P^     S 

1^  1^  1^  pS  S  *? 

5^    CO    to    •<    ■<      S 


fe  W  <1 

Ph  Ph  Pi 

P-i  Ph  P-i 

(In   Pli  P-i  Pli 

Pi  Pi  Pi  Pi 

fEH     <lj  <1  <tj 


pli  5-1 

S  S"  &^ 

Zi  p^  p^ 

1^  Pi  Pi 

io  <5  <! 


m 


M 

£ 


P-l      O 

12;  ^ 


w  O 


0) 

s 

5 

TJ 

O) 

W 

O 

Q 

'« 

<5l 

>< 

^ 

ANALOGS   OF   NICOTINAMIDE  511 


t3 


<»  'St  }2 


2         -^  S 

W  ^  S  H?  -  ^ 


o  o 


w,     ^ 

D5  ^  05 

2i 

Oh   M   CL, 

CLi 

Ph    CL|    fL| 

Ph 

P^   Ph   P^ 

oi 

Pi  Pi  P5 

<1 

<  <  < 

Pj         gPHPHP^><-flHPH£PH  gCMCMfl^PM      2SSi=^f^i=^'^  IxIS^ 

e^       p^a^a^pHS^p^Q^p^         ^pucuPmPm     ^^^phP^plhP^  £S 

>^       "^pitfp^^Spigpi     p5-?pipipipi     l^^^piPipiPi  ^>^ 


2.       »  ::i  M 


^    2    :^  § 

«   "f.    .5P  .3' 

*^  pH  (l] 


°  ft  J 

Q  op 


a 

■8 

3 

o 

Ti 

K 

X 

01 

Q-i 

o 

o 

Q 

s 

t3 

*i1 

>> 

X 

^ 

o 

o 

512 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


<A 


a 
>> 
H 


!^ 


s 


3 
O 

m 


S 


CLi  Ah  CL| 

P5  fi2  Ci^ 

*?  *?  ^ 

c<i  eo  ic 


.     flH 


tf    Oh    05    »5 
«55    <1    <tl    <| 


|Tl 

jS 

£ 

(1h 

o 

O 

-1^ 

3 

TJ 

*f 

X 

^ 

M 

o 

0 

^ 

CO 

f3 

OS 

i-H 

ri<! 

-»J 

< 

^ 

=3 

'^ 

a 

c6 

e8 

W 

is  — 

13 

'C    1— ( 

cS 

C3    O 

^ 

w 

o 

hA 

1-5 

O    ^    O    IC    (M    OJ    t^ 


xxxxxxxxx 


Ph 
Ph 


fel 


•s  § 

SI 


12^  ^  !?;  o 


O    O    Pi    o 


5  05  i 

Oh  Pm  Oh 

Ph  Oh  Ph 

PLi  Ph  Oh 

Oh  Oh  Oh 


I  1 

O     cS 


rS  9  C 


102  o 


a  s 


c 
'c 

=*  Ph  o  St3 
II     /S-g  £ 
o  « 


(D    <3j 


.S  .S  o  ."S 


cS  •-<    <D 


*3  ^^  /-*    ^ 


?3  P^ 


<]    O    O) 


S  2  s  «  1^ 

K^    tc    c«    cS    "    5 

2  =*H  .£  ■>,  2  "a 

t^    -+:i  "-H  ~  '-^  *iH 
0)     o    C    eS    r    PH 


<a  ia 


JZ;     ^  -S  .2  -^  '^   -^ 


(U 


CO   -S 


O      „T3  Ph 

?:  oj  cs  o  c 


>     CO    g  <tj 


&-" 


ce 


o 


tUD    O    O 


on    cu    i 


-  O 

2   C 


■^     >    e3    0)    C    C     <u 
cS     P    S  1^    5?  •-  T3 


^^^ 


c6  .2 


cS 


2  •-  ^    oj 

„^^;^- >> 

.3    C  9^  "^ 
lU    2    >  i-H  Ph    S    C 

S  ^  «  :3  Oh        -S 

|£  II  §x-2a 

>^     3      t4       tH 

O    cS    O    eS 


ANALOGS   OF   NICOTINAMIDE  513 

form  ADP  and  ATP  leads  to  about  1.2  kcal/mole  increase  in  binding  energy. 
The  addition  of  the  2'-phosphate  to  form  NADP  from  NAD  increases  the 
binding  markedly  and  changes  the  nature  of  the  inhibition.  It  is  rather 
strange  that  the  addition  of  nicotinamide  riboside  to  ADP  lowers  the  bind- 
ing energy  about  1  kcal/mole. 

Williams  (1952)  found  that  malate  dehydrogenase  is  inhibited  by  adenine, 
adenosine,  and  ATP.  From  this  observation  he  concluded  that  such  nor- 
mally occurring  substances  may  well  affect  dehydrogenases  and  other  en- 
zymes in  the  cell.  His  work  stemmed  from  the  report  of  Raska  (1946)  that 
administration  of  300-500  mg/day  adenine  to  dogs  on  normal  diets  leads 
to  the  development  of  multiple  avitaminosis  after  10-20  days;  signs  of  ni- 
cotinate  deficiency,  such  as  black  tongue,  were  noted.  The  many  more  data 
now  available  serve  to  strengthen  Williams'  conclusion,  since  even  more 
potent  inhibitors  have  been  reported.  There  has  been  much  speculation 
concerning  the  regulation  of  oxidative  reactions  by  adenine  nucleotides  me- 
diated through  coupled  phosphorylation.  It  is  quite  possible  that  other  more 
direct  effects  on  dehydrogenases  occur,  both  in  the  cell  (particularly  in  the 
compartmentalized  mitochondria)  and  in  experimental  enzyme  prepara- 
tions where  the  concentrations  of  added  nucleotides  are  often  high  enough 
to  inhibit  appreciably.  From  Table  2-31  we  see  that  five  enzymes  are  inhi- 
bited from  22%  to  50%  by  ATP  at  concentrations  from  1  mM  to  3.4  mM; 
ATP  is  commonly  added  to  mitochondrial  preparations  at  these  or  higher 
concentrations.  An  experimental  survey  of  dehydrogenase  inhibitions  by 
nucleotides  would  be  valuable.  Chen  and  Plant  (1963),  on  the  basis  of  the 
fairly  potent  inhibitions  exerted  by  certain  nucleotides  on  the  NAD-link- 
ed  isocitrate  dehydrogenase  (Table  2-31),  felt  that  some  regidation  of  cycle 
activity  may  be  exerted,  and  if  such  does  occur  it  would  be  a  very  impor- 
tant factor  in  understanding  not  only  the  effects  of  nucleotide  analogs  but 
also  of  many  inhibitors  which  either  primarily  or  secondarily  alter  the  levels 
of  cellular  or  mitochondrial  nucleotides.  Another  interesting  point  has  been 
brought  out  by  Dalziel  (1962)  in  connection  with  possible  impurities  in 
preparations  of  the  coenzymes.  Although  one  might  expect  in  many  cases 
an  insignificant  inhibitory  effect  of  certain  analogs  because  they  have  been 
shown  to  bind  less  tightly  than  the  normal  coenzyme  to  the  apoenzyme, 
Dalziel  correctly  states  that  it  is  the  relative  values  of  K^  and  K^^  which 
are  important,  and  K,,^  can  be  much  higher  than  K^..  He  calculated  that 
the  presence  of  an  analog  of  NADH  as  a  3%  impurity  can  produce  as  much 
as  70%  inhibition  of  liver  alcohol  dehydrogenase  if  the  analog  and  NADH 
have  the  same  affinity  for  the  apoenzyme. 

An  NAD(P)  analog  which  would  bind  to  the  NAD(P)  site  on  dehydroge- 
nases and  then  react  chemically  with  some  group  at  the  site  might  well 
be  of  some  value  in  labeling  these  sites.  Such  an  analog  was  investigated 
by  van  Eys  et  al.  (1962)  on  the  basis  that  thiazole  rings  often  open  at  al- 


514  2.  ANALOGS  OF  ENZYME  KEACTION  COMPONENTS 

kaline  pH  to  generate  a  free  SH  group.  The  NAD  analog  with  4-methyl-5- 
(/?- hydroxy  ethyl  )thiazole  replacing  nicotinamide  was  found  to  behave  in 
this  manner  and  to  form  a  disulfide  bond  with  SH  groups  at  the  site  of 
dehydrogenases,  this  binding  being  competitive  with  NAD.  In  the  case  of 
horse  liver  alcohol  dehydrogenase,  2  moles  of  this  analog  are  bound  tightly 
to  each  mole  of  enzyme. 

One  of  the  most  interesting  studies  of  dehydrogenase  inhibition  by  nu- 
cleotides is  that  of  the  complex  effects  of  GTP  on  glutamate  dehydrogenase 
(Frieden,  1962,  1963).  The  K^  is  0.0003-0.0005  mM  and  the  kinetics  being 
uncompetitive  point  to  different  sites  for  NADP  and  GTP.  Furthermore, 
GTP  not  only  inhibits  directly  but  increases  the  ability  of  NADH  to  inhibit. 
Since  the  NADH  inhibition  is  due  to  the  dissociation  of  the  enzyme  into 
four  subunits,  it  is  likely  that  GTP  enhances  the  process,  and  this  was  dem- 
onstrated ultracentrifugaUy.  The  dissociation  of  the  tetramer  enzyme  it- 
self is  not  necessarily  the  basic  cause  of  the  loss  of  activity;  it  is  possible 
that  structural  changes  brought  about  by  NADH  and  GTP  produce  both 
dissociation  and  reduced  catalytic  activity.  The  behavior  can  be  explained 
adequately  on  the  basis  of  three  binding  sites:  (1)  a  coenzyme  site,  (2)  a 
purine  nucleotide  site  with  which  GTP  and  activating  nucleotides  react, 
and  (3)  a  NADH-binding  site.  The  following  complexes  are  thus  possible 
—  EC,  ECI,  ECg,  ECgl,  and  EI  —  where  C  represents  the  coenzyme.  The 
binding  of  GTP  to  the  enzyme  depends  on  the  presence  of  NADP  at  a 
vicinal  site,  the  EI  complex  probably  not  being  of  much  importance.  The 
importance  of  this  situation  for  the  regulation  of  cell  metabolism  is  obvious, 
particularly  since  this  enzyme  plays  a  central  role  in  many  pathways.  Frie- 
den pointed  out  the  likely  relationship  between  glutamate  dehydrogenase 
and  the  or-ketoglutarate  step  in  the  cycle;  GDP  is  required  for  the  conver- 
sion of  succinyl-CoA  to  succinate  and  GTP  is  formed,  which  can  suppress 
the  activity  of  glutamate  dehydrogenase,  an  enzyme  which  under  certain 
conditions  controls  the  steady-state  level  of  a-ketoglutarate  in  the  cycle. 
He  also  suggests  that  ammonia  formation  by  the  liver,  protein  synthesis, 
and  glyconeogenesis  can  all  be  regulated  by  this  inhibition  involving  a  feed- 
back site. 

ANALOGS  OF  THIAMINE 

Thiamine  functions  in  metabolism  in  the  pyrophosphorylated  form  as  the 
coenzyme  in  various  reactions  where  a  bond  adjacent  to  a  carbonyl  group  is 
broken  (a-cleavage),  the  active  complex  in  each  case  being  an  aldehyde- 
thiamine-PP-enzyme  structure  wherein  a  C — C  bond  is  formed  at  the  2- 
position  of  the  thiazole  ring.  These  reactions  would  include  (1)  a-keto  acid 
decarboxylation  (e.g.  pyruvate  decarboxylase),  (2)  a-keto  acid  oxidation 
(e.g.  pyruvate  and  a-ketoglutarate  oxidases),  (3)  the  phosphoroclastic  reac- 
tion of  pyruvate,  and  (4)  a-ketol  formation  (e.g.  transketolase  and  phos- 


ANALOGS   OF   THIAMINE 


515 


phoketolase).  Thus  three  major  metabolic  sequences  —  the  pentose-P  path- 
way, the  tricarboxylate  cycle,  and  photosynthetic  carbon  dioxide  fixation  — 
are  dependent  on  thiamine-PP,  since  a-cleavage  occurs  in  all,  and  a  variety 
of  other  metabolic  processes  can  be  secondarily  affected.  Thiamine  defi- 
ciency, or  interference  with  the  formation  or  function  of  thiamine-PP,  can 
produce  profound  metabolic  and  physiological  disturbances.  Animals  re- 
quire preformed  thiamine,  most  plants  can  synthesize  the  entire  thiamine 
molecule,  and  microorganisms  vary  widely  from  complete  dependence  on 
exogenous  supply  to  complete  synthetic  ability.  The  responses  of  organisms 
to  thiamine  analogs  wiU  depend  on  these  factors  as  well  as  the  role  of  thia- 
mine in  metabolism.  The  pathways  of  thiamine  biosynthesis  are  not  com- 
pletely understood  and  the  accompanying  scheme  is  to  be  taken  as  pro- 
visional and  not  necessarily  applicable  to  all  organisms.  Thiaminase  is  ap- 
parently absent  or  relatively  inactive  in  most  tissues  and  thus  the  reactions 
catalyzed  by  this  enzyme  are  probably  not  common  or  important.  The  most 
important  reaction  is  the  pyrophosphorylation  of  thiamine  since  certain 
analogs  can  interfere  here  or  be  similarly  phosphorylated.  Thiamine-PPP 
has  been  included  because  its  formation  from  thiamine  in  yeast  has  been 
demonstrated  (Kiessling,  1956),  although  it  is  coenzymically  inactive.  It 
may  be  noted  that  ATP  is  required  for  thiamine-PP  synthesis  and  that 

"Pyrimidine"  "Thiazole" 


+  PP 
'pyrimidine -PP" 


+  P 
"thiazole- P' 


"pyrimidine-B" 
+  : 

"thiazole" 

"pyrimidine-B" 

+ 
"thiazole-PP" 


thiamine- P 


+  B 


Thiamine 


+  H,0 


thiamine-PP 


+  H,0 


"pyrimidine" 
:  + 

"thiazole" 

"pyrimidine" 

:  + 

"thiazole-PP' 


thiamine- PPP 


enzyme-thiamine-  PP 


(The  pyrimidine  portion  of  thiamine  is  indicated  in  quotes  and  is  2- 
methyl-4-amino-5-hydroxymethylpyrimidine;  the  thiazole  portion  is 
designated  likewise  and  is  4-methyl-5-{2-hydroxyethyl)thiazole.  B  is 
any  base  that  can  replace  the  thiazole  in  the  exchange  reaction  catalyzed 
by  thiaminase. ) 


516  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

thiamine-PP  is  involved  in  metabolic  reactions  leading  to  ATP,  so  that 
interference  with  thiamine-PP  formation  or  function  will  tend  to  deplete 
the  cells  of  ATP  and  perhaps  further  depress  thiamine-PP  synthesis. 

The  possible  sites  of  action  for  thiamine  analogs  can  be  broadly  classified 
as  (1)  inhibition  of  thiamine-PP  synthesis,  either  on  the  formation  of  thia- 
mine or  its  pyrophosphorylation,  (2)  interference  with  the  formation  of 
complexes  between  thiamine-PP  and  enzymes,  and  (3)  inhibition  of  thia- 
minase.  In  any  case  the  inhibition  may  be  exerted  by  either  the  analog  or 
its  phosphorylated  derivatives.  There  is  no  evidence  that  any  significant 
effects  of  any  of  the  analogs  studied  can  be  attributed  to  thiaminase  inhi- 
bition, so  the  first  two  mechanisms  are  undoubtedly  the  most  important 
in  the  induction  of  thiamine  deficiency  symptoms.  There  is  some  evidence, 
which  will  be  discussed  later,  that  thiamine  may  have  a  function  or  func- 
tions unassociated  with  coenzyme  activity,  particularly  in  the  nervous  sys- 
tem, and,  if  this  is  true,  one  might  consider  the  interference  by  analogs 
in  this  function. 

It  would  appear  that  most  of  the  groups  in  the  thiamine  molecule  parti- 
cipate in  either  the  binding  or  the  catalysis  inasmuch  as  the  structure  can 
not  be  significantly  altered  without  loss  of  activity,  and  the  number  of  ef- 
fective analogs  is  rather  small.  The  first  report  of  enzyme  inhibition  by  a 
thiamine  analog  was  by  Buchman  et  al.  (1940),  who  found  yeast  pyruvate 
decarboxylase  activity  to  be  depressed  by  4-methyl-5-hydroxyethylthiazole 
diphosphate  (which  they  called  "thiazole  pyrophosphate"),  the  phosphory- 
lated thiazole  portion  of  thiamine.  Neither  the  nonphosphorylated  com- 
pound nor  the  monophosphate  is  inhibitory.  It  requires  about  10  times  as 
much  analog  as  thiamine-PP  to  inhibit  50%,  but  if  the  analog  is  added 
before  the  thiamine-PP,  the  inhibition  is  more  pronounced.  These  results 
point  to  the  importance  of  the  phosphate  groups  in  the  binding.  They 
state,  "We  conclude  that  there  has  been  demonstrated  here  a  not  hitherto 
recognized  type  of  competitive  inhibition  of  enzyme  reactions,  caused  by 
competition  not  between  substrate  and  inhibitor  but  between  coenzyme 
and  inhibitor."  (See  formulas  on  page  517). 

Either  the  pyrimidine  portion  or  the  thiazole  portion  of  the  thiamine 
molecule  can  be  altered  to  form  analogs.  Replacement  of  the  thiazole  ring 
with  a  similarly  substituted  pyridine  ring  gives  pyrithiamine,  which  was 
shown  by  Robbins  (1941)  to  inhibit  the  growth  of  certain  fungi,  and  by 
Woolley  and  White  (1943  b)  to  produce  thiamine  deficiency  symptoms  in 
mice.  Replacement  of  the  pyrimidine  amino  group  with  a  hydroxyl  group 
leads  to  oxy thiamine,  found  by  Bergel  and  Todd  (1937)  to  lack  vitamin 
activity,  and  by  Soodak  and  Cerecedo  (1944)  to  be  quite  toxic  to  mice. 
These  two  analogs  have  been  studied  the  most  thoroughly  of  the  thiamine- 
like  compounds  and  remain  the  most  frequently  used  to  produce  experi- 


ANALOGS   OF  THIAMINE 


517 


X 

^^ 

1 

X    1 

o^ 

o 

o^ 

a 

a? 

a? 

u 

o 

4. 

sf 

K 

1 

X 

^^■^^ 

^u 

^O 

OT 

^u 

QJ 

c 

.3 

u 

7^ 

^  /- 

+2: 

"u 

c 

s 

+z 

^U 

S 

~"^; 

^u^ 

1 

K 

E 

2 

1 

s" 

ci 

K 

u 

5 

u 

a 
o 

X 

u 

5 

§ 
o 

^ 

U 

ll^ 

0^ 

n^ 

1 

^^ 

1  1                 ■ 

z^z 

■z 

r^ 

o 

o 

u 

af 

sT 

1 

X 

1 

s 
o 

•      I 

0-P-=0 

I 
o 

.     I 


4. 

as* 


k^ 


+  Z 


J3  . 


a? 
o 

X 

■^    /^ 

a? 
o 

1 
u 

aT 

rn                  -^ 

af 

8 

o 

1 

o 

L 

w 
o 

X 

a? 
a. 

1 

a? 

CO ( 

a 

c 
S 

0) 

c 

^ 

X 

X 

0) 

c 

+  2 

af 

X 

.2 

+z 

aT 

+z 

4, 

X 

o 

rt 

u 

"i^ 

p 

U 

>. 

u 

§5 

aT       1 

1 

o 
(1) 

^ 

X 

a! 

< 

aT 
z 

Jx 

4^-H 

S  o 

¥^ 

o 

¥^ 

6 

¥Y 

E 

Z"«>Z 

1  1                   1 

z^  ^z 

o 

Z^  /;2 

2^^::^'' 

\3^ 

Y 

Y 

T 

o 

o 

u 

o 

af 

af 

af 

af 

518  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

mental  disturbances  in  thiamine  function.*  Two  other  types  of  analog  per- 
haps deserve  more  attention  than  they  have  received:  the  imidazole  analog 
(the  thiazole  ring  replaced  with  a  similarly  substituted  imidazole  ring)  exerts 
an  antivitamin  effect  on  bacterial  growth  (Erlenmeyer  et  al.,  1948),  and  the 
2' -n-butylpyrimidine  analog  produces  thiamine-deficiency  states  in  rats  (Em- 
erson and  South  wick,  1945),  both  substances  being  roughly  of  the  same 
potency  as  pyrithiamine  and  oxythiamine.  A  large  number  of  interesting 
analogs,  such  as  those  synthesized  by  Livermore  and  Sealock  (1947)  and 
by  Mano  and  Tanaka  (1960),  have  not  yet  been  adequately  examined. 

Effects  on   Enzymes  Dependent  on  Thiamine-PP 

It  is  clear  that  neither  oxythiamine  nor  pyrithiamine  is  inhibitory  to 
pyruvate  decarboxylase,  but  that  the  diphosphate  esters  can  interfere  with 
the  binding  of  thiamine-PP  to  the  apoenzyme.  Thus  the  yeast  decarboxylase 
is  inhibited  by  oxy thiamine-PP  but  not  by  oxythiamine  (Eusebi  and  Ce- 
recedo,  1950;  Navazio  et  al.,  1956),  and  wheat  germ  decarboxylase  behaves 
similarly  (Eich  and  Cerecedo,  1954,  1955).  Pyrithiamine  is  noninhibitory 
whereas  pyrithiamine-PP  is  as  effective  as  oxythiamine-PP  (Woolley,  1951; 
Eich  and  Cerecedo,  1954).  Oxythiamine-PPP  is  inhibitory  (Velluz  and  Her- 
bain,  1951;  Navazio  et  al.,  1956)  but  it  is  possible  that  some  of  the  action 
results  from  the  splitting  off  of  a  phosphate  to  form  oxythiamine-PP,  just 
as  the  cocarboxylase  activity  of  thiamine-PPP  has  been  found  to  depend 
on  hydrolysis  to  thiamine-PP  (Kiessling,  1956).  It  is  not  certain  if  oxy- 
thiamine-P  is  inhibitory  and  contradictory  results  have  been  obtained. 
These  data  indicate  the  importance  of  the  phosphate  groups  for  the  bind- 
ing, and  this  is  supported  by  the  observation  of  Wiethoff  et  al.  (1957)  that 
pyrophosphate  inhibits  wheat  germ  decarboxylase  competitively  with  thia- 
mine-PP (the  pyrophosphate  must  be  added  before  the  thiamine-PP). 

Thiamine-PP  and  the  phosphorylated  analogs  are  bound  fairly  tightly  to 
the  apoenzyme  {K„^  for  thiamine-PP  is  usually  near  0.001-0.003  mM)  and 
thus  the  order  of  addition  of  coenzyme  and  analog  is  important,  especially 
as  the  analogs  are  bound  about  1.5-3.0  kcal/mole  less  tightly  (Woolley, 
1951;  Stewart,  1957).  If  thiamine-PP  is  added  30  min  before  oxythiamine- 
PP  there  is  no  inhibition,  if  they  are  added  simultaneously  there  is  slight 
inhibition,  but  if  oxythiamine-PP  is  added  30  min  before  the  coenzyme 
appreciable  inhibition  may  be  exerted  (Eich  and  Cerecedo,  1954).  Although 

*  It  was  found  later  that  the  material  originally  designated  as  pyrithiamine  was  a 
mixture  (Wilson  and  Harris,  1949).  The  pyridine  analog  was  synthesized  in  pure 
form  and  named  "neopyri thiamine"  but  since  the  active  compound  in  the  early 
preparation  was  this  substance,  it  has  been  generally  agreed  that  the  original  name  be 
restored.  The  general  results  of  the  early  work  are  not  invalidated,  but  the  true  po- 
tency of  pyrithiamine  is  greater  than  indicated  there. 


ANALOGS   OF   THIAMINE  519 

some  exchange  between  bound  and  free  coenzyme  and  analog  must  occur, 
it  is  too  slow  for  equilibrium  to  be  obtained  easily.  To  determine  the  maximal 
inhibiting  power  of  an  analog  it  is  advisable  to  incubate  the  apoenzyme  with 
the  analog  previous  to  addition  of  the  coenzjTne. 

Pyruvate  oxidase  is  inhibited  similarly  to  the  decarboxylase,  as  expected, 
and  in  the  case  of  pyrithiamine-PP  it  would  appear  to  be  competitive 
(Woolley,  1951).  Oxythiamine-PPP  inhibits  pyruvate  oxidation  in  pigeon 
breast  muscle  extracts  but  this  may  be  mediated  through  the  diphosphate 
(Onrust  et  al.,  1952).  The  formation  of  acetoin  from  pyruvate  is  also  inhib- 
ited by  oxythiamine-PP  (Eich  and  Cerecedo,  1954).  However,  Kuratomi 
(1959)  noted  that  oxythiamine,  like  thiamine,  can  form  acetoin  from  py- 
ruvate. 

Transketolase  from  yeast  is  strongly  inhibited  by  oxythiamine-PP  (Datta 
and  Eacker,  1961)  but  oxythiamine  itself  has  no  effect  (Dreyfus  and  Moniz, 
1962).  At  0.036  mM  and  0.072  mM  the  inhibitions  are  60%  and  80%,  re- 
spectively, when  the  analog  is  added  previous  to  thiamine-PP,  but  if  the 
oxythiamine-PP  is  added  2  min  after  the  coenzyme,  no  inhibition  is  ob- 
served. Addition  of  higher  concentrations  of  thiamine-PP  cannot  reverse 
the  inhibition.  Thus  it  is  difficult  for  either  the  analog  or  the  coenzyme  to 
displace  each  other  from  the  apoenzyme.  Oxythiamine-PP  is  bound  more 
tightly  than  thiamine-PP  to  the  enzyme  but  it  requires  2-3  hr  to  inhibit 
50%  when  the  enzyme  initially  contains  thiamine-PP.  The  rate  of  dis- 
placement for  transketolase  is  even  less  than  for  decarboxylase,  since 
thiamine-PP  was  found  to  reverse  oxythiamine-PP-inhibited  enzyme  30% 
in  20  min. 

The  inhibition  of  thiamine-PP-dependent  enzymes  by  oxythiamine  and 
pyrithiamine  will  depend  on  whether  these  analogs  can  be  phosphorylated 
or  not.  Thus  in  the  experiments  of  Kunz  (1956),  where  oxythiamine  and 
pyrithiamine  at  10  vaM  inhibited  pyruvate  oxidation  in  rat  liver  mitochon- 
dria 95%  and  35%,  respectively,  one  is  not  certain  if  there  is  direct  inhibi- 
tion or  if  the  depression  was  due  to  the  formation  of  small  amounts  of  the 
phosphorylated  esters.  Acetylthiamine  inhibits  to  about  the  same  degree 
as  pyrithiamine  and  this  analog  cannot  be  phosphorylated  (unless  it  is  first 
deacetylated),  so  it  would  seem  that  these  rather  weak  inhibitions  may  be 
to  some  extent  exerted  directly.  It  is  also  interesting  that  oxythiamine 
and  pyrithiamine  have  no  effect  on  pyruvate  oxidation  in  brain  mitochon- 
dria, this  being  attributed  by  Kunz  to  a  different  structure  or  permeability 
compared  to  liver  mitochondria;  different  phosphorylative  capacities  might 
also  play  a  role.  Phosphorylation  of  these  analogs  mediated  through  thia- 
mine kinase  seems  to  occur  in  most  tissues  and  it  is  likely,  as  pointed  out 
by  Woolley  (1951),  that  the  toxic  reactions  observed  with  oxythiamine 
and  pyrithiamine  in  animals  are  produced  primarily  by  the  phosphorylated 
compounds. 


520  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Accumulation  of  Intermediates  and  in  Vivo  Effects 

Inasmuch  as  the  oxidation  of  pyruvate  requires  thiamine-PP,  one  would 
expect  some  accumulation  of  pyruvate  in  animals  treated  with  thiamine 
analogs  if  pyruvate  oxidase  is  indeed  inhibited  in  vivo.  Such  has  been  ob- 
served in  rats  with  oxythiamine  (Frohman  and  Day,  1949;  Gubler,  1961) 
and  pyrithiamine  (de  Caro  et  al.,  1954).  For  example,  rats  injected  intra- 
peritoneally  with  150  //g  oxythiamine  show  an  elevation  in  blood  pyruvate 
of  1.3  to  5.1  mg%;  blood  lactate  also  increases  from  11.5  to  42.8  mg%.  In 
mice,  pyrithiamine  raises  the  blood  pyruvate  somewhat  but  oxythiamine 
has  no  effect  (de  Caro  et  al.,  1956),  possibly  indicating  a  species  difference 
since  in  rats  oxythiamine  is  more  effective  than  pyrithiamine  (Gubler, 
1961).  The  administration  of  oxythiamine  to  dogs  at  6  mg/kg  in  three 
doses  leads  to  a  marked  rise  in  blood  pyruvate  (0.4  to  5.7  mg%)  and  thia- 
mine is  able  to  counteract  this  effectively  (Wilson  et  al.,  1962).  Simultan- 
eously there  is  a  severe  fall  in  liver  glycogen  (13  to  0.4  mg/g).  Rats  and 
cats  respond  similarly  but  are  less  sensitive.  Growth  of  Neurospora  in  the 
presence  of  oxythiamine  is  accompanied  by  pyruvate  accumulation  and  a 
simultaneous  reduction  in  pyruvate  decarboxylase  activity  is  demonstrable 
(Sankar,  1958).  Administration  of  increasing  amounts  of  thiamine  partially 
or  completely  counteracts  these  effects  on  pyruvate  levels,  in  all  instances 
where  it  has  been  tested. 

We  shall  now  turn  to  evidence  of  enzyme  inhibition  in  the  tissues  of 
analog-treated  animals.  It  may  be  calculated  from  the  data  of  Von  Holt  et  al. 
(1955)  that  feeding  pjTithiamine  to  rats  at  10  mg/kg  for  7-12  days  results 
in  some  63%  reduction  of  pyruvate  oxidation  in  liver  homogenates.  A 
thorough  investigation  of  the  changing  patterns  of  keto  acid  oxidation  in 
deficient  and  analog-treated  rats  has  been  made  by  Gubler  (1958,  1961); 
his  results  are  summarized  in  Table  2-32.  It  is  seen  that  the  oxidation  of 
pyruvate  is  more  sensitive  than  that  of  «-ketoglutarate  to  both  dietary 
deficiency  and  the  analogs;  this  could  relate  to  different  displacing  rates 
in  the  two  oxidases,  or  to  different  dependencies  of  enzyme  activity  on 
thiamine-PP  level.  Oxidation  of  /5-keto  acids,  as  expected,  is  not  affected. 
The  effects  of  oxythiamine  and  pyrithiamine  are  roughly  the  same  on  all 
tissues,  with  the  exception  of  brain  in  which  pyrithiamine  is  more  effective. 
The  reason  for  this  is  not  understood  —  it  would  seem  unlikely  that  oxy- 
thiamine is  unable  to  penetrate  the  blood-brain  barrier  —  but  it  may  be 
correlated  with  the  fact  that  only  pyrithiamine  is  able  to  produce  poly- 
neuritis in  rats.  Another  difference  between  these  two  analogs  lies  in  the 
ability  of  thiamine-PP  added  in  vitro  to  the  mitochondrial  suspensions  to 
counteract  the  depression  of  pyruvate  oxidation.  The  loss  of  activity  from 
dietary  deficiency  of  thiamine  is  readily  reversed  by  adding  thiamine-PP, 
as  anticipated;  the  loss  due  to  pyrithiamine  is  surprisingly  weU  reversed 
(to  about  90%  of  the  control  values  in  brain  and  kidney);  the  loss  due  to 


ANALOGS   OF  THIAMINE 


521 


Table  2-32 

Effects  of  Thiamine  Deficiency  and  Thiamine  Analogs 
ON  THE  Oxidative  Decarboxylation  of  Keto  Acids  in  Mitochondria  " 


Substrate 

%  Change 

Tissue 

Dietary 
deficiency 

Oxythiamine 

Pyrithiamine 

Liver 

Pyruvate 

-75 

-56 

-57 

a-Ketoglutarate 

-38 

-   1 

-11 

a-Ketoisovalerate 

+  3 

-22 

-13 

a-Keto-^-methylvalerate 

-11 

-11 

-12 

/S-Hydroxybutjrrate 

+  4 

+   2 

+  11 

Brain 

Pyruvate 

-24 

-18 

-51 

a-Ketoglutarate 

+21 

+  6 

-44 

Kidney 

Pyruvate 

-56 

-52 

-61 

a-Ketoglutarate 

-55 

+  5 

-37 

Heart 

PjTuvate 

-32 

-67 

-58 

a-Ketoglutarate 

-24 

-37 

-17 

°  The  analogs  were  administered  to  rats  at  oxythiamine/thiamine  =  200,  and 
pyrithiamine/thiamine  —  5,  these  doses  producing  polyneuritis  in  several  days.  The 
figures  give  the  changes  observed  in  the  oxidation  of  the  substrates  indicated  in 
mitochondrial  suspensions.  (From  Gubler,  1961.) 


oxythiamine  is  reversed  poorly.  The  most  obvious  explanation  of  this  is 
that  pyrithiamine  blocks  the  synthesis  of  thiamine-PP,  so  that  the  tissues 
are  primarily  deficient  in  cocarboxylase,  whereas  oxythiamine  may  exert 
its  inhibition  mainly  by  binding  to  pyruvate  oxidase  in  the  form  of  its 
diphosphate  ester.  These  problems  wiU  be  discussed  after  further  effects 
of  these  analogs  have  been  presented. 

Feeding  pyrithiamine  to  pigeons  leads  to  a  50%  reduction  in  the  p>Tuvate 
decarboxylase  activity  in  breast  muscle,  and  this  can  be  reversed  by  the 
addition  of  thiamine-PP  to  the  homogenates  (Koedam  et  al.,  1956).  This  is 
accompanied  by  a  marked  reduction  in  the  thiamine-PP  content  of  muscle, 
so  that  it  was  concluded  that  there  is  no  essential  difference  between  di- 
etary thiamine  deficiency  and  pjTithiamine  feeding.  Pyruvate  dismutation 
and  acetoin  formation  in  breast  and  heart  muscle  are  likewise  depressed 
by  pyrithiamine  feeding  (Koedam,  1958).  A  large  single  dose  of  pyrithia- 
mine (2.5  mg)  leads  to  a  rapid  inhibition  of  acetoin  formation  and  even  after 
8  days  the  activity  does  not  return  to  normal. 

The  respiratory  quotient  of  rats  treated  with  pyrithiamine  (5  mg/day 


522  2.  ANALOGS  or  enzyme  reaction  components 

intraperitoneally  for  5-6  days)  is  lowered  and,  in  contrast  to  normal  or 
thiamine-deficient  animals,  administration  of  glucose  does  not  raise  it  (see 
accompanying  tabulation)  (de  Caro  et  al.,  1954).  There  is  thus  an  inhibition 


R.Q. 

Before  glucose      After  glucose 

Controls 

Diet-deficient 

Pyrithiamine-treated 

0.81                     0.93 
0.78                     0.92 
0.72                     0.77 

of  the  total  oxidation  of  carbohydrate,  which  can  be  directly  attributed  to 
a  block  in  pyruvate  oxidation;  however,  it  is  difficult  to  explain  the  greater 
effect  in  the  diet-deficient  animals.  The  only  work  showing  an  impairment 
of  transketolase  function  is  that  of  Wolfe  (1957).  Rats  deprived  of  thiamine 
or  given  oxythiamine  show  a  depression  of  the  pentose-P  pathway  in  the 
erythrocytes,  pentose  accumulating,  whereas  pyrithiamine  produces  no 
changes  in  transketolase  activity  even  when  the  animals  are  paralyzed. 

Some  interference  with  amino  acid  metabolism  by  thiamine  analogs,  me- 
diated through  changes  in  the  utilization  of  the  a-keto  acids,  might  be  ex- 
pected, but  little  study  of  this  has  been  made.  Pyrithiamine  inhibits  the 
formation  of  aspartate  and  asparagine  from  glutamate  in  germinating  Pha- 
seolus  seeds  and  there  is  an  accumulation  of  ammonia,  possibly  due  to  the 
lack  of  oxalacetate  (Sivaramakrishnan  and  Sarma,  1954).  These  effects  can 
be  attributed  to  a  reduction  in  a-ketoglutarate  oxidase.  Oxythiamine  and 
pyrithiamine  both  inhibit  the  growth  of  Vibrio  cholera  and  lead  to  the  accu- 
mulation of  alanine,  aspartate,  and  glutamate.  The  effects  on  the  levels  of 
such  amino  acids  will  depend  to  a  great  extent  on  the  pattern  of  metabolism, 
i.e.,  on  the  over-all  direction  of  transamination  reactions. 

Effects   on   Thiamine-PP  Synthesis 

Pyrithiamine  has  been  found  to  inhibit  thiamine  kinase  from  chicken 
blood  (Woolley,  1950  a),  rat  liver  (Eich  and  Cerecedo,  1954;  Mano  and  Ta- 
naka,  1960),  rat  intestine  (Cerecedo  et  al.,  1954),  and  pigeon  liver  (Koedam, 
1958),  and  the  inhibition  appears  to  be  competitive  with  thiamine.  This  in- 
hibition is  fairly  potent:  When  the  ratio  pyrithiamine/thiamine  is  1  the  in- 
hibition is  around  50%,  at  a  ratio  of  5  it  is  75%,  and  at  a  ratio  of  10  it  is 
90%  in  rat  liver  and  intestine  (thiamine  concentrations  0.01-0.1  raM).  On 
the  other  hand,  oxythiamine  inhibits  thiamine  kinase  much  less  strongly 
or  not  at  all  when  present  in  similar  ratios  to  thiamine  (Eich  and  Cerecedo, 
1954;  Cerecedo  et  al,  1954;  Koedam,  1958;  Mano  and  Tanaka,  1960).  The 


ANALOGS   OF  THIAMINE  523 

difference  in  inhibitory  activity  between  these  two  analogs  possibly  indi- 
cates the  importance  of  the  pyrimidyl  4 '-amino  group.  This  was  substan- 
tiated in  the  demonstration  by  Cerecedo  and  Eich  (1955)  that  oxypyri- 
thiamine,  in  which  the  4'-amino  group  is  replaced  by  hydroxyl  group,  does 
not  inhibit  rat  liver  thiamine  kinase. 

Mano  and  Tanaka  (1960)  studied  a  large  series  of  thiamine  analogs  with 
respect  to  their  abilities  to  be  phosphorylated  by  a  rat  liver  thiamine  kinase 
system  and  their  inhibitory  potencies  on  thiamine  phosphorylation  (see  ac- 
companying tabulation).  The  analogs  and  thiamine  were  all  at  0.1  milf. 


Relative  activity  %  Inhibition  of 

(thiamine   —   100)  thiamine  kinase 


Pyrithiamine 

0 

53 

2'-Ethylthiamine 

11 

38 

2'-/i-Butylthiamine 

0 

32 

Oxythiamine 

0 

6 

Diacetylthiamine 

3 

5 

Thiothiamine 

0 

Stim     2 

Dibenzoylthiamine 

2 

Stim     3 

0-Acetylthiamine 

7 

Stim     6 

Thiamine  disulfide 

0 

Stim  12 

The  inhibitions  by  pyrithiamine  and  the  2'-alkylthiamines  are  competitive 
and  the  following  values  of  K^  were  calculated:  pyrithiamine  0.033  m.M, 
2'-ethylthiamine  0.041  mM,  and  2'-butylthiamine  0.043  mM.  The  inability 
of  the  enzyme  to  catalyze  the  phosphorylation  of  oxythiamine  and  pyri- 
thiamine is  noteworthy  in  view  of  the  theory  that  these  analogs  may  exert 
their  effects  in  the  diphosphate  form.  It  may  be  recalled  that  Woolley  (1951 ) 
failed  to  demonstrate  the  synthesis  of  pyrithiamine-PP  in  chicken  blood. 
There  has  been  surprisingly  little  attention  to  this  important  problem  of 
analog  phosphorylation  in  organisms. 

Inhibition  of  Thiaminase 

The  importance  of  this  enzyme  in  the  mammalian  metabolism  of  thiamine 
is  not  known.  The  only  evidence  for  its  possible  function  is  the  appearance 
in  the  urine  of  the  pyrimidine  and  thiazole  moieties  of  thiamine  following 
administration  of  thiamine.  It  would  seem  unlikely  that  inhibition  of  this 
enzyme  is  an  important  factor  in  the  toxicity  of  thiamine  analogs,  but  it 
could  be  of  some  significance  in  determining  the  effects  on  tissue  levels  of 
thiamine-PP.  It  was  stated  by  Soodak  and  Cerecedo  (1944)  that  oxythia- 
mine inhibits  carp  thiaminase  but  no  data  were  given.  Pyrithiamine  inhibits 


524  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

this  enzyme  around  40%  when  analog  and  thiamine  are  both  0.5  mM  (Sea- 
lock  and  White,  1949).  Pyrithiamine  is  split  by  the  enzyme  but  at  a  slower 
rate  than  is  thiamine.  Apparently  it  is  bound  more  tightly  than  thiamine, 
but  reacts  more  slowly,  since  in  mixtures  of  the  two  only  the  splitting  of 
thiamine  is  depressed.  Thus  too  little  is  known  of  the  effects  of  these  analogs 
on  thiaminase  to  evaluate  the  importance  of  the  inhibitions. 

Sealock  examined  the  effects  of  a  series  of  substituted  methylthiazolium 
ions  on  fish  thiaminase  and  found  that  3-o-aminobenzyl-4-methylthiazole 
(ABMT)  is  a  particularly  potent  inhibitor  (see  tabulation)  (Sealock  and 
Goodland,  1944).  The  inhibition  is  competitive,  with  K^^  =  0.0831  mM  and 
Ki  =  0.00197  mM,  possibly  indicating  that  ABMT  is  bound  2.3  kcal/mole 


Analog 


3-o-Aminobenzyl-4-methylthiazole 

3-^-Aminoethyl-4-methylthiazole 

3-/3-Phthalimidoethyl-4-methylthiazole 

3-o-Nitrobenzyl-4-methy]thiazole 

3-Ethyl-4-methylthiazole 

3-Phenyl-4-methylthiazole 

3-Phenyl-2-methyl-4-methylthiazole 

3-Ethyl-2-methyl-4-methylthiazole 


Concentration 
(mM) 

Inhibition 

Relative  "X/' 

0.5 

100 

<  0.005 

0.5 

48 

0.54 

0.5 

6 

7.8 

0.5 

2 

24 

10 

9 

101 

10 

5 

190 

5 

0 

>495 

10 

0 

>990 

more  tightly  than  thiamine.  The  amino  group  is  probably  quite  important, 
since  the  replacement  with  a  nitro  group  reduces  the  inhibition  so  markedly. 
The  aminobenzylthiazoles  were  later  studied  in  greater  detail  (Sealock  and 
Livermore,  1949)  and  the  position  of  the  amino  group  was  shown  to  be 
critical,  only  the  ortho  compound  being  inhibitory  (see  accompanying  ta- 
bulation). On  the  other  hand,  the  position  of  the  thiazole  methyl  group  is 
not  critical.  Thiamine  was  0.5  mM  in  these  experiments.  Kenten  (1958)  has 


Amino  position  in  Methyl  position  in  Concentration  „,    ^  i  -i  •  • 

,  ,     .  XI,-       1       •  /     jiT\  /o  Inhibition 

benzyl  ring  thiazole  ring  (mM) 


ortho 
meta 
para 
ortho 
meta 
para 
ortho 


4 

0.5 

78.1 

4 

0.5 

Stim 

4 

0.5 

1.7 

2 

0.5 

89.2 

2 

0.5 

Stim 

2 

1 

2.0 

2,4 

0.5 

30.2 

ANALOGS    OF   THIAMINE  525 

found  the  thiaminase  from  bracken  {Pteridium  aquilinum)  to  be  very  sen- 
sitive to  ABMT,  15-20%  inhibition  being  given  by  0.002  mM  and  almost 
complete  inhibition  by  0.05  mM.  The  inhibition  is  probably  basically  com- 
petitive since  it  proceeds  faster  in  the  absence  of  thiamine.  As  far  as  I  know, 
this  interesting  compound  has  not  been  tested  in  whole  animals  to  deter- 
mine if  thiaminase  inhibition  can  be  achieved  and  how  this  will  alter  thia- 
mine metabolism. 

Effects   on    Excretion   and   Tissue   Levels  of  Thiamine 

If  these  analogs  displace  thiamine  or  thiamine-PP  from  the  tissues  in 
any  way,  or  inhibit  the  transport  or  metabolism  of  thiamine,  an  increased 
urinary  excretion  of  thiamine  would  be  expected,  and  this  has  been  found 
to  occur  in  rats  given  50 //g  oxythiamine  (Frohman  and  Day,  1949).  One 
might  also  predict  that  tissue  levels  of  thiamine  or  its  diphosphate  would 
be  reduced,  and  this  has  been  demonstrated  for  both  pyrithiamine  and  oxy- 
thiamine in  mice,  pigeons,  and  rats.  The  depression  of  tissue  thiamine-PP 
seems  to  be  generally  associated  with  a  rise  in  blood  pyruvate,  so  that  at 
least  part  of  this  depletion  is  related  to  enzymes  involved  in  the  oxidation 
or  decarboxylation  of  a-keto  acids.  Inasmuch  as  theories  for  the  mechan- 
isms by  which  these  analogs  act  depend  on  the  changes  in  tissue  thiamine 
levels,  it  will  be  necessary  to  examine  the  results  with  some  care. 

Pyrithiamine  markedly  depletes  the  tissues  of  thiamine-PP  in  pigeons. 
Controls  were  fed  100  //g  thiamine  per  day  and  another  group  was  fed  623  //g 
pyrithiamine  each  day  in  addition;  after  an  average  survival  time  of  19  days, 
the  thiamine-PP  levels  in  the  tissues  were  those  shown  in  the  following 
tabulation  (Koedam  et  al.,  1956).  The  pyruvate  decarboxylase  activity  in 
muscle  is  reduced  around  50%  and  adding  thiamine-PP  restores  activity. 


Thiamine-PP  content  (/<g/g) 


Tissue 




%  Change 

Controls 

Pyrithiamine-fed 

4.55 

1.00 

-78 

2.61 

0.78 

-70 

3.54 

1.09 

-69 

4.42 

1.80 

-59 

4.36 

2.08 

-52 

Heart 

Brain 

Liver 

Breast  muscle 

Kidney 


The  fall  in  thiamine-PP  level  is  quite  rapid;  at  4  days  it  is  mainly  complete 
in  most  tissues,  and  from  the  data  on  pjTuvate  utilization  it  would  appear 
that  a  marked  decrease  occurs  within  1  day  (Koedam,  1958).  Pigeons  given 
a  single  large  dose  of  pyrithiamine  (10  mg)  and  examined  64  days  later  show 


526  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

no  permanent  effect  on  the  tissue  thiamine-PP  levels.  An  important  obser- 
vation was  that  pyrithiamine  induces  a  more  rapid  depletion  of  tissue  thia- 
mine-PP than  does  elimination  of  exogenous  thiamine.  A  comparison  be- 
tween dietary  deficiency  and  pyrithiamine  administration  in  rats  was  report- 
ed by  de  Caro  et  at.  (1954).  The  results  in  the  accompanying  tabulation  were 


Total  thiamine  content  (//g/g) 


Tissue 


Controls  Avitaminotic  Pyrithiamine-fed 


Liver  7.42  2.82  1.33 

Muscle  1.60  0.80  0.71 

Brain  3.38  2.89  0.57 


obtained  from  rats  injected  with  5  mg  pyrithiamine  daily  for  5-6  days  and 
rats  subjected  to  a  thiamine-free  diet  for  a  comparable  time.  In  all  cases 
pyrithiamine  produces  a  greater  effect  than  simple  elimination  of  thiamine 
intake;  the  effect  in  brain  is  particularly  striking  and  possibly  correlated 
with  the  polyneuritic  symptoms  produced  by  pyrithiamine. 

Oxythiamine,  on  the  other  hand,  does  not  seem  to  be  so  active  in  reducing 
the  tissue  levels  of  thiamine-PP  (Steyn-Parve,  1954).  This  analog  at  1  mg/ 
day  for  15  days  to  pigeons  produces  the  changes  summarized  in  the  accom- 
panying tabulation.  No  deficiency  symptoms  were  noted  and  none  would 


Thiamine-PP  content  (fig/g) 


Tissue 


Controls  Oxythiamine-fed  %  Change 


Heart 
Muscle 
Cerebrum 
Liver 


6.70 

3.40 

-49 

6.25 

4.60 

-26 

4.15 

3.75 

-10 

5.35 

4.90 

-  8 

be  expected  at  these  tissue  levels.  The  author  believed  that  the  change  in 
the  liver  is  not  significant;  it  is  also  possible  that  the  large  drop  in  the  heart 
thiamine-PP  is  too  great,  since  in  another  experiment  with  twice  the  above 
oxythiamine  dosage  the  level  decreases  only  34%.  The  relative  ineffective- 
ness of  oxythiamine  was  confirmed  by  de  Caro  et  al.  (1956)  in  mice,  where 
0.5-2  mg/day  certainly  produces  little  effect  on  the  thiamine  levels  in  muscle 
and  brain,  although  some  decrease  in  liver  is  observed.  There  are  likewise  no 
significant  change  in  blood  pyruvate.  These  results  were  confirmed  and  ex- 
tended by  Gurtner  (1961),  who  administered  pyrithiamine  at  250  /yg/day 


ANALOGS   OF  THIAMINE  527 

and  oxy thiamine  at  10  mg/day  to  rats  intraperitoneally  for  29  days;  a 
thiamine-deficient  group  was  also  included  (see  accompanying  tabulation). 


Pyrithiamine 

Oxythiamine 

^i-deficient 

Weight  (%  change) 

-15 

-43 

-44 

Cardiac  rate  (%  change) 

—  5 

-18 

-28 

Paralysis  (%  occurrence) 

93 

0 

46 

Convulsions  (%  occurrence) 

73 

0 

20 

Blood  pyruvate  (%  change) 

+  15 

+288 

+95 

Tissue  thiamine-PP  (%  change) 

Liver 

-35 

+26 

-97 

Heart 

-76 

-20 

-95 

Brain 

-86 

-  6 

-85 

The  differences  in  the  actions  of  the  two  analogs  is  well  illustrated  here; 
pyrithiamine  produces  marked  neurological  symptoms  without  much  effect 
on  weight,  cardiac  rate,  or  blood  pyruvate,  although  there  is  a  very  signi- 
ficant fall  in  tissue  thiamine-PP,  whereas  oxythiamine  causes  bradycardia 
and  weight  loss  without  neurological  effects,  while  the  blood  pyruvate  is 
elevated  greatly  without  significant  changes  in  tissue  thiamine-PP. 

Tissue  Levels  of  Pyrithiamine 

The  pyrithiamine  in  rat  tissues  following  the  injection  of  1  mg  intraperi- 
toneally was  determined  microfluorimetrically  by  Rindi  and  Perri  (1961) 
and  Rindi  et  al.  (1961)  and  the  results  are  plotted  in  Fig.  2-18.  These  rats 
had  been  maintained  on  a  thiamine-deficient  diet  and  within  1  day  the  pyri- 
thiamine content  of  the  tissues  corresponded  closely  to  the  normal  thiamine 
content,  indicating  that  the  analog  probably  occupies  the  binding  sites  nor- 
mally occupied  by  thiamine  or  its  diphosphate.  It  was  also  shown  that  prac- 
tically all  of  the  pyrithiamine  in  the  liver  is  phosphorylated.  The  concen- 
tration in  the  brain  increases  progressively  throughout  the  12  days  of  the 
experiment  and  it  is  likely  that  this  reflects  a  transference  of  pyrithiamine 
from  the  liver,  the  analog  gradually  replacing  the  thiamine  in  the  brain. 
Daily  oral  administration  of  33  //g  thiamine  and  210  //g  pyrithiamine  leads 
to  a  slow  but  very  definite  rise  in  tissue  pyrithiamine  over  20  days,  the  levels 
eventually  reached  being  higher  than  following  the  single  intraperitoneal 
injection.  It  is  unfortunate  that  we  have  no  data  on  oxythiamine  distribu- 
tion in  the  tissues,  since  this  might  well  help  to  answer  some  of  the  problems 
as  to  why  its  effects  are  often  quite  different  from  those  of  pyrithiamine. 


528 


2,  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


TIME  (DAYS) 


Fig.  2-18.  Pyrithiaraine  concentrations  in  rat  tissues   after  intra- 
peritoneal injection  of  1  mg.  (From  Rindi  and  Perri,  1961.) 


Effects  on  the  Growth  of  Microorganisms 

The  degree  of  inhibition  of  various  bacteria  and  fungi  by  pyrithiamine 
has  been  related  to  the  pattern  of  thiamine  biosynthesis  (Robbins,  1941; 
Woolley  and  White,  1943  c).  Sensitivity  to  pyrithiamine  was  correlated  with 
a  requirement  for  intact  thiamine,  whereas  those  organisms  able  to  syn- 
thesize thiamine  completely  are  poorly  inhibited.  If  the  organisms  require 
only  part  of  the  thiamine  molecule,  the  growth  depression  by  pyrithiamine 
is  intermediate.  If  pyrithiamine  interferes  with  either  the  formation  or 
enzymic  function  of  thiamine-PP,  it  would  be  difficult  to  understand  how 
the  manner  of  obtaining  thiamine  could  determine  sensitivity  to  the  analog. 
However,  another  factor  must  be  considered.  It  was  shown  that  pyrithia- 
mine-resistant  organisms  possess  an  enzyme  capable  of  cleaving  pyrithia- 
mine, probably  a  thiaminase,  while  sensitive  thiamine-requiring  organisms 
do  not.  At  least  part  of  the  resistance  might  be  attributed  to  the  ability 
of  these  organisms  to  inactivate  the  analog;  the  pyridine  portion  split  from 
pyrithiamine  would  not  be  inhibitory  and  the  pyrimidine  portion  can  ac- 


ANALOGS   OF   THIAMINE  529 

tually  be  utilized  in  thiamine  synthesis.  Supporting  the  importance  of  a 
pyrithiaminase  in  resistance  is  the  observation  by  Woolley  (1944  a)  that  a 
pyrithiamine-resistant  strain  of  Endomyces  vernalis,  obtained  by  subcul- 
turing  in  increasing  concentrations  of  the  analog  and  capable  of  withstand- 
ing 25  times  the  concentration  initially  depressing  growth  50%,  contains 
such  an  enzyme.  Indeed,  pyrithiamine  is  capable  of  stimulating  growth  in 
the  absence  of  thiamine  since  the  pyrimidine  portion  (which  is  all  that  is 
required  by  this  organism)  is  provided  by  the  splitting  reaction.  However, 
destruction  of  pyrithiamine  is  not  the  only  factor  involved,  since  enough 
of  the  analog  remains  unsplit  to  inhibit  completely  the  parent  strain.  This 
enzyme  may  be  functional  in  the  pathway  biosynthesizing  thiamine,  which 
would  be  the  reason  for  the  correlation  with  thiamine  requirements.  It  might 
also  be  well  to  consider  another  possible  mechanism  of  inhibition,  a  block 
of  the  transport  of  thiamine  into  the  cells;  only  those  organisms  requiring 
intact  thiamine  would  be  susceptible.  A  strain  of  S.  aureus  adapted  to  py- 
rithiamine exhibits  a  variety  of  changes:  the  pigment  color  changes  from 
orange  to  lemon  yellow,  glucose  utilization  is  severely  depressed,  and  ace- 
tate utilization  is  increased  (Das  and  Chatterjee,  1962).  A  partial  blocking 
of  the  pentose-P  pathway  was  also  observed.  These  results  indicate  the 
complex  alterations  occurring  during  the  development  of  resistance. 

Some  of  the  effects  on  microorganisms  will  be  briefly  summarized,  since 
most  of  this  work  has  no  direct  bearing  on  the  mechanism  of  inhibition. 
In  most  cases  the  growth  depression  by  the  analogs  is  counteracted  by  thia- 
mine, as  in  the  inhibition  of  Neurospora  crassa  by  oxythiamine  (Sankar, 
1958),  and,  at  least  in  some  cases,  the  inhibition  is  formally  cimpetit"'e 
with  thiamine  (Quesnel,  1956).  Growth  depression  can  depend  on  various 
factors.  For  example,  Phycomyces  blakesleeanus  becomes  more  resistant  to 
pyrithiamine  with  culture  age,  the  concentration  required  for  50%  inhibi- 
tion being  8  times  greater  at  13  days  than  at  4  days  (Fluri,  1959).  Is  this 
due  to  an  alteration  of  thiamine  metabolism  with  age,  or  to  different  me- 
tabolic requirements  for  thiamine  ?  No  change  in  sensitivity  to  oxythiamine 
with  age  was  noted.  Furthermore,  oxythiamine  seemed  to  induce  a  thiamine 
deficiency,  determined  by  changes  in  carbohydrate  content,  whereas  pyri- 
thiamine did  not.  Euglena  gracilis  occurs  in  a  normal  green  form  and  a 
white  form  (chlorophyll-deficient  from  streptomycin  treatment):  The  white 
form  is  about  5  times  more  sensitive  to  pyrithiamine  than  is  the  green  form 
(Schopfer  and  Keller,  1951).  Thiamine  analogs  have  been  considered  as 
possibly  useful  in  certain  infections.  The  growth  of  Microsporum  audouini 
is  very  strongly  inhibited  by  0.0012  mM  pyrithiamine  and  the  use  of  the 
analog  in  tinea  capitis  was  suggested  (Ulrich  and  Fitzpatrick,  1951).  The 
infection  of  wheat  with  leaf  rust  (Puccinia)  might  be  controlled  with  oxy- 
thiamine inasmuch  as  this  substances  exerts  a  selective  action  on  the  fungus 
when  isolated  infected  leaves  are  tested  (Samborski  and  Forsyth,  1960). 


530  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

A  concentration  of  0.75  n\M  inhibits  rust  development  completely  and  does 
not  exhibit  phytotoxicity.  Vibrio  cholera  is  inhibited  moderately  by  both 
oxythiamine  and  pyrithiamine  (Chatterjee  and  Haider,  1960),  Lactobacillus 
fermentum  is  inhibited  by  the  imidazole  analog  of  thiamine  (Erlenmeyer  et 
al.,  1948),  and  E.  coli  is  inhibited  50%  by  the  methylthio  analog  of  thia- 
mine at  a  ratio  of  100  with  respect  to  thiamine  (Ulbricht  and  Gots,  1956). 
Growth  inhibition  by  thiamine  analogs  has  been  reviewed  by  Rogers  (1962), 

Toxic  and  Thiamine-Deficiency  Effects  in  Animals 

Both  pyrithiamine  and  oxythiamine  are  toxic  to  animals  and  produce 
states  apparently  related  to  thiamine  deficiency.  These  analogs  are  of  com- 
parable potency;  in  most  species  pyrithiamine  may  be  slightly  more  active 
on  a  weight  basis.  The  usual  daily  doses  to  induce  the  characteristic  toxic 
reactions  and  eventual  death  are  usually  between  0.01  and  0.1  mg,  but  this 
depends  on  the  thiamine  intake,  the  effective  ratios  of  analog/thiamine  be- 
ing around  5  to  50.  The  sequence  of  reactions  following  administration  of 
pyrithiamine  to  mice  or  rats  may  be  summarized  as:  decreased  food  intake 
(this  may  be  noted  within  24  hr),  inactivity  and  a  hunched  position,  ner- 
vousness, tremors  and  occasional  convulsions,  spasticity  followed  by  weak- 
ness of  the  legs,  incoordination,  and  paralysis.  Death  usually  occurs  within 
24  hr  after  the  development  of  polyneuritis.  These  are  essentially  the  symp- 
toms seen  in  thiamine  deficiency  but  they  occur  more  rapidly  after  the  ana- 
logs. Full  polyneuritis  and  death  may  be  produced  within  5-12  days  de- 
pending on  the  dose.  Pyrithiamine  also  produces  typical  thiamine-deficiency 
polyneuritis  in  pigeons.  The  effects  of  oxythiamine  in  mice  and  rats  are 
somewhat  different,  although  death  may  occur  in  approximately  the  same 
time  as  from  pyrithiamine.  There  is  also  anorexia  and  weight  loss,  and  the 
animals  may  become  nervous,  convulsive,  and  incoordinated  during  the 
first  24  hr,  but  the  later  characteristic  symptoms  of  polyneuritis  do  not 
occur.  Descriptions  of  the  later  reactions  to  oxythiamine  have  generally 
been  inadequate.  In  chicks  apparently  both  analogs  can  induce  polyneuritic 
states.  The  above  summary  is  derived  mainly  from  the  work  of  Woolley 
and  White  (1943  b),  Eusebi  and  Cerecedo  (1949),  Daniel  and  Norris  (1949), 
Frohman  and  Day  (1949),  Cerecedo  et  al.  (1951),  Naber  et  al.  (1954),  and 
Wolfe  (1957). 

The  differences  between  the  effects  of  pyrithiamine  and  oxythiamine  in 
mice  and  rats  have  been  emphasized  by  several  workers,  particularly  the 
absence  of  polyneuritis  during  treatment  with  oxythiamine,  and  have  ini- 
tiated speculations  on  the  different  mechanisms  of  action.  It  must  be  made 
clear  that  the  toxic  effects  of  oxythiamine  are  not  nonspecific  and  unre- 
lated to  thiamine  function,  since  the  reactions  to  both  analogs  may  be 
counteracted  by  administration  of  thiamine  (Woolley  and  White,  1943  b; 
Jones  et  al.,  1948;  Daniel  and  Norris,  1949;  Cerecedo  et  al.,  1951;  and  others). 


ANALOGS   OF  THIAMINE  531 

Oxypyrithiamine  reduces  the  survival  time  of  mice  but  does  not  produce 
polyneuritis,  so  that  it  appears  to  behave  like  oxythiamine,  indicating  the 
importance  of  the  4'-amino  group  on  the  pyrimidine  ring  for  the  effects  on 
the  nervous  system  (Cerecedo  and  Eich,  1955).  The  2'-w-butyl  analog  of 
thiamine  suppresses  the  growth  of  rats  and  leads  to  polyneuritis,  these  ef- 
fects being  antagonized  by  increased  thiamine  administration,  so  that  this 
analog  superficially  acts  like  pyrithiamine  (Emerson  and  Southwick,  1945). 
Another  substance  possibly  interfering  with  thiamine  metabolism  is  2,4- 
diamino-5-phenylthiazole  (amiphenazole,  Daptazole),  a  drug  used  as  a  res- 
piratory stimulant.  Rats  on  a  thiamine-free  diet  given  injections  of  ami- 
phenazole and  the  pyrimidine  portion  of  thiamine  in  low  doses  do  not  show 
deficiency,  indicating  some  ability  to  replace  the  normal  thiazole  compo- 
nent, but  at  higher  doses  deficiency  signs  appear  sooner  (Shulman,  1956). 
Amiphenazole  alone  even  at  high  doses  produces  no  effects.  An  abnormal 
thiamine  analog  is  apparently  synthesized  in  the  animals.  The  2-trifluoro- 
methyl  analog  of  thiamine  (trifluorothiamine)  administered  at  100  mg/kg/ 
day  to  mice  on  the  thiamine-deficient  diet  leads  to  weight  loss,  paralysis, 
convulsions,  and  an  inhibition  of  the  growth  of  transplanted  carcinoma 
(Barone  et  al.,  1960).  It  inhibits  the  growth  of  Bacillus  subtilis  more  po- 
tently than  does  oxythiamine  or  pyrithiamine;  this  effect  is  antagonized 
by  thiamine,  but  is  enhanced  by  either  the  pyrimidine  or  thiazole  moieties 
of  thiamine.  Further  study  of  this  interesting  analog  will  be  awaited  with 
anticipation. 

A  few  miscellaneous  observations  relative  to  the  pharmacological  effects 
of  thiamine  analogs  on  neuromuscular  function  will  be  summarized  because 
of  the  importance  of  such  effects  in  developing  theories  of  the  mechanisms 
of  action.  It  has  long  been  known  that  thiamine  is  involved  in  the  forma- 
tion of  acetylcholine  (providing  the  acetyl  radical  from  pyruvate),  and  that 
brain  acetylcholine  concentration  falls  during  thiamine  deficiency;  it  is  pos- 
sible that  some  of  the  effects  of  the  analogs  are  mediated  by  a  depression 
of  acetylcholine  synthesis  at  synapses.  Some  have  claimed  that  thiamine 
is  functional  in  axon  conduction  through  its  role  in  the  synthesis  of  acetyl- 
choline, and  others  have  reported  a  release  of  thiamine  during  nerve  stim- 
ulation. It  is  also  established  that  thiamine  in  rather  high  concentration 
inhibits  cholinesterase  and  can,  under  certain  circumstances,  augment  the 
action  of  acetylcholine.  Pyrithiamine  at  1-3  roM  decreases  the  rates  of  de- 
polarization and  repolarization  during  the  action  potential  in  frog  nerve, 
whereas  oxythiamine  at  these  or  higher  concentrations  produces  no  effect 
(Kunz,  1956).  This  was  interpreted  as  a  blocking  of  the  Na+  carrier  me- 
chanism and  as  evidence  for  the  participation  of  thiamine  in  Na+  transport. 
Depolarization  is  associated  with  Na+  entry,  but  repolarization  in  nerve  is 
not  connected  directly  to  Na+  flux.  These  concentrations  are  much  higher 
than  occur  following  administration  to  animals.  The  tibialis  twitch  response 


532  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

in  the  cat  is  depressed  by  thiamine,  pyrithiamine,  and  any  of  the  pyrithia- 
mine  analogs  having  a  hydroxyl  group  on  the  pyridinium  ring;  furthermore, 
the  neuromuscular  blocks  produced  by  tubocurarine  and  decamethonium 
are  antagonized  by  these  compounds  (Ngai  et  al.,  1961).  In  the  absence  of 
a  hydroxyl  group,  there  is  a  potentiation  of  twitch  tension.  Changes  in 
blood  pressure  parallel  those  in  twitch  tension.  These  compounds,  of  course, 
may  bear  some  relationship  to  acetylcholine  because  of  the  quaternary  ni- 
trogens and  other  functional  groups  an  appropriate  distance  away,  and  it 
is  likely  that  these  acute  effects  are  unrelated  to  the  metabolic  aspects  of 
thiamine.  It  is  interesting,  however,  that  oxythiamine  is  without  activity 
on  the  neuromuscular  junction  and  respiration,  although  it  causes  a  fall  in 
blood  pressure.  No  interference  by  the  analogs  in  the  actions  of  thiamine 
was  reported.  Injections  of  thiamine,  thiamine-PP,  pyrithiamine,  and  oxy- 
thiamine into  frogs  cause  a  miosis,  which  was  interpreted  as  a  direct  effect 
on  the  iris  (Ber  and  Singer- Altbeker,  1961).  It  is  possible  that  this  is  me- 
diated through  inhibition  of  cholinesterase,  and  it  is  not  necessary  to  pos- 
tulate special  neural  or  muscular  functions  for  thiamine.  It  would  be  more 
valuable  to  study  the  possible  changes  in  neuro-muscular  activity  during 
administration  of  the  analogs  chronically  and  when  there  are  evident  motor 
disturbances. 

Mechanisms  of  Action  and  Comparison  of  Pyrithiamine  and  Oxythiamine 

Pyrithiamine  has  been  shown  to  do  the  following:  (1)  produce  polyneur- 
itis as  in  dietary  thiamine  deficiency,  (2)  deplete  various  tissues  of  thia- 
mine-PP and  increase  its  renal  excretion,  (3)  inhibit  a-keto  acid  metabolism 
in  vivo,  which  is  reversed  by  adding  thiamine-PP,  (4)  cause  elevation  of 
blood  pyruvate,  (5)  inhibit  the  phosphorylation  of  thiamine  (thiamine  ki- 
nase), (6)  apparently  be  phosphorylated  in  the  tissues  to  pyrithiamine-PP, 
(7)  in  the  diphosphate  form  inhibit  pyruvate  decarboxylase,  pyruvate  oxi- 
dation, and  probably  transketolase,  and  (8)  be  picked  up  by  the  tissues  to 
about  the  same  extent  as  is  thiamine  normally.  Most  of  these  effects  are 
counteracted  by  the  administration  of  sufficient  thiamine.  It  is,  therefore, 
not  difficult  to  establish  possible  sites  of  pyrithiamine  inhibition;  most  of 
the  reactions  to  pyrithiamine  can  be  explained  on  the  basis  of  either  a 
block  in  thiamine  phosphorylation  or  a  direct  inhibition  of  the  enzymes 
utilizing  thiamine-PP  through  its  diphosphate  ester.  If  the  site  were  only 
on  the  kinase,  the  fundamental  effect  would  be  a  depletion  of  thiamine-PP 
such  as  occurs  in  dietary  deficiency,  and  addition  of  thiamine-PP  to  tissue 
preparations  should  restore  the  activity  of  pyruvate-metabolizing  enzymes 
completely.  The  reversal  is,  however,  only  partial  (Gubler,  1961)  and  by 
no  means  as  great  as  in  diet-deficient  animals.  It  is  thus  likely  that  both 
mechanisms  play  a  role.  The  appearance  of  pyrithiamine  mainly  in  the  di- 
phosphate form  in  tissues  (Rindi  and  Perri,  1961)  also  points  to  the  im- 


ANALOGS   OF  THIAMINE  533 

portance  of  direct  a-keto  acid  enzyme  inhibition.  The  rapidity  with  which 
pyrithiamine  exerts  its  toxicity  (Eusebi  and  Cerecedo,  1949),  relative  to 
dietary  deficiency,  would  indicate  an  effect  other  than  a  block  of  thiamine- 
PP  synthesis.  The  exact  role  of  thiamine  kinase  inhibition  cannot  be  eval- 
uated at  this  time. 

When  we  turn  to  oxythiamine  the  problem  becomes  more  complex.  The 
major  differences  from  the  actions  of  pyrithiamine  may  be  summarized  as 
follows:  (1)  typical  polyneuritis  is  not  produced,  (2)  it  is  not  as  effective 
in  reducing  tissue  thiamine-PP  levels,  particularly  in  the  brain,  (3)  it  does 
not  produce  a  significant  depression  of  pyruvate  oxidation  in  the  brain  in 
vivo,  (4)  inhibition  of  thiamine  kinase  is  slight  or  absent,  and  (5)  its  toxic 
effects  are  more  readily  overcome  by  thiamine.  The  most  obvious  explana- 
tion would  be  that  oxythiamine  has  generally  the  same  actions  as  pyrithia- 
mine in  most  tissues  but  for  some  reason  does  not  interfere  readily  with 
thiamine  function  in  the  nervous  system.  Failure  to  penetrate  into  nerve 
tissue  would  adequately  account  for  this  but  there  is  no  direct  evidence  for 
this,  unless  the  failure  of  oxythiamine  to  affect  the  membrane  potentials 
of  frog  nerve,  although  pyrithiamine  is  effective,  is  interpreted  in  this  way. 
One  aspect  that  has  usually  not  been  considered  is  the  metabolism  of  these 
analogs  in  the  tissues,  although  Cerecedo  et  al.  (1951)  felt  that  oxythiamine 
is  more  rapidly  metabolized  than  pyrithiamine.  Just  as  in  resistant  bacteria, 
resistant  tissues  may  contain  a  thiaminase-like  enzyme  capable  of  destroy- 
ing the  analog,  and  it  is  possible  that  a  particular  tissue  can  inactivate  one 
of  these  analogs  more  than  the  other. 

If  the  major  pathway  of  thiamine  in  animal  tissues  is  simply  (1 )  the  trans- 
port of  thiamine  into  the  cells,  (2)  its  phosphorylation,  and  (3)  the  com- 
bination of  the  thiamine-PP  with  the  apoenzymes,  the  analogs  must  act  on 
one  of  these  steps.  Pyrithiamine  inhibits  steps  (2)  and  (3),  while  oxythia- 
mine affects  only  (3).  No  examination  of  interference  with  transport  systems 
has  been  reported  but  this  should  be  done.  The  question  of  whether  thiamine 
has  actions  additional  to  its  function  in  cc-cleavage,  especially  in  nervous 
tissue,  must  arise  (Woolley  and  Merrifield,  1952;  Gubler,  1961),  but  I  doubt 
if  the  evidence  from  the  use  of  analogs  is  sufficient  at  the  present  time  to 
imply  mechanisms  other  than  on  the  established  systems.  One  cannot  judge 
the  state  of  a  tissue  function  from  changes  in  blood  pyruvate  (which  re- 
flects changes  throughout  the  whole  animal  and  perhaps  particularly  in  the 
liver);  if  central  nervous  system  effects  are  to  be  evaluated,  alterations  in 
the  metabolism  in  the  nerve  cells  must  be  determined.  Also  one  must  al- 
ways consider  the  relationship  between  cell  function  and  thiamine-PP  level. 
To  what  degree  must  thiamine-PP  in  the  brain  fall  before  symptoms  occur? 
This  has  often  been  judged  by  experiments  in  diet-deficient  animals,  but 
in  analog-treated  animals  it  is  not  necessary  that  thiamine-PP  levels  be 
reduced  to  the  same  degree  to  obtain  the  same  functional  disturbances. 


534  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Finally,  the  arguments  of  Gubler  (1961),  that  pyrithiamine  must  induce 
disturbances  other  than  in  o;-keto  acid  metabolism,  I  feel  are  not  entirely 
valid.  He  states  that  since  the  a-keto  acid  oxidase  activities  are  appreciably 
lower  in  thiamine-deprived  rat  livers  than  in  the  livers  of  analog-treated 
rats,  some  other  disturbances  in  physiological  function  must  contribute  to 
the  deficiency  symptoms  and  death.  However,  it  is  unlikely  that  the  changes 
in  liver  metabolism  have  much  to  do  with  either  the  symptoms  or  death, 
and  he  actually  found  that  the  a-keto  acid  oxidase  activities  in  the  brain 
are  lower  in  pyrithiamine-treated  rats  than  in  thiamine-deprived  rats.  The 
other  argument,  that  pyrithiamine  causes  a  polyneuritis  that  is  difficult 
or  impossible  to  reverse  by  administration  of  thiamine  whereas  a-keto  acid 
oxidase  activities  can  be  readily  restored  in  tissue  extracts  by  adding  thia- 
mine-PP,  may  be  significant,  but  these  data  could  be  just  as  easily  explained 
by  an  inhibition  of  thiamine  kinase  (preventing  the  synthesis  of  thiamine- 
PP  in  the  animal)  or  a  very  slow  rate  of  exchange  between  thiamine  and 
enzyme-bound  pyrithiamine-PP  in  the  intact  nervous  tissue. 


ANALOGS  OF  RIBOFLAVIN  AND  FAD 

Riboflavin  functions  in  metabolism  as  riboflavin-5'-phosphate  (flavin  mo- 
nonucleotide, FMN)  and  flavin-adenine  dinucleotide  (FAD)  in  various  oxi- 
dizing enzymes  and  electron  transport.  The  flavin  coenzymes  are  usually 
very  tightly  bound  to  their  respective  apoenzymes  and  are  not  dissociated 
during  extraction  of  the  enzyme  preparations.  Indeed,  in  some  cases,  such 
as  succinate  oxidase,  the  flavin  component  can  be  liberated  only  by  pro- 
teolytic digestion,  with  fragments  of  peptides  attached,  and  the  activity 
cannot  be  restored  by  addition  of  any  flavin  compound.  In  most  cases  it 
is  thus  difficult  for  analogs  to  replace  or  compete  with  the  flavin  coenzyme, 
particularly  in  preparations  from  animal  tissues,  although  in  microorganisms 
the  flavoenzymes  are  generally  more  readily  dissociable.  The  binding  of 
FAD  seems  to  involve  the  isoalloxazine  ring  (perhaps  the  imino  group  at 
position  3),  possibly  the  ribityl  portion,  the  phosphates,  and  the  adenine 
ring.  Chelation  to  apoenzyme-bound  metal  ions,  such  as  iron,  is  likely  be- 
cause most  flavoenzymes  contain  such  metal  ions,  but  there  is  still  some 
doubt  as  to  whether  the  metal  ions  function  primarily  in  binding  or  in  elec- 
tron transfer. 

Animals  and  a  few  bacteria  depend  on  exogenous  riboflavin  but  it  is 
synthesized  in  plants  and  most  microorganisms.  The  pathway  of  riboflavin 
biosynthesis  is  not  well  understood  and  has  been  studied  mainly  in  a  few 
microorganisms  used  for  the  commercial  production  of  riboflavin;  the  reac- 
tions by  which  riboflavin  is  transformed  into  active  coenzymes  are  better 
documented.  An  abbreviated  scheme  of  biosynthesis  and  breakdown  is  re- 
presented here  to  facilitate  discussion  of  the  actions  of  analogs. 


ANALOGS  OF  RIBOFRAVIN  AND  FAD 
Diaminouracil 


6,7-dimethyl-8-fo-l'-ribityl)-lumazine 


535 


Riboflavin 


(3) 


lumichrome  +  ribito' 


riboflavin- 5 '-P  +  AMP 


(1) 

r 

riboflavin- 5 '-P 


(4) 


^   riboflavin  +  P 


(1)  flavo kinase 

(2)  FAD  pyrophosphorylase 

(3)  riboflavinase 


(4)  phosphomonoesterase 

(5)  nucleotide  pyrophosphatase 


(Reaction  (6)  represents  a  possible  formation  of  FAD  directly  from  riboflavin  and 
ATP,  as  postulated  by  Masuda  (1955)  in  E.  ashbyii;  reactions  (1)  and  (2)  also  require 
ATP) 


Many  analogs,  in  which  various  parts  of  the  riboflavin  or  FAD  structure 
have  been  modified,  have  been  examined  and  very  few  are  able  to  replace 
riboflavin,  possibly  due  both  to  inability  to  form  the  coenzyme  analogs  and 
to  relative  inactivity  of  the  coenzyme  analogs  when  formed.  Most  of  these 
analogs  are  not  significantly  inhibitory  to  riboflavin  function,  indicating 
the  high  degree  of  specificity  in  these  reactions.  Beinert  (1960)  has  classified 
the  structural  modifications  producing  most  of  the  interesting  analogs  into 
those  where  (1)  the  riboflavin  is  altered  (as  by  replacement  of  ribitol  by 
other  sugar  alcohols  or  alkyl  groups,  replacement  of  the  isoalloxazine  ring 
by  other  ring  systems,  or  substitutions  in  the  isoalloxazine  ring),  (2)  ribo- 
flavin esters  other  than  the  5'-phosphate  are  synthesized  (as  the  diphos- 
phate, acetylphosphate,  or  sulfate),  and  (3)  dinucleotides  of  riboflavin 
have  the  AMP  replaced  by  other  nucleotides  (such  as  GMP,  IMP,  CMP, 
or  UMP). 

In  this  chapter  we  shall  be  concerned  with  the  rather  simple  and  obvious 
analogs  of  riboflavin  and  FAD.  There  are  several  types  of  drug  and  me- 
tabolic inhibitor  that  probably  interfere  in  one  way  or  another  with  ri- 
boflavin metabolism  or  function  —  such  as  the  promazines,  the  antibiotic 
tetracyclines,  the  acridine  antiseptics,  certain  antimalarials,  and  perhaps 
some  of  the  chelators  of  the  phenanthroline  and  bipyridine  types  —  and 
these  will  be  discussed  for  the  most  part  in  future  chapters.  Quinacrine 
only  will  be  treated  in  some  detail  at  the  end  of  this  section  because  it  is 
the  best  studied  riboflavin  analog. 


536 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


HX 


HX 


Ribityl 


H,C 


Ribityl- 5^P04 


Riboflavin 


o 

Riboflavin- 5 '-P 


HX 


HX 


Dulcityl 

1 


Galactoflavin 


H3C 
HX 


Lyxityl 


NH 


O 


Lyxoflavin 


HX 


HX 


HX 


HX 


CH2CH2OCOCH3 


NH 


Lumiflavin 


U-2112 


H3C 
HX 


CH,CHXCOCHXH,C(X) 


^^^■> 


U-6538 


NH 


HX 


Ribityl 

I 
N. 


NH 
CH3  O 

Isoriboflavin 


ANALOGS  OF  KIBOFLAVIN  AND  FAD 


537 


Ribityl 


H,CCH. 


H3CCH; 


6,  7 -Diethyl  analog 
of  riboflavin 


Ribityl 


6-Chloro  analog 
of  riboflavin 


Ribityl 

I 


H,C 


A^. 


H  NH, 


2,  4-Diamino-7,  8- 

dimethyl-10- 

ribityl-5,  lO-dihydrophenazine 


Sorbityl 


Flavotin 


H,N 


NH, 


H,N 


Acriflavine 
(trypaflavine,  euflavine) 


NH, 


Proflavine 


CH, 


N-^    ^ 

II 

0 

Toxoflavin 

Effects  on 

Growth 

A  number  of  analogs  have  been  found  to  inhibit  the  growth  of  Lacto- 
bacillus casei,  the  standard  test  organism:  isoriboflavin,  the  6-chloro  and  7- 
chloro  analogs  of  riboflavin  (Lambooy,  1955),  the  2,4-diaminophenazine 
analog  (Woolley,  1944  b),  and  riboflavin-5'-sulfate  (Egami  et  al.,  1956). 
Other  bacteria  are  occasionally  inhibited,  for  example  S.  aureus  and  Str. 
plantarum  by  dichlororiboflavin  (Lambooy,  1955).  The  acridines,  proflavine 
and  5-aminoacridine,  suppress  the  growth  of  L.  casei  but  this  is  not  coun- 
teracted by  increasing  the  riboflavin  concentration,  so  the  mechanism  is 


538  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

not  clear  (Madinaveitia,  1946).  Toxoflavin  is  a  poisonous  substance  from 
Pseudomonas  cocovenenans  and  responsible  for  some  fatal  food  poisonings  in 
Java.  It  quite  potently  inhibits  the  growth  of  E.  coli,  S.  aureus,  B.  subtilis, 
and  Shigella,  as  well  as  being  toxic  to  experimental  animals  (Latuasan  and 
Berends,  1961).  The  mechanism  is  unknown  but  it  was  postulated  that  tox- 
oflavin is  an  effective  electron-acceptor  and  removes  electrons  from  the 
transport  chains  to  form  hydrogen  peroxide,  possibly  competing  with  the 
natural  flavin  coenzymes  (or  accepting  electrons  from  them).  Relatively 
little  has  been  done  on  the  induction  of  riboflavin  deficiency  in  animals. 
Isoriboflavin  almost  stops  the  growth  of  young  rats  at  2  mg/day  and  ribo- 
flavin is  able  to  reverse  this  (Emerson  and  Tishler,  1944),  and  galactoflavin 
appears  to  produce  very  similar  effects  (Emerson  et  al.,  1945).  The  6-chloro 
and  7-chloro  analogs  of  riboflavin  (Lambooy,  1955)  and  the  dinitrophena- 
zine  analog  (Woolley,  1944  b)  produce  mild  deficiency  states  in  rats  and 
mice.  Effects  have  usually  been  determined  by  growth  rates,  and  typical 
symptoms  of  riboflavin  deficiency  have  seldom  been  noted,  mainly  because 
it  requires  a  fairly  long  time  to  deplete  the  tissues  of  enzyme-bound  co- 
enzymes. 

Some  interesting  and  suggestive  reports  on  the  carcinostatic  activity  of 
certain  analogs  have  appeared.  The  6,7-dichloro-9-(r-D-sorbityl)isoalloxa- 
zine  analog  causes  regression  of  mouse  lymphosarcoma,  the  ribityl  compound 
has  slight  activity,  and  the  other  sugar  alcohol  derivatives  are  inactive 
(Holly  et  al.,  1950).  Replacement  of  the  ribityl  group  with  nonsugar  residues 
gives  riboflavin  antagonists  which  inhibit  L.  casei  and  produce  deficiency 
states  in  rats.  The  9-hydroxyethyl  analog  of  riboflavin  (U-2113)  weakly 
suppresses  mouse  adenocarcinoma  (Shapiro  et  al.,  1956),  and  the  9-aceto- 
xyethyl  analog  (11-2112)  behaves  similarly  (Lane  et  al.,  1958).  However, 
U-2112  given  to  patients  with  various  types  of  cancer  (0.25-6  g/day  for 
5-84  days,  with  total  doses  1.25-226  g)  exhibits  no  beneficial  action  and 
no  evidence  of  riboflavin  deficiency  is  seen;  this  was  attributed  to  the  rapid 
hydrolysis  of  this  substance  in  man.  The  9-hemisuccinoxyethyl  analog  (U- 
6538)  does  not  inhibit  L.  casei  but  depresses  growth  in  rats  at  10  mg/kg/ 
day,  which  is  reversible  with  riboflavin  (Lane  et  al.,  1959).  It  is  quite  ef- 
fective against  lymphosarcoma  in  rats  and  at  5  mg/kg/day  leads  to  a  66% 
inhibition  of  tumor  growth  without  depressing  the  over-all  growth  rate. 
Two  of  four  patients  receiving  the  compound  showed  changes  suggestive 
of  riboflavin  deficiency  but  no  evidence  of  carcinostasis  was  observed,  per- 
haps due  to  the  terminal  nature  of  the  disease  and  the  metabolism  of  the 
analog.  Galactoflavin  is  known  to  cause  tumor  regression  in  rodents.  It  is 
tolerated  by  patients  at  a  dose  of  1  g  every  8  hr  for  2-5  months,  and  de- 
ficiency symptoms  do  not  occur  unless  the  diet  is  low  in  riboflavin  (Lane 
and  Brindley,  1964).  Presumably  reports  on  its  carcinostatic  activity  will 
be  published. 


ANALOGS  OF  RIBOFLAVIN  AND  FAD  539 

Metabolism   of   Riboflavin    Analogs   and    Effects   on    Riboflavin    Metabolism 

Certain  analogs,  such  as  the  6,7-diethyl  derivative,  are  able  to  replace 
riboflavin  to  some  extent  at  low  concentration  but  are  inhibitory  at  higher 
concentration.  This  analog  supports  the  growth  of  L.  casei,  and  Lambooy 
(1950)  believed  that  it  must  be  phosphorylated.  This  was  demonstrated  in 
the  rat  where  6,7-diethylriboflavin-5'-P  was  found  in  the  liver  although  no 
FAD  analog  was  demonstrable  (Aposhian  and  Lambooy,  1955).  L.  lactis  is 
able  to  incorporate  lyxoflavin  into  lyxoflavin-5'-P  and  the  corresponding 
dinucleotide  (Huennekens  et  al.,  1957  b).  Scala  and  Lambooy  (1958)  were 
able  to  modify  L.  casei  by  prolonged  riboflavin  deficiency  and  high  con- 
centrations of  the  6-chloro  analog  so  that  the  organism  could  use  either  ribo- 
flavin or  the  analog.  They  believe  that  the  analog  inhibits  the  phosphor- 
ylation of  riboflavin.  It  is  interesting  that  the  adapted  organism  cannot 
use  the  7-chloro  analog. 

Flavokinase  catalyzes  the  phosphorylation  of  riboflavin  and  shows  a  high 
degree  of  specificity  toward  substrates.  The  yeast  enzyme  phosphorylates 
dichlororiboflavin  as  well  as  riboflavin,  arabitylflavin  poorly,  and  all  other 
analogs  tested  not  at  all  (including  isoriboflavin,  galactoflavin,  dulcitylfla- 
vin,  and  sorbitylflavin)  (Kearney,  1952).  The  only  analog  that  inhibits  the 
enzyme  is  lumiflavin  (35%  at  0.18  mM  with  riboflavin  0.051  mM)  and  this 
occurs  only  when  the  analog  is  in  excess  of  the  riboflavin.  McCormick  (1962) 
has  extended  this  work  to  partially  purified  rat  liver  flavokinase  and  found 
similar  behavior,  only  riboflavin,  dichlororiboflavin,  and  arabitylflavin  be- 
ing phosphorylated  (all  with  KJs  between  0.012  and  0.017  mM).  Four 
analogs  were  found  to  be  inhibitory:  lumichrome  {K^  =  0.048  mM),  lumi- 
flavin {K^  =  0.031  mM),  the  9-formylmethyl  analog  {K^  =  0.0097  mM), 
and  the  9-(2'-hydroxyethyl)  analog  {K^  =  0.0068  mM).  The  following  are 
not  phosphorylated  and  do  not  inhibit:  isoriboflavin,  galactoflavin,  sorbityl- 
flavin, dichloroarabitylflavin,  7-methylmannitylflavin,  and  7-methyldulci- 
tyLflavin.  The  fact  that  most  analogs  are  not  attacked  by  flavokinase  is  per- 
haps the  primary  reason  for  the  failure  of  these  compounds  to  replace  ribo- 
flavin. It  is  also  clear  that  the  data  are  insufficient  to  draw  conclusions 
relative  to  the  possibility  of  some  of  the  most  commonly  used  analogs  inhi- 
biting the  phosphorylation  of  riboflavin,  but  what  evidence  we  have  would 
indicate  that  such  inhibition  is  unlikely  to  be  important.  The  synthesis  of 
riboflavin  from  6,7-dimethyl-8-(r-D-ribityl)lumazine  by  an  enzyme  system 
from  Ashbya  gossypii  is  potently  inhibited  by  a  variety  of  analogs  of  this 
precursor,  of  which  the  6,7-dihydroxy  derivative  is  the  most  active  {K^  = 
0.000009  mM)  (Winestock  et  al,  1963).  It  is  interesting  that  5'-deoxyribo- 
flavin  is  fairly  inhibitory  {K,  =  0.019  mM)  and,  indeed,  it  was  concluded 
that  the  sugar  moiety  is  necessary  for  inhibition. 

Very  little  information  on  the  effects  of  analogs  on  the  tissue  levels  of 
riboflavin  compounds  is  available.  Rats  given  galactoflavin  for  10-28  days 


540  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

show  a  70-75%  depression  of  liver  mitochondrial  flavin,  and  dietary  ribo- 
flavin restriction  also  reduced  the  level  (Beyer  et  al.,  1961).  The  evidence 
from  depression  of  enzyme  activity  will  be  discussed  later.  The  thorough 
analysis  of  the  effects  of  riboflavin  deficiency  on  rats  by  Burch  et  al.  (1956) 
has  shown  that  various  tissues  differ  markedly  in  ability  to  retain  FMN 
and  FAD  (see  accompanying  tabulation).  Deficiency  of  23-day  duration 


%   Change  in  deficient  rats 

Tissue 

FMN 

FAD 

NADH  oxidase 

D-Amino  acid 
oxidase 

Xanthine 
oxidase 

Brain 

-24 

-19 

0 

+  8 

Liver 

-86 

-61 

+  10 

-66 

-22 

Kidney 

-41 

-19 

-11 

-17 

-19 

Heart 

0 

-32 

2.3 

1    (•) 

0 

has  little  effect  on  brain  flavins  while  liver  levels  drop  rapidly.  No  necessary 
correlation  between  total  FMN  or  FAD  and  enzyme  activity  is  evident, 
indicating  that  some  enzymes  will  lose  their  FMN  or  FAD  much  more 
readily  than  others  and  that  a  fraction  of  the  cellular  flavin  may  be  nonen- 
zymically  bound.  On  the  basis  of  these  results,  one  might  expect  riboflavin 
analogs,  when  active,  to  exert  differential  effects  on  the  various  flavoen- 
zymes,  and  it  is  likely  that  analyses  for  total  flavins  will  not  provide  me- 
tabolically  significant  figures. 

Effects  on  Flavoenzymes 

The  enzymes  most  commonly  used  to  test  the  inhibitory  activity  of 
riboflavin  analogs  are  those  with  dissociable  flavin  coenzymes,  such  as  the 
old  yellow  enzyme  and  D-amino  acid  oxidase,  and  in  such  cases  an  inhibi- 
tion of  a  competitive  nature  is  not  unexpected.  However,  there  are  many 
instances  of  the  inhibition  of  enzymes  which  have  very  tightly  bound  co- 
enzymes and  these  are  more  difficult  to  interpret.  Some  inhibitions  of  both 
types  are  shown  in  Table  2-33.  Inhibitions  by  riboflavin,  FMN,  and  FAD 
are  also  included  because  these  show  that  it  is  not  alwaj^s  necessary  to 
consider  inhibition  as  resulting  from  structural  analogs. 

Most  of  these  inhibitions  appear  to  be  noncompetitive.  Thus  the  inhibi- 
tions of  L-galactono-y-lactone  dehydrogenase  by  riboflavin,  L-amino  acid 
oxidase  by  riboflavin  and  its  analogs,  and  D-amino  acid  oxidase  by  ribo- 
flavin are  not  reduced  by  increasing  concentrations  of  FMN  or  FAD.  How- 
ever, the  inhibitions  of  kidney  D-amino  acid  oxidase  by  FMN  and  ribo- 
flavin-5' -sulfate  are  competitive  with  respect  to  FAD,  and  the  inhibition 


ANALOGS  OF  RIBOFLAVIN  AND  FAD 


541 


K 


o 


W 


^^^ 

Ti 

■ 

c 

cS 

a 

-*J 

bX 

o 

S 

M 

CO 

^    t<; 


PS 


•*     O     -*     Ci 


O    — '     o    o 


cS 

c3 

i^ 

c3 

*    ^ 

■t: 

c3 

« 

o    ^^ 

o 

u^ 

O 

^ 

:a  s 

rO 

"3 

-Q 

§ 

Ci  fe 

S 

M 

S 

fe 

K         ^ 


cr; 


^iS       Q  <: 


T3 


542 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


P5 


w 


W    W 


73 

-^ 

C 

e6 

S 

tH 

tH 

<D 

<» 

M 

tiC 

_g 

c 

CC 

CB 

cs     -r     03 


^        O     -r 


hH       Ph      I— I 


O     Q 


O 


^    ^ 


A 


m 


—     CX) 

^      C£) 

^ 

C5 

o 

-*     O 

TjH 

lO 

o 

-H     <N     lO 

odd 


A        A 


O 


o 


fq 


O 

o    ^ 

^ 

-Q     '^ 

•-     (^ 

S 

P5     P^ 

s   s 


w 


P5 


&►, 

C 

<D 

o 

^ 

o 

CO 

03 

o 

^ 

oj 

cS 

c* 

^ 

o 

s 

^ 

ft 

O 

-5 

O 

'2 

■ft 

6 
c 
o 

bJO 

2 

C5J 

o 

& 

'3 

T3 

O 
O 

Q 

>^. 

^ 

4^ 

•Jl 

O 

"S 
-d 

O 

Hi 

hJ 

fR 

J 

0 

5 

O 

o 


t3 
C 

^3 

« 

T3 

-a 

fl 

.^^ 

C 

c 

c3 

1) 

(S 

c3 

fi 

>> 

<A 

>> 

>> 

o 

0) 

^ 

<U 

aj 

to 

ft 

3 

3 

3 

pq 


^ 


ANALOGS  OF  EIBOFLAVIN  AND  FAD 


543 


IC 

05 

00 

05 

(-• 

1— 1 

^-^ 

*« 

« 

e 

o 

^ 

.4^ 

1 

^ 

CO 

IS 

C 

]B 

3 

6 

O 

v 

s 

w 

< 

.K 

cS 

o 

c 

o 

cS 

^ 

o 

^ 

^ 

^ , 

Ol 

^5 

^ 

(^ 

o 

<s  o 

d 

eo  05 

fe< 


w 


r::r  o  rzr 


o   o 


C<l    <M    (M 

odd 


■*     GO     Oi 
CO    (M    -H    d 


2   ^ 


« 


Pi 


cs         j:  -^ 


K 


2  a>  S 
'^  Ti  ^ 


0)     -tJ 

Cm     '^^ 


P5  fe 


'2 

o  ^  Q 

o  ?;    c«    > 

3 

:e  S  <^ 

:a  §  -d  ^ 

J 

P5  f!^   P=H 

«  fe  O  fe 

.5>  i 


c    5- 


K3 


.??     S 


g 

^ 

^ 

^ 

s 

o 

'bib 

73 

o 

o 

"tJ 

s 

;-i 

0) 

O 

"x 

~ 

33 

-c 

o 

c3 

£ 

c 

eo 

O 
3 
t3 

"3 

IS 

^ 

'o 

ffi 

£ 

_C 

^ 

P-i 

O 

"o 

fi 

r2 

03 

< 

'x 
o 

o 

to 
1 

'2 

ID 
O 

O 

C     o 

>  2 

J 

c3      M 

eS 

P5 


«    52 


P5 


544  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

of  glutamate  racemase  by  riboflavin  is  reduced  by  FAD.  It  is  clear  that 
there  is  not  much  information  on  the  inhibition  of  enzymes  by  riboflavin 
analogs,  especially  by  their  phosphates,  or  FAD  analogs.  The  mechanism 
for  the  inhibition  of  enzymes  in  which  the  flavin  coenzyme  is  tightly  bound 
is  not  known.  However,  it  is  possible  to  suggest  three  mechanisms.  It  is 
now  known  that  various  flavins  and  their  nucleotides  form  molecular  com- 
plexes with  one  another,  and  the  formation  of  such  complexes  with  bound 
FMN  or  FAD  may  occur,  preventing  the  normal  interactions  of  the  coen- 
zymes in  oxidation.  In  some  instances  the  inhibitors  may  interfere  with 
the  experimental  electron  acceptor,  particularly  when  this  is  a  dye.  Lastly, 
one  must  consider  the  possibility  of  nonspecific  binding  of  these  polyhetero- 
cyclic  compounds  to  the  enzymes;  one  might  predict  that  a  number  of  en- 
zymes not  involving  flavins  would  be  inhibited  by  such  analogs,  but  few 
have  been  examined.  None  of  the  riboflavin  analogs  in  Table  2-33  is  a  po- 
tent inhibitor  and  it  is  unlikely  that  these  inhibitions  are  responsible  for 
any  of  the  in  vivo  effects  observed. 

The  coenzyme  of  the  old  yellow  enzyme  is  FMN,  and  riboflavin-5'-sulfate 
does  not  interfere  with  its  binding  to  the  apoenzyme,  which  Theorell  et  al. 
(1957)  explain  by  the  less  negative  charge  on  the  sulfate  group.  Riboflavin- 
5'-sulfate,  however,  inhibits  D-amino  acid  oxidase  (Egami  and  Yagi,  1956), 
so  that  the  structural  requirements  for  binding  must  be  different  in  these 
two  enzymes.  Yagi  and  Nagatsu  (1960)  have  studied  the  effects  of  ribofla- 
vin-5'-sulfate  on  rat  liver  mitochondrial  oxidations  of  a-ketoglutarate,  suc- 
cinate, malate,  and  D-alanine,  and  found  that  no  inhibition  is  exerted  at 
0.1  mM,  which  they  interpret  as  due  to  the  tight  binding  of  the  FAD  in 
the  mitochondria.  Aged  mitochondria  are  stimulated  by  FAD  and  here  in- 
hibition by  riboflavin-5' -sulfate  can  be  demonstrated.  The  FAD  analogs  of 
the  various  flavins  have  not  been  studied  often  but  Huennekens  et  al. 
(1957  b)  found  lyxoflavin-5'-phosphate  to  be  active  in  the  NADPH-cyto- 
chrome  c  reductase  (although  less  than  FMN)  and  lyxoflavin  dinucleotide 
to  be  active  in  the  D-amino  acid  oxidase  (but  less  than  FAD).  We  have 
seen  that  riboflavin  deficiency  leads  to  reduction  in  the  activities  of  certain 
enzymes  in  the  tissues.  Administration  of  galactoflavin  to  rats  for  15-28 
days  leads  to  an  approximately  40%  reduction  in  glutamate  and  /5-hy- 
droxybutyrate  oxidation  in  liver  mitochondria,  but  no  change  in  succinate 
oxidation  or  in  the  P  :  0  ratios  (Beyer  et  al,  1961).  U-2113,  the  9-hydrox- 
yethyl  analog  of  riboflavin,  causes  a  slight  (15%)  decrease  in  tumor  xan- 
thine oxidase  in  mice.  It  is  not  known  if  this  is  due  to  FAD  depletion  or  a 
more  direct  inhibition.  5-Hydroxytryptamine  (serotonin)  is  metabolized  by 
monoamine  oxidase  to  5-hydroxyindoleacetate;  both  riboflavin  deficiency 
and  galactoflavin  increase  the  urinary  excretion  of  this  product,  indicating 
that  one  of  the  other  metabolic  pathways  for  serotonin  is  depressed  by 
interference  with  flavin  function  (Wiseman  and  Sourkes,  1961).  These  mis- 


ANALOGS  OF  RIBOFLAVIN  AND  FAD  545 

cellaneous  observations  do  not  provide  a  satisfactory  basis  for  understand- 
ing the  metabolic  effects  of  riboflavin  analogs. 

We  have  been  discussing  analogs  of  the  riboflavin  portion  of  FAD  and 
some  mention  of  the  adenine  nucleotides  as  inhibitors  should  be  made.  The 
D-amino  acid  oxidase  of  sheep  kidney  is  inhibited  competitively  by  various 
purines  and  nucleotides  (see  accompanying  tabulation)  (Burton,  1951  a). 


Inhibitor 

{mM) 

5'-AMP 

1.05 

S'-AMP 

No  inhibition 

ADP 

13 

ATP 

11 

Adenosine 

45 

Adenine 

22 

Hypoxanthine 

24 

Caffeine 

11 

It  was  shown  that  complexes  between  riboflavin  and  purines  are  formed 
and  have  the  following  dissociation  constants:  caffeine  10  n\M,  adenosine 
30  mM,  AMP  40  mM,  ADP  37  mM,  and  ATP  39  mM.  The  formation  of 
such  complexes  might  account  for  the  enzyme  inhibition  in  the  case  of 
adenosine  and  caffeine,  but  cannot  account  for  the  more  potent  effects  of 
AMP  and  ADP,  these  latter  substances  competing  with  FAD  for  the  apo- 
enzyme  site.  The  D-amino  acid  oxidase  from  pig  kidney  is  likewise  inhibited: 
50%  inhibition  is  given  by  0.4-0.6  mM  AMP,  ADP,  ATP,  and  IMP;  by 
6  mM  adenine,  adenosine,  and  hypoxanthine;  and  by  15-20  mM  uracil,  cyto- 
sine,  and  ribose-5'-P  (FAD  0.00014  mM  in  all  cases)  (Walaas  and  Walaas, 
1956).  The  K,  for  the  competitive  5'-AMP  is  0.64  mM.  Crandall  (1959) 
determined  K^  for  AMP  as  0.1  mM  for  this  enzyme.  Flavokinase  is  inhi- 
bited strongly  by  5'- AMP  {K^  =  0.025  mM),  and  it  is  possible  that  the 
adenine  portion  of  the  nucleotide  competes  with  the  alloxazine  ring  of  ribo- 
flavin for  the  active  center  (Kearney,  1955). 

Quinacrine  (Mepacrine,  Atabrine,  Atebrin) 

Quinacrine  is  an  acridine  derivative  introduced  by  the  Germans  in  1932 
for  malaria  therapy  as  a  suppressive  agent  and  is  more  effective  than  qui- 
nine on  the  asexual  forms  of  the  plasmodia.  Since  the  observation  by  Wright 
and  Sabine  (1944)  that  FAD  counteracts  the  inhibition  of  tissue  respiration 
by  quinacrine,  it  has  been  commonly  assumed  that  quinacrine  exerts  its 


546  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

primary  effects  by  interference  with  flavoenzymes,  and  has  come  to  be  the 
most  widely  used  substance  to  detect  the  participation  of  a  flavin  com- 
ponent in  an  enzyme  or  metabolic  system.  Quinacrine  not  only  inhibits  the 


CH,CH,CH3-N^^'^' 


CH,0 


Quinacrine 

malarial  organism  but,  in  common  with  other  acridines,  suppresses  the 
growth  of  various  bacteria.  For  example,  Lactobacillus  casei  is  inhibited 
and  this  seems  to  be  related  to  flavin  metabolism  since  the  maximal  con- 
centration at  which  growth  occurs  is  0.12  niM  when  riboflavin  is  0.00066  mM 
and  0.49  mM  when  riboflavin  is  10  times  the  previous  concentration  (Ma- 
dinaveitia,  1946).  Spore  germination  of  B.  subtilis  (Falcone  et  at.,  1959)  and 
B.  coagvlans  (Amaha  and  Nakahara,  1959)  induced  by  L-alanine  is  inhibited 
58%  by  0.1  mM  quinacrine  and  nearly  completely  by  1  mM. 

The  toxic  efl^ects  observed  in  experimental  animals  and  man  —  for  ex- 
ample, gastrointestinal  (abdominal  pain,  diarrhea,  nausea),  dermatological 
(eczematoid  dermatitis,  lichen  planus),  central  nervous  system  (psychoses), 
and  others  —  do  not  appear  to  be  related  to  riboflavin  deficiency,  and  the 
typical  syndrome  of  deficiency  has  never  been  produced  by  quinacrine. 
Thus  one  must  assume  that  other  actions  are  probably  of  more  importance 
in  animals.  One  characteristic  of  quinacrine  is  its  remarkable  ability  to  be 
accumulated  in  the  tissues  during  chronic  administration,  and  eventually 
the  tissue  levels  are  hundreds  or  thousands  of  time  higher  than  in  the  serum 
(see  Table  1-8-1).  This  is  evident  from  the  yellow  coloration  of  the  tissues. 
These  levels,  of  course,  do  not  represent  free  quinacrine  and  the  accumula- 
tion is  due  to  the  high  affinity  of  various  tissue  components  for  quinacrine. 
It  is  bound  in  the  cytoplasm,  the  mitochondria,  and  the  nucleus;  equilibra- 
tion of  isolated  nuclei  with  quinacrine  results  in  a  200-fold  concentration 
differential  (Reiner  and  Gellhorn,  1956).  Proteins,  nucleic  acids,  and  nu- 
cleoproteins  bind  quinacrine  strongly,  and  some  of  the  inhibitory,  muta- 
genic, and  carcinostatic  effects  have  been  attributed  to  such  binding.  It 
is  thus  clear  that  quinacrine  can  be  bound  at  many  loci  in  the  cell. 

Effects  of  Quinacrine  on   Enzymes 

Many  enzymes  are  inhibited  by  quinacrine  (Table  2-34).  In  some  cases 
the  inhibition  is  competitive  (or  at  least  reduced  by  increasing  the  concen- 
tration of  FMN  or  FAD)  and  in  others  it  is  not.  Quinacrine  has  come  to 


ANALOGS  OF  RIBOFLAVIN  AND  FAD  547 

be  the  most  commonly  used  detector  for  the  participation  of  flavins  in 
enzyme  systems,  but  before  this  is  subjected  to  analysis  we  shall  discuss 
some  of  the  results  which  have  been  obtained.  The  following  summary  pre- 
sents the  data  reported  on  antagonism  but  does  not  in  any  case  imply  a 
truly  competitive  mechanism. 

Enzymes  in  which  inhibition  is  reduced  by  FMN  or  FAD 

Adenosinetriphosphatase  (mitochondrial  NAD-activated):  FMN  and  FAD 
(Low,  1959) 

Aldehyde  oxidase  (rat  and  monkey  liver):  FAD  (Lakshmanan  et  al.,  1964; 
Mahadevan  et  al,  1962) 

Aliesterase  (liver):  FMN  (Hemker  and  Hiilsman,  1960) 

D- Amino  acid  oxidase  (lamb  and  sheep  kidney):  FAD  (Hellerman  et  al., 
1946;  Burton.  1951  a) 

Catechol  oxidase  (spinach):  FAD  (Nair  and  Vining,  1964) 

Choline  dehydrogenase  (rat  liver):  FAD  (Bargoni,  1963) 

Cytochrome  reductase:  FMN  (Haas,  1944) 

Hydroxylamine:  cytochrome  c  oxidoreductase  {Nitrosomonas):  FMN  and 
FAD  (Aleem  and  Lees,  1963) 

Lactate  dehydrogenase  {Lactobacillus  and  yeast):  FMN  and  FAD  (Snos- 
well,  1959;  Iwatsubo  and  Labeyrie,  1962) 

NADH:  nitrite  oxidoreductase  (Neurospora):  FAD  (Nicholas  et  al.,  1960) 

NADH  oxidase  {Azotobacter,  Clostridium,  and  Lactobacillus):  FAD  (Re- 
paske  and  Josten,  1958;  Dolin,  1959;  C.  F.  Strittmatter,  1959) 

Nitrate  reductase  (Pseudomonas):  FAD  (Fewson  and  Nicholas,  1961) 

Succinate  oxidase  {Tetrahyrnena  and  Xanthomonas):  FMN  and  FAD  (Ei- 
chel,  1956  b;  Madsen,  1960) 

Enzymes  in  which  inhibition  is  not  reduced  by  FMA  or  FAD 

Adenosinetriphosphatase  (mitochondrial  sonicate)  (Beyer,  1960) 

Allohydroxy-D-proline  oxidase  {Pseudomonis)  (Yoneya  and  Adams,  1961) 

L-Amino  acid  oxidase  (moccasin  venom)  (Singer  and  Kearney,  1950) 

Ethanolamine  oxidase  (Arthrobacter)  (Narrod  and  Jakoby,  1964) 

L-Galactono-5^-lactone   dehydrogenase    (cauliflower)   (Mapson  and  Bres- 
low,  1958) 

Lactate  dehydrogenase  (Propionibacteritim)  (Molinari  and  Lara.  1960) 

Nitrate  reductase  {E.  coli)  (Heredia  and  Medina,  1960) 

Old  yellow  enzyme  (yeast)  (Kistner,  1960) 


548 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


is 


« 

P- 

s 

g 

>H 

o 

N 

o 

!Z 

W 

c 

_o 

P^ 

'-3 

o 

c 

cS 

:z; 

0) 

o 

Ph 

H 

m 

W 

^ 

a 


CD     J3 


e 

-d 

fl 

ITt 

03 

^ 

(S 

CO 

a 

^ 

jJ3  I— H 


^^       M 


?o  ^- 


05      05 


s  o 


s  ^  o 


^    c 


O   ^ 


S>  ^ 


u     .3 


c    ft 


S    a 


Ph 


o5  ::i 


;ta         oi 


^    o    o   1^2 

^  t^  w  ^ 


^    »o    O    C<1 


Tj*     «^    00     Ol 


Ph    Ph 

Ph 

Ph 

Ph 

Ph 

e^  pli  e^ 

CM 

H    H 

H 

H 

H 

H 

H    b^    b^ 

H 

<1    <3 

< 

<3 

<1 

<i5 

<    <    < 

< 

T3     T3 
O      O 


s  s 


CO      a3      aj      r- 


P5     P5 


N 


pIH 


O    ^    S^l     <M 
S    S    <Z>     O 


P 


W 


-c  -^  -^  V.  ts 

-S  -2  -2  «  ^ 

eS  eS  cj  o  d 

P5  P5  pq  W  >^ 


w 


ANALOGS  OF  RIBOFLAVIN  AND  FAD  549 


el 

OS 

eS 

f-H 

K 

'O 

"e 

c 

e3 

« 

iH 

c 

eg 

^ 

"S 

13 

P< 

^_^ 

ft 

O 

-* 

o 

M 

CD 

bJD 

_o3 

OS 

2, 

'c? 

'5° 

P5 

P5 

C5      ;2r 


-O 

<» 

_c 

,_! 

W 

, — . 

J3 

lO 

X! 

05 

^73 

05 

> 

^ 

05 

-d 

c 
o 

c 

^^ 

cS 

;h 

^ 

t-i 

^ 

C 

:0 

'3 

M 

CO 

h^l 

^ 

fc< 


O-hO^  lO  ICtMCOO  o 


o  o  o  o 


O  ^-v  -^ 


73 


c  -so 


^^ 

,,^ 

o 

^ 

-* 

^ 

o 

o 
o 

o 
o 

o 

o 

q 

o 

o 

'c 

o 

d 

o 

_ce 

Q 

"3 

<5 

8- 

ft^ 

hj 

H      <1  ft^  J  <3     Q     O 


S  >r        ^  J     «^  '^  kS?  .2     o      ft     1 


X 

0) 

0 

eg 

-53 

'2 

0 

'S 

s   S  >$,  <  <       ft  -ti 


o 


I  i  i I  I        §     ^  1  ;     I  II  I 


"o 

0 
0 

JU 

J3 

>>  s 

>. 

0 

0 

0 

0 

550 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


a 


O      u 


K 


P 
-S 

O 

"m 

M 

S" 

■* 

t3 

t3 

"* 

C 

C 

C5 

c3 

e3 

S 

73 
g 

C 
O 

CO 

(S 

£^ 

P^ 

eS 

cS 

ce 

ffi 

^ 

S 

K 


W 


O    -^    <M 


•i    - 


o        ts 


o 

« 

>, 

n^ 

01 

s 

o 

c 

p 

cS 

4J 

y 

O 

Tl 

o 

C! 

>> 

c3    ^ 


OS 

o 

05 

c 

^,^ 

^ 

e 

Tl 

so 

b4 

s. 

OJ 

O 

O 

O 

W 

o 


:3 


T*  C^         fl 


o     o 


t-    o 


ei^ 

Ph 

Q 

o 

< 

<: 

^ 

;3 

t3 

'C 

C 

03 

S 

Ph 

'^ 

«p 

o 

<D 

i 

i 

Ol 

m 

CO 

O 

O 

O 

o 

o 

o 

S 

p 

5 

O 

CD 

w 


CO    P 

o 


-S       c3      -^^ 


M 


WW      ffi   c 


o    o  o 

CO     o    o 


lO      O    CO 
O      O    Ci 


^    lO    CO    »o 


O    CD    O    CO 


c     c     c 


03        ^       -:=     -=        -i     -= 


o  o  o 


"O 

fc< 

C3 

"? 

to 
> 

73 

S 

>» 

;3 

s 

&c 

IB 

M 

s 

^ 

■ft 

c 
'2 

'& 

c3 
0) 

^ 

0) 

2 

s 

M 

c 

p 

3 

O 

'3 

s 

O 

Q 

O 

o 


o 


ANALOGS  OF  RIBOFLAVIN  AND  FAD  551 


fe 

^ ^ 

{i^ 

M 

Si 

t3 
c 

05 
C5 

S 

o 


c 


2 

O  S    h5    S      niS  C^ 


Q  2     :b     S      -5  "q 


c 

05 

'd 

-o 

C 

c 

^-^ 

eS 

o 

o 
2 

02 

<s 

Is 

o 
c 

"o 

5 

w 

§ 

iC    o 

o 

o 

o 

-* 

(M 

C^ 

O    CO 

■<t   lo 

CO 

r^ 

1— 1 

5^ 

CO 

ir- 

CO   CO 

o    .S     s^ 


«o    iC  ^  o   ^    »0        1 

OcO-*|Tt<Tt<  ^^,^(_J  OOO'KI         oooooo 


Clh 

Ph 

H 

H 

<: 

<J 

Ti 

Pm 

T3 

c 

H 

C 

c« 

<i1 

eS 

IC 

T3 

lO 

cq 

C 

C^ 

3.  CO 


8  W  c 


ft 

<1 

"^ 

T^ 

G 

a 

lO 

c3 

""— ^ 

-t^ 

P-I 

o 

H 

1 

c 

T3 


C!  CO        CO        ^  ;S      5s         -^ 

O  ><    >H    Q.      Q  ^      ^ 


s 

s 

-w 

H 

r^ 

o 

g 

.^ 

(*. 

05 

>> 

Cl 

2 

M 

^ 

>. 

=  H 

w 

j  ^ 

^ 

_o 

^ 

^ 

^ 

g 

P5 

e 

^ 

o 

e 

i 

o 

eS 

bo 

^ 

o 

-^ 

o 

>, 

43 

ji 

ft 

v 

^  s 


C       r"       '^  S! 

J  i  ^     * 
c  Si,        -s®         &>S«2 

-^  O  L^rt  OOOo 

45  _2  '^       Z. 
'O  +^  •—       j; 


cS 


552 


2.  ANALOGS  OF  ENZYME  KE ACTION  COMPONENTS 


rt 


o  -s 


P5 


« 


M    W 


^         CO  -H 


tS 

-2 

i 

cS 

0) 

■t^ 

•4^ 

o 

c3 

c3 

-i-i 

t-1 

i-:i    1-^ 


12      '^ 


•^ 

-ss 

Si 

a 

fi, 

s 

e 

•i 

s 

^ 

i» 

*4i 

S 

ts 

;»! 

tJS 

tO 

e 

o 

>i  ^ 


a. 


fe 


eo 

Oi 

^ 

oo 

Oi 

CO 

>c 

OS 

^-' 

CO 

05 

^^ 

c 

C5 

~— ' 

c 

S 

o 

^_^ 

t-l 

o 

tc 

OJ 

tc 

2 

w 

C 
c3 

O 

-3 

o 

c 
W 

c 
a 

3 

1 

e 

C 

s 

03 

1— 1 

CZ3 

o 

,M 

dS 

s 

3 

u 

o 

>-, 

(B 

O 

cS 

<?: 

M 

Q 

W 

PlH 

P5 

P5 

CO 

o 

CO 

o 

o 

00 

00 

t- 

CO 

CO 

eo 

w 


w 


-2   ^        WO 

Is  :2     p  :2 


W    W    W    |ij 
P    Q    fi    Q 

^   ^   ^   ^ 


^-    t§    ^'   M 


T3 

c3 

a 

« 

c3 

C 

CO 

a 

4h 

§:^ 

>> 

So 

O 

W    2 

t3 

9  § 

>. 

S^ 

w 

s     s  '^ 


o 


fM 


>> 

o 

2 

U 

^ 

s 

c3 

cS 

cS 

4^ 

o 
3 

-4^ 

3 

-C 

T5 

T3 

« 

?^ 

W 

O 

W 

S 

P 

1^ 

P 

'S 

< 

'x 

<: 

'X 

^ 

o 

^ 

O 

ANALOGS  OF  RIBOFLAVIN  AND  FAD 


553 


P^      Ph 


o 
m 

35 

c 

c3 

'— ' 

^ 

o 

X! 

"e 

^ 

_s 

"cS 

05 

« 

W 

05 

^ 

02 
CO 

"o 

c3 

,c 

s 

[q 

.^ 

o 

O 

"o 

05 

S 

Q 

ft 

o    ^ 


P,       CJ     O       p^H 


,  Dh        ~ 


«  ffi 


=J        N     ^ 


o    '^ 


O        05 


O      OOO^      O      — 'lO 


-^  ■*     QO 


<3 

® 

m 

fa 

+i 

•4J 

'C 

'C 

T3 

-ti 

■+^ 

-3 

'2 

'^ 

s 

s 

o 

tc 

ffi 

M 

W 

fi 

a 

Q 

P 

< 

<i: 

< 

*tj 

12; 

^ 

^ 

S5 

o 

d 

ft 

<5 

fa 

^^ 

'2 

00 

o 

3 

w 

ffi 

<?a 

ffi 

w 

Q 

Q 

Q 

p 

p 

< 

<1 

-< 

< 

< 

"z, 

^ 

2; 

;2; 

^; 

s 

cj 

^3 

c 

« 

s 

g 

'^ 

t- 

CO 

c 

o 

c3 

d 

d 

W 

W 

ffi 

^ 

P 

P 

P 

c 
o 

< 

< 

< 

"^ 

^ 

"Z 

g; 

PM 


^      o 


^ 

3 

» 

fl 

c5 

.^ 

+3 

o 

r^ 

o 

o 

^ 

£ 

-3 

c 

Oi 

W 

O 

W 

O 

P 

'S 

P 

^ 

<: 

'x 

<! 

'X 

^2; 

O 

^S 

o 

T3 

i 

s 

^ 

s 

"S 

;^ 

.g 

CS. 

=0 

5~ 

s 

« 

o 

U 

S 

e 

o 

►o 

rg 

o 

1 

^ 

X 

^  fp 


-t3 

P2 

'a 

g 
5 

s 

.2 

XD 

c 

c 
-3 

"3) 
o 

a 

SI 

1 

■ft 

S 

m 

s 

OQ 

o 

3 
TJ 

ffi 

ffi 

3 
t3 

w 

s 

T3 

0) 

<B 

« 

w 

o 

Cli 

£" 

O 

Ph 

O 

p 

-o 

P 

P 

V, 

p 

r2 

< 

'x 

<1 

-^ 

"x 

o 

< 

'x 
o 

^ 

o 

^ 

^ 

!z; 

554 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


P3 


A 


c« 

^ 

C 

■o 

S 

s 

^ 

-d 

,_^ 

T3 

c 

CO 

c 

eS 

o 

cS 

05 

.2 

c 

'S 

' — ■ 

9 

^ 

0) 

,^ 

CO 

o 

u 

b< 

5s 

+= 

IB 

o 

ce 

w 

w 

P^ 

W 

»0      CD 

d    d 


:?;  ^  ^;  13 


ft    e     s 


t^ 


-e 


D^    tt5      ^ 


g 

>^, 

p 

N 

73 

c 

« 

W 

u 

PM 


o 

+i 

o 

_S 

r* 

CD 
05 

"w 

Tl 

^•^ 

C 

c3 

c3 

O 
O 

S 

g 

O 

'5° 

m 

^ 

5 


fe; 


*? 

■s 

i 

s 

-d 

01 

5S 

Eh 

OJ 

O 

CO 

o 
t3 

t3 

^ 
^ 

c3 
O 

"3 

a; 

o 

o 

O 

j2 
ft 
ai 
O 

p 

j3 

(1 

P^ 

S 

ffO 

ft 

O      -H 

d  d 


P5 


P5 


o  t-  ■>* 

^    (M    -^ 


-H  O      r-H      O 


eS 

-d 

(U 

>J 

fi 

^ 

j3 

rj-J      ^ 

m 

o 

0) 

"o 

2 

4^     ^^ 

c 

t 

"S 

^-K  ^-^ 

^S 

^ 

(M 

<1 

(N  .a 

.■§  w 

ft 
o 

d 

Ph 

o 

< 

s 

^  > 

O    ^ 

<>- 

m 

PL, 

<: 

i, 

a 

Cm 

^s 

> 

fe; 


P5 


ANALOGS  OF  RIBOFLAVIN  AND  FAD  555 


™  3        fJ     X<     rS  OS 

T3  cS        y-,     ;;^    Ci  j^ 

fe  C  2        O  ^,  05 


(35 

'ao 

'm 

cS 

^^ 

c3 

" — ' 

t^ 

>H 

'e 

G 

"e 

T3 

■u 

ei 

■" 

cS 

s    -- 


rt    ffC    CO  fC        --  CC  (M 


Pi 

Ph 

s 

O 

B 

Q 

ai 

© 

73 

9*  •< 

B 

w  ^ 

a 

9^ 

'S 

B      C 

o 

^  I-     «        _.        3     5     ^ 


.2 

ID 

X 

fe 

o 

t. 

T3 

;:;• 

C 

H 

03 

^ 
e 

d 

-t 

g 

"3 
C 

S 

o 

o 

« 
-S 

pq 

s 

g 

O      S 


c3 

^ 

ao 

a 

o 

(D 

">> 

TS 

GO 

X 

>> 

o 

^ 

<D 

'x 

-S 

T3 

O 

>~. 

X 

QJ 

» 

+J 

-»^ 

73 

C 

1 

'o 

'o 

o 

o 

-is 

p 

3 

OQ 

CO 

iZ} 

3     t< 
a: 

!z;  2 


73 

<» 

« 

cS 

(i 

t« 

<u 

3 

-fj 

cc 

s 

O 

3 

IS 

CO 

H 

A 


P5  q'=Q'^f|         II 


556  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Some  of  the  enzymes  inhibited  by  quinacrine  do  not  involve  flavin  co- 
enzymes, as  emphasized  by  Hellerman  et  al.  (1946)  and  Hemker  and  Hiils- 
man  (1960),  so  that  no  direct  antagonism  with  FMN  or  FAD  would  be 
expected.  Nevertheless,  reversal  of  the  inhibition  is  sometimes  seen  and 
may  be  due  to  the  formation  of  a  complex  between  the  quinacrine  and  the 
added  coenzyme,  as  was  shown  to  occur  between  quinacrine  and  FMN  in 
the  inhibition  of  aliesterase,  this  presumably  removing  some  of  the  quina- 
crine from  the  enzyme.  In  any  case,  the  degree  of  reversal  or  even  whether 
reversal  occurs  with  FMN  or  FAD  will  depend  on  the  way  in  which  the  ex- 
periment is  run  and  the  relative  concentrations.  If  interaction  of  the  en- 
zyme and  quinacrine  is  allowed  to  occur,  the  chance  of  reducing  the  inhi- 
bition by  adding  coenzyme  is  less  than  if  both  are  added  together,  since  it 
may  be  difficult  to  reach  equilibrium  due  to  the  tight  binding  of  quinacrine. 

The  effects  of  quinacrine  on  ATPases  and  oxidative  phosphorylation  are 
interesting  and  perhaps  important  in  explaining  some  of  the  metabolic 
changes.  Myofibrillar  ATPase  does  not  depend  on  any  flavin  component 
and  yet  is  inhibited  rather  well  by  quinacrine  (Kaldor,  1960).  Furthermore, 
increasing  the  ATP  concentration  from  1  raM  (where  1  mM  quinacrine  in- 
hibits about  50%)  to  7  vaM  almost  abolishes  the  inhibition.  The  inhibition 
is  potentiated  by  Mg++  and  counteracted  by  Ca++,  so  that  a  quinacrine-Mg 
complex  was  postulated  as  the  possible  active  inhibitor.  Interactions  of 
Mg++  and  ATP  with  quinacrine  were  shown  by  fluorescence  changes.  Irvin 
and  Irvin  (1954)  had  found  that  quinacrine  forms  complexes  with  AMP 
and  ATP  at  physiological  pH's.  The  dissociation  constant  for  the  ATP  com- 
plex is  1.38  X  10"^,  so  that  1  raM  quinacrine  would  lower  the  ATP  con- 
centration from  1  vaM  to  0.67  raM,  whereas  it  would  have  a  negligible  ef- 
fect on  ATP  around  7  raM.  Increasing  the  ATP  might  also  reduce  the  Mg++ 
available  for  a  quinacrine  complex,  and  if  this  is  the  active  inhibitor,  the 
inhibition  would  be  lessened.  The  ATPase  activity  in  mitochondrial  pre- 
parations may  be  quite  different  from  the  myofibrillar  enzyme  and  could 
involve  a  flavin  component.  Quinacrine  inhibits  the  ATPase  of  beef  heart 
mitochondria  and  simultaneously  uncouples  oxidative  phosphorylation  even 
more  potently  (Penefsky  et  al.,  1960),  while  the  DNP-stimulated  ATPase 
of  rat  liver  mitochondria  is  stimulated  by  quinacrine  at  lower  concentra- 
tions (0.75  mM)  and  inhibited  by  higher  (3  raM)  (Low,  1959  a),  the  P,- 
ATP  exchange  reaction  and  oxidative  phosphorylation  being  depressed. 
Low  felt  that  this  provides  evidence  for  the  participation  of  flavin  in  such 
ATPase  activity,  especially,  as  FMN  and  FAD  can  reverse  the  inhibition, 
but  other  interpretations  are  possible  (e.g.,  complexes  between  quinacrine 
and  ATP  or  the  flavins).  There  is  no  doubt,  however,  that  DNP  alters  the 
response  of  mitochondrial  ATPase  to  quinacrine,  but  no  DNP  effect  is  ob- 
served in  muscle  ATPase  (Pennington,  1961).  Uncoupling  by  quinacrine 
was  first  reported  by  Loomis  and  Lipmann  (1948)  and  recent  work  seems 


ANALOGS  OF  RIBOFLAVIN  AND  FAD  557 

to  establish  a  true  uncoupling  action,  although  by  no  means  so  specific  as 
with  DNP  since  Og  uptake  is  usually  reduced  simultaneously  with  the 
P  :  0  ratio. 

Baltscheffsky  (1960  b)  found  that  light-induced  phosphorylation  in  spin- 
ach chloroplasts  is  strongly  inhibited  by  quinacrine,  0.04  mM  producing 
almost  complete  block,  and  in  cell-free  extracts  of  Rhodospirillum  rubrum 
less  potently  (Baltscheffsky  and  Baltscheffsky,  1958;  Baltscheffsky,  1960  a). 
The  inhibition  is  reversed  by  FMN  and  FAD  in  the  the  bacterial  extracts, 
but  not  at  all  in  the  chloroplasts;  indeed,  in  the  latter  FMN  and  FAD  are 
quite  potent  inhibitors.  It  was  suggested  that  an  endogenous  flavin  is  a 
necessary  component  of  the  system.  Photophosphorylation  has  recently  been 
found  to  be  very  sensitive  to  quinacrine.  In  Rhodospirillum  chromatophores 
quinacrine  begins  to  depress  the  photophosphorylation  at  0.0001  mM,  in- 
hibits 65%  at  0.028  mM,  and  blocks  completely  at  0.1  mM,  the  K,  being 
0.003  mM  (Horio  and  Kamen,  1962  a).  The  characteristic  response  to  ribo- 
flavin is,  however,  not  prevented  and  it  was  postulated  that  quinacrine 
binds  at  some  locus  in  the  respiratory  chain.  Quinacrine  at  concentrations 
around  0.05  mM  uncouples  all  types  of  photophosphorylation  in  Swiss  chard 
chloroplasts  and  simultaneously  stimulates  the  photoreduction  of  dichloro- 
phenolindophenol  (Gromet-Elhanen  and  Avron,  1963).  Similar  effects  were 
observed  in  spinach  chloroplasts,  quinacrine  at  0.02  mM  inhibiting  photo- 
phosphorylation 61%  and  at  0.05  mM  inhibiting  completely,  at  the  same 
time  stimulating  the  photoreduction  of  trimethyl-l,4-benzoquinone  (Dilley 
and  Vernon,  1964).  Changes  in  light  absorption  and  scattering  indicate  a 
relationship  between  photosphosphorylation  and  structural  alterations  in 
the  chloroplasts,  but  it  is  not  known  if  quinacrine  modifies  directly  these 
structural  changes. 

The  determination  of  the  inhibitor  constant,  K„  for  quinacrine  and  sim- 
ilar substances  is  somewhat  more  complex  than  with  most  inhibitors,  due 
to  the  fact  that  equilibrium  is  difficult  to  achieve  and  mutual  depletion 
kinetics  must  be  applied,  the  free  concentrations  of  both  quinacrine  and 
FAD  being  much  lower  than  the  total  concentration.  HeUerman  et  al.  (1946) 
considered  these  problems  relative  to  the  inhibition  of  D-amino  acid  oxidase 
and  described  a  very  useful  technique  with  the  appropriate  equations  for 
the  calculation  of  K^,  The  ^fad  is  0.00057  mM,  and  K^  for  quinine  is 
0.67  mM  (means  for  two  enzyme  preparations).  The  quinine  inhibition  is 
quite  competitive  but  quinacrine  behaves  atypically  and  its  K^  varies  with 
the  experimental  conditions  (it  is  somewhat  smaller  than  the  K^  for  qui- 
nine). 

An  interesting  example  of  the  effects  of  pH  on  inhibition  was  reported 
by  Molinari  and  Lara  (1960)  for  the  lactate  dehydrogenase  of  Propionihac- 
terium  pentosaceum  (Fig.  2-19).  Increase  of  pH  augments  the  inhibition  by 
quinacrine  whereas  the  opposite  effect  is  seen  on  Dicumarol  inhibition, 


558 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


but  these  relationships  are  reasonable  if  one  assumes  that  the  negatively 
charged  Dicumarol  reacts  with  positively  charged  enzyme  groups,  and  the 
positively  charged  basic  quinacrine  reacts  with  enzyme  anionic  groups. 

We  must  finally  evaluate  the  reliability  of  quinacrine  as  an  indicator  of 
flavin  participation  in  enzyme  reactions.  Certainly  the  mere  inhibition  of 
an  enzyme  by  quinacrine  does  not  imply  involvement  of  a  flavin  coenzyme. 


100 


80 


60 


40 


n    20 


% 

INH 


Fig.  2-19.  Effects  of  pH  on  the  inhibitions  of  lactate  dehydrogenase 

from  Propionibacterium  pentosaceum  by  quinacrine  and  Dicumarol 

at  0.1   mM.  (From  Molinari  and  Lara,  1960.) 


The  observations  of  Hemker  and  Hiilsman  (1960)  support  the  opinion  of 
Hellerman  et  al.  (1946)  that  quinacrine  is  a  relatively  nonspecific  inhibitor, 
due  to  its  affinity  for  proteins  in  general.  If  a  reversal  of  the  inhibition  by 
FMN  or  FAD  is  demonstrated,  there  is  more  likelihood  for  the  participation 
of  a  flavin,  but  even  here  one  must  consider  the  possible  complexing  of  the 
quinacrine  by  the  reversing  flavin.  Also  it  seems  that  flavin  nucleotides 
which  are  not  coenzymically  active  are  often  as  good  reversors  as  the  co- 
enzyme itself,  indicating  some  other  mechanism  than  competition  for  the 
active  center.  It  would  seem  to  me  that  quinacrine  would  be  one  of  the  least 
likely  specific  antagonists  of  FMN  or  FAD,  since  structurally  it  is  not  as 
close  as  many  other  analogs.  An  ideal  indicator  for  flavins  would  be  phos- 


ANALOGS   OF   RIBOFLAVIN    AND    FAD  559 

phorylated  or  in  the  form  of  a  dinucleotide  analog,  and  such  has  not  yet 
been  found. 

Many  flavin-dependent  enzymes  bind  FAD  or  other  flavin  coenzymes  very 
tightly  and  it  is  difficult  to  understand  how  quinacrine  could  displace  these, 
or  how  exogenous  FMN  or  FAD  could  antagonize  the  action  of  quinacrine 
if  this  is  the  case.  If  an  enzyme  after  extraction  is  flavin-dependent  and  is 
catalytically  active,  it  must  have  very  tightly  bound  coenzyme;  perhaps 
quinacrine  can  react  with  the  bound  coenzyme  but  there  is  no  evidence  for 
this.  Certainly  the  failure  of  quinacrine  to  inhibit  should  not  be  taken  as 
evidence  for  the  absence  of  a  flavin  component.  Actually  it  must  be  said 
that  many  of  the  experiments  with  quinacrine  have  not  been  done  properly. 
In  some  instances  one  concentration  of  quinacrine  has  been  used  and,  if 
inhibition  of  any  degree  is  noted  it  is  stated  that  this  is  evidence  for  a 
flavoenzyme,  even  though  no  antagonism  has  been  demonstrated;  it  should 
be  obvious  that  no  conclusions  can  be  drawn  from  these  results.  In  other 
cases  two  experiments  have  been  run,  one  with  quinacrine  alone  and  one 
with  both  quinacrine  and  either  FMN  or  FAD;  if  the  inhibition  is  less  in 
the  presence  of  the  FMN  or  FAD  it  is  concluded  that  quinacrine  is  inter- 
fering with  the  normal  function  of  a  flavin  coenzyme.  A  control  with  FMN 
or  FAD  alone  must  also  be  run,  since  in  many  cases  these  substances  will 
stimulate  activity.  The  results  of  Bargoni  (1963)  are  difficult  to  interpret 
in  that  FAD  would  prevent  the  inhibition  by  quinacrine  only  if  the  FAD 
were  incubated  with  the  enzyme  for  30  min  before  the  inhibitor  is  added. 
The  binding  of  quinacrine  to  some  enzymes  is  readily  reversible,  but  yeast 
lactate  dehydrogenase  is  irreversibly  inactivated;  FAD  will  slow  this  inac- 
tivation  but  will  not  reactivate  (Iwatsubo  and  Labeyrie,  1962).  Several 
suggestions  as  to  the  design  of  such  antagonism  tests  may  be  made:  (1)  use 
a  flavin  derivative  which  is  the  most  likely  coenzyme  involved,  (2)  use  co- 
enzyme-dissociated  and  reconstituted  enzymes  whenever  possible,  (3)  al- 
ways have  a  control  with  the  reversor  alone,  and  (4)  attempt  to  establish 
competitive  relations  between  the  quinacrine  and  the  reversor. 

Effects  of  Quinacrine  on  Metabolism 

Quinacrine  and  other  antimalarials  were  found  by  Fulton  and  Christo- 
phers (1938)  to  depress  the  respiration  of  trypanosomes.  Concentrations  as 
low  at  0.004  mM  exert  inhibitory  effects  on  multiplication  but  it  requires 
0.3  mM  to  inhibit  the  respiration  11.5%,  at  which  concentration  the  count 
is  reduced  26.3%.  These  observations  and  several  others  afterward  dem- 
onstrate that  concentrations  presumably  higher  than  are  present  in  vivo 
must  be  used  to  depress  the  respiration  of  these  organisms.  Wright  and 
Sabine  (1944)  studied  the  effects  of  quinacrine  on  the  respiration  of  rat 
tissue  slices.  In  most  instances  there  is  an  initial  stimulation  followed  by  a 
slowly  developing  inhibition,  often  not  complete  after  100-200  min.  Liver 


560  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

slice  respiration  is  depressed  only  slightly  by  0.5  n\M,  even  after  2  hr, 
while  1-2  mM  produces  a  maximal  depression  of  around  65%  after  1  hr. 
Brain  respiration  is  more  sensitive  and  is  reduced  around  75%  by  0.5  mM. 
Following  inhibition  by  quinacrine,  addition  of  pyruvate,  lactate,  citrate, 
fumarate,  or  malate  does  not  restore  oxygen  uptake,  but  addition  of  succi- 
nate brings  about  a  rapid  rise  in  respiration,  indicating  only  that  succinate 
oxidase  is  not  blocked  significantly  at  these  concentrations.  Because  of  the 
reversal  of  D-amino  acid  oxidase  inhibition  by  coenzyme,  they  suggested 
that  the  block  may  be  around  the  flavoenzyme  locus  in  the  respiratory 
chain,  but  actually  there  is  no  evidence  for  this.  The  respiration  of  Plasmo- 
dium lophurae  with  different  substrates  is  depressed  15-24%  by  0.1  mM 
quinacrine  and  somewhat  above  50%  by  1  mM  (Bovarnick  et  al.,  1946). 
Actually  it  is  very  difficult  to  compare  in  vitro  and  in  vivo  effects  and  con- 
centrations because  of  the  progressive  binding  and  accumulation  of  quina- 
crine; in  other  words,  the  free  plasma  concentration  of  quinacrine  means 
very  little,  nor  does  total  tissue  concentration  necessarily  relate  to  any 
enzyme  effects. 

A  glycolytic  inhibition  by  quinacrine  was  suggested  by  the  early  work 
of  Marshall  (1948),  who  found  a  depression  of  glucose  utilization,  a  decrease 
in  glucose- 1-P,  an  increase  in  glucose-6-P,  some  decrease  in  triose-P's,  and 
decreases  in  pyruvate  and  lactate  in  washed  chick  erythrocytes  parasitized 
by  Plasmodium  gallinaceum.  It  is  difficult  to  separate  the  metabolic  effects 
on  parasite  and  erythrocyte,  but  it  is  probable  that  the  major  fraction  of 
the  glucose  utilization  was  due  to  the  parasites.  The  most  marked  effect  of 
quinacrine  is  an  accumulation  of  ATP,  which  Marshall  attributed  to  an  in- 
hibition of  hexokinase. 

The  only  analysis  of  the  effects  of  quinacrine  on  metabolism  was  made 
by  Bowman  et  al.  (1961).  The  glucose  utilization  of  P.  bergkei  free  parasites, 
parasitized  reticulocytes,  and  reticulocytes  was  determined,  and  low  con- 
centrations of  quinacrine  (claimed  to  be  near  those  found  therapeutically) 
exhibit  a  selective  action  on  the  parasites.  The  glucose  utilization  over  1  hr 
is  reduced  34%  by  0.0125  mM  and  94%  by  0.035  mM  quinacrine.  There 
is  no  effect  on  the  pattern  of  glucose-1-C^*  and  glucose-6-C^^  distribution. 
The  amount  of  lactate  formed  from  glucose  is  reduced  and  there  is  an  accu- 
mulation of  hexose-6-P,  so  it  was  concluded  that  quinacrine  inhibits  some 
enzyme  which  is  involved  in  the  utilization  of  hexose-6-P  and  is  normally 
rate-limiting  in  the  free  parasites;  this  enzyme  may  be  phosphofructokinase. 
Depression  of  respiration  could  thus,  in  part,  be  attributed  to  a  glycolytic 
inhibition  and,  if  so,  is  probably  not  related  to  a  flavoenzyme. 

There  is  probably  need  for  more  investigation  of  the  effects  of  quinacrine 
on  tissues  and  parasites  in  animals  given  the  drug  for  varying  times,  be- 
cause of  the  difficulty  in  estimating  the  proper  in  vitro  concentrations  to 
use.  It  may  well  be  that  some  enzyme  system  not  previously  examined  is 


ANALOGS    OF    PYRIDOXAL  561 

most  potently  inhibited.  With  regard  to  the  inhibition  of  growth,  it  might 
be  well  to  consider  more  carefully  the  changes  resulting  from  complexes 
formed  between  quinacrine  and  nucleotides  or  nucleic  acids. 


ANALOGS  OF  PYRIDOXAL 

A  group  of  substances,  including  pyridoxol,  pjTidoxal,  pyridoxamine,  and 
their  phosphates,  possess  vitamin  Bg  activity  and  these  will  be  designated 
as  pyridoxine  in  accordance  with  Braunstein  (1960)  and  the  Commission 
on  Chemical  Terminology  of  the  International  Union  of  Pure  and  Applied 
Chemistry  (the  substance  previously  called  pyridoxine  now  being  pyridoxol). 
These  substances  are  converted  to  pyridoxal  which  is  metabolically  func- 
tional in  the  form  of  its  phosphate.  Pyridoxal-P  is  the  coenzyme  for  a  large 
number  of  enzymes  involved  in  the  decarboxylation,  transamination,  oxi- 
dative deamination,  racemization,  a,/?-cleavage,  and  /?-  and  y-substituent 
replacement  in  amino  acid  metabolism,  and,  in  addition,  may  be  active  in 
amino  acid  transport.  Disturbances  in  pyridoxal  metabolism  or  functions 
will  thus  bring  about  primarily  alterations  in  the  biosynthesis  and  degra- 
dation of  amino  acids,  and  indirectly  will  affect  protein  synthesis  and  a 
variety  of  other  metabolic  pathways.  Possibly  the  most  important  biochem- 
ical defect  will  be  the  reduction  in  transaminations  involving  glutamate, 
inasmuch  as  these  reactions  are  central  in  amino  acid  metabolism.  More 
recently  it  has  been  found  that  phosphorylase  a  contains  pyridoxal-P,  per- 
haps bound  in  an  aldamine  linkage,  and,  although  initially  it  was  believed 
that  it  is  enzymically  nonfuctional,  the  demonstration  by  Illingworth  et  al. 
(1958)  that  pyridoxal  and  5-deoxy pyridoxal  will  prevent  the  binding  of 
pyridoxal-P  and  enzyme  activity  points  to  some  role  of  the  pyridoxal-P 
in  the  catalysis. 

Animals  generally  require  pyridoxine  whereas  plants  and  most  micro- 
organisms can  synthesize  pyridoxal.  Little  is  known  about  the  biosynthesis 
of  pyridoxal,  but  the  rather  complex  interrelationships  between  the  pyri- 
doxines  and  their  phosphates  are  now  fairly  clear.  The  pathways  and  the 
enzymes  involved  are  summarized  in  the  accompanying  diagram  (Braun- 
stein, 1960;  Wada  and  Snell,  1961).  It  is  believed  that  the  major  pathway 
for  the  formation  of  pyridoxal-P  is 

Pyridoxol  -^  pyridoxol-P  ->  pyridoxal-P 

The  primary  excretory  metabolite  of  pyridoxal  is  4-pyridoxate  and  its  lac- 
tone. Most  of  the  pyridoxine  in  tissues  is  present  as  pyridoxal-P  bound  quite 
tightly  to  enzymes  and  other  proteins.  Certain  analogs  can  inhibit  the  for- 
mation of  pyridoxal-P  and  may  act  partly  in  this  way,  while  other  analogs 
may  be  phosphorylated  and  compete  with  pyridoxal-P. 


562 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

(2)       ^ 


Pyridoxol 


4-pyridoxate 


(6) 


(1) 

r 

pyridoxal 

(5)      (4) 


(3) 
(2) 


pyridoxol- P 


(5) 


pyridoxamine  ^^ 


(3) 
(2)    , 


pyridoxal -P 


(5)^ 


(4) 


pyridoxamine- P 


(1)  pyridoxol  oxidase 

(2)  pyridoxal  kinase 

(3)  phosphatase 


(3) 

(4)  transaminase 

(5)  pyridoxol- P  oxidase 

(6)  aldehyde  oxidase 


Two  general  types  of  analog  are  theoretically  possible  —  those  resulting 
from  alteration  of  the  pyridine  ring  and  those  in  which  the  substituent 
groups  are  modified,  replaced,  or  eliminated  —  but  practically  it  has  been 
found  thus  far  that  only  analogs  of  the  second  type  are  effective.  Actually, 
not  many  really  effective  analogs  have  been  found.  The  most  commonly 
used  analog  has  been  deoxypyridoxol*  and  we  shall  limit  our  discussion 
mainly  to  this  substance.  Unless  otherwise  noted,  the  name  deoxypyridoxol 
will  refer  only  to  the  4-derivative. 

Deoxypyridoxol  was  found  to  have  no  vitamin  Bg  activity  by  Unna  (1940) 
and  to  be  an  antagonist  of  pyridoxine  in  the  chick  by  Ott  (1946).  Chicks 
on  a  low  pyridoxine  diet  can  be  killed  by  as  little  as  16  //g  deoxypyridoxol, 
whereas  normal  chicks  on  an  adequate  pyridoxine  diet  can  withstand  as 
much  as  600  //g.  By  varying  the  relative  doses  of  both  vitamin  and  analog, 
Ott  showed  that  approximately  2  molecules  of  analog  can  counteract  1 
molecule  of  pyridoxine.  Deoxypyridoxol  has  since  been  found  to  inhibit  cer- 
tain microorganisms  and  to  produce  symptoms  of  vitamin  Bg  deficiency  in 
animals,  including  man.  4-Methoxymethylpyridoxol  (usually  called  methox- 
ypyridoxine)  was  found  by  Unna  to  have  slight  vitamin  activity  in  the 
rat,  but  Ott  (1947)  demonstrated  a  potent  inhibitory  effect  in  the  chick. 
The  ability  of  rats  to  use  this  analog  is  related  to  its  transformation  to 
pyridoxal  in  these  animals  (Porter  et  al.,  1947).  In  the  chick  it  is  about  25 
times  as  effective  as  deoxypyridoxol.  These  are  apparently  the  only  analogs 
so  far  tested  that  can  produce  rather  typical  pyridoxine-deficiency  symp- 
toms in  animals,  although  several  others  can  inhibit  microbial  growth  by 
disturbing  pyridoxal  function.  Toxopyrimidine  is  undoubtedly  toxic  to  ani- 
mals and  can  be  antagonized  by  pyridoxine,  but  it  is  debatable  whether 

*  This  is  4-deoxypyridoxol  and  has  previously  been  called  desoxypjTidoxine  or 
deoxypyridoxine.  However,  if  we  are  to  conform  to  the  modern  nomenclature,  the 
specific  compound  must  be  deoxypyridoxol.  Deoxypyridoxine  might  be  used  to  refer 
to  the  entire  group  of  deoxy  substances  exhibiting  vitamin  Bg  activity  antagonism. 


ANALOGS    OF    PYRIDOXAL 


563 


CH20H 

CHO 

CHO 

H0^J:;;;^^CH20H 

HO.^^4^^^ 

CH2OH 

HO  ^^^Av^  CHjOPd; 

H3C'^N+ 
^            H+ 

Pyridoxol 
(pyridoxine) 

Pyridoxal 

Pyridoxal-P 

CH2NH3" 

HO^    >\    ^CH,OH 


H,C         N. 


CHjOH 

CH2OH 


H3C         N 
H 


CH3 
HO^^X/CH^OH 


XJ 


H3C  N' 


Pyridoxamine 


3-Deoxypyridoxol 


4-Deoxypyridoxol 
(deoxypyridoxine) 


CH2OH 
HO^J^CH3 


CH2OCH3 
H0^A^.^CH30H 


•XJ 


H^ 


CH2OH 
HO,      X^     ^CH,OH 


H,C— CKT  ^NI 


5  -  Deoxypy  r  idoxo  1 


4  -  Methoxy  methyl  - 
pyridoxol 


w-Methylpyridoxol 


CH2OCH2CH3  CH=NOH 

^^^..^^'Cv^CHaNHa  HO^    JX    /CH,OH 


--L  y^ 


H,C— CH,        Nl 

3  -2  JJ+ 


"XT' 

H^ 


nh: 


H3C       "N 
H 


CH2OH 


2-Ethyl-3-amino- 

4-ethoxyniethyl-5- 

aminomethylpyridine 


Pyridoxal  oxime 


CHO 


Toxopyrimidine 


CH3 
HOv.    A^  /OH 


HaC^     Te 


NO, 


4-Nitrosalicylaldehyde  ,    2,  4-Dimethyl- cvcjo- 

•'  ^  telluropentane-3, 5-dione 


564  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

it  can  be  termed  a  true  pyridoxine  analog;  this  substance  will  be  discussed 
separately  at  the  end  of  this  section. 

Effects  on   Pyridoxine  Metabolisnn  and  Tissue  Levels  of  Pyridoxine 

The  only  analog  found  in  the  early  work  to  inhibit  pyridoxal  kinase  strong- 
ly is  2-ethyl-3-amino-4-ethoxymethyl-5-aminomethylpyridine  (Hurwitz, 
1952).  This  substance  cannot  be  phosphorylated  because  of  the  lack  of  a 
hydroxyl  group  at  position  5,  but  has  a  rather  high  aJBinity  for  the  yeast 
enzyme  {K^  =  0.073  mM)  and  competitively  inhibits  the  phosphorylation 
of  pyridoxal.  Certain  analogs  can  be  phosphorylated  by  this  enzyme  (e.g., 
deoxypyridoxol,  3-deoxypyridoxol,  and  the  3-amino  analog  of  pyridoxol) 
(Umbreit  and  Waddell,  1949;  Hurwitz,  1955  b),  and  presumably  could  re- 
duce the  phosphorylation  of  pyridoxal  through  substrate  competition.  Phos- 
phorylation was  shown  to  require  a  hydroxymethyl  group  at  the  5-position 
and  the  absence  of  a  substituent  at  position  6.  The  more  recent  work  of 
McCormick  and  Snell  (1961 )  demonstrated  other  potent  inhibitors  and  made 
it  clear  that  the  kinases  from  different  sources  vary  markedly  with  respect 
to  affinity  for  the  analogs  (Table  2-35).  Furthermore,  the  relative  affinities 
are  fairly  well  correlated  with  the  abilities  to  inhibit  growth  of  the  various 
bacteria.  The  inhibitions  are  not  always  competitive  and,  in  some  instances, 
increase  with  pyridoxal  concentration.  The  3-hydroxy  group  can  be  replaced 
by  an  amino  group  or  omitted  without  affecting  affinity  adversely,  but 
substitution  in  the  6-position  reduces  the  affinity  without  necessarily  abol- 
ishing it.  It  is  interesting  that  A^-methylpyridoxal  is  completely  inactive  as 
an  inhibitor.  The  most  potent  inhibitors  of  pyridoxal  kinase  are  derivatives 
obtained  by  the  reaction  of  pyridoxal  with  various  carbonyl  agents.  The 
beef  brain  kinase  is  inhibited  50%  by  0.00005  mM  pyridoxal  semicarbazone 
and  by  0.000065  mM  pyridoxal  azine  (the  product  of  the  reaction  of  2 
pyridoxals  with  hydrazine),  but  the  discussion  of  such  inhibitions  is  more 
pertinent  to  the  subject  of  the  carbonyl  agents. 

Various  enzymes  oxidizing  pyridoxol,  pyridoxol-P,  or  pyridoxamine-P 
have  recently  been  found  in  liver,  and  are  occasionally  inhibited  by  analogs. 
The  oxidation  of  pyridoxol  is  competitively  inhibited  by  deoxypyridoxol 
(67%  when  pyridoxol  is  10  mM  and  the  analog  is  12.5  mM)  (Morino  et  al., 
1960),  while  the  oxidation  of  pyridoxol-P  is  competitively  inhibited  by 
deoxypyridoxol-P  {K„,  =  0.02  mM,  and  K,  =  0.35  mM)  (Morisue  et  al, 
1960).  These  enzymes  thus  follow  the  general  rule  that  nonphosphorylated 
analogs  inhibit  the  reactions  of  nonphosphorylated  substrates,  and  phos- 
phorylated analogs  inhibit  the  reactions  of  phosphorylated  substrates.  The 
oxidative  deamination  of  pyridoxamine-P  is  inhibited  by  pyridoxamine  (the 
latter  is  also  deaminated  at  a  slower  rate)  and  rather  weakly  by  pyridoxol 
(Pogell,  1958).  Wada  and  Snell  (1961)  examined  the  competitive  inhibitions 
of  pyridoxol-P  oxidase  by  a  variety  of  substances  (see  accompanying  tab- 


ANALOGS    OF    PYRIDOXAL 


565 


PQ 


^        o 
d        d 


d        d 


2Q 


o        o 

d        d 


in 

O  lO 

d        -« 


^ 


-r        -A    >> 


o 

o 

o 

n 

';-i 

^ 

>•. 

>i 

p 

fO 

Ph 

ft 

^^ 

^ 

>■- 

>. 

t>> 

>■. 

X 

X 

^ 

f^ 

o 

O 
IB 

1 

-* 

lO 

3 

<N 

-a 


P-i 


o 

« 

o 

^ 

r2 

'E 

^ 

>> 

o 

ft 

^^ 

t3 

, 

C 

"« 

eS 

^.^ 

CO 

c3 

X 

-c 

O 

c 

'2 

cS 

'E 

^ 

ft 

•p 

t-i 

c 

o 

(-> 

Ch 

a 

in 

0) 

_2 

a 

> 

2 

« 
^ 

&^ 

H 

566 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


%  Inhibition  of  oxidation  of: 

Inhibitor 

Concentration 
(mM) 

Pyridoxol-P 

Pyridoxamine-P 

(0.3  mM) 

(0.3  mM) 

Pyridoxal 

1 

6 

7 

Pyridoxol 

1 

<3 

<3 

Pyridoxamine 

1 

<3 

<3 

4-Pyridoxate 

1 

3 

5 

4-PyTidoxate-P 

0.2 

23 

52 

1 

33 

62 

Deoxypyridoxol 

1 

0 

0 

Deoxypyridoxol-P 

0.2 

34 

54 

1 

42 

70 

Pyridoxal  oxime 

0.001 

17 

31 

0.01 

57 

67 

ulation)  and  noted  that  only  the  phosphorylated  derivatives  are  signifi- 
cantly inhibitory.  Whether  inhibition  of  these  oxidases  by  analogs  plays  a 
role  in  the  depressant  or  toxic  effects  produced  is  at  present  unknown,  but 
it  must  be  admitted  that  for  deoxypyridoxol  and  its  phosphate  none  of 
the  inhibitions  is  probably  potent  enough  to  be  important  in  vivo. 

The  effects  of  analogs  on  the  tissue  levels  of  the  vitamin  Bg  group  are 
particularly  important  in  certain  arguments  relative  to  the  mechanisms  by 
which  these  analogs  are  toxic.  Umbreit  (1955  a)  believes  that  deoxypyri- 
doxol exerts  actions  other  than  the  antagonism  of  vitamin  Bg  function. 
The  basis  for  this  is  principally  that  deoxypyridoxol  accelerates  the  ap- 
pearance of  deficiency  symptoms  when  animals  are  on  a  diet  lacking  pyri- 
doxine  and  yet  does  not  reduce  the  tissue  levels  of  pyridoxal  coenzymes. 
He  has  also  pointed  out  that  in  some  cases  there  is  also  no  fall  in  transamin- 
ase or  decarboxylase  activity  during  the  "acute"  deficiency  produced  by 
deoxypyridoxol.  Nevertheless,  it  is  admitted  that  the  toxic  effects  of  de- 
oxypyridoxol can  be  readily  counteracted  by  pyridoxine  administration. 
Actually  there  is  very  little  published  on  tissue  levels  of  vitamin  Bg  as  af- 
fected by  deoxypyridoxol.  Stoerk  (1950)  reported  that  dietary  deficiency 
lowers  liver  pyridoxine  content  but  that  deoxypyridoxol  produces  no  fur- 
ther lowering  despite  a  more  rapidly  appearing  deficiency  syndrome.  Similar 
results  were  obtained  by  Beaton  and  McHenry  (1953)  in  rats  exhibiting 
acrodynia  following  deoxypyridoxol  feeding  (see  accompanying  tabulation). 
Umbreit  also  cites  unpublished  data  supporting  these  results.  Effects  of 
deoxypyridoxol  on  enzyme  activity  in  vivo  will  be  taken  up  in  the  following 


ANALOGS    OF    PYRIDOXAL  567 

section,  but  there  is  now  sufficient  evidence  that  the  activities  of  certain 
pyridoxal-P-dependent  enzymes  are  reduced. 


„     .J      .  -p.  J       1  Liver  vitamin  Br 

ryndoxine  JJeoxypyridoxol 


(/'g/g) 


-  -  6.3 
+  -  11.0 

-  +  6.3 

+  +  9.6 


Interpretation  of  this  apparent  discrepancy  between  toxic  reactions  and 
insignificant  changes  in  liver  vitamin  Bg  during  deoxypyridoxol  treatment 
can  be  made  along  several  lines.  In  the  first  place,  it  is  generally  believed 
that  much  of  the  tissue  pyridoxal  is  bound  to  nonenzyme  protein,  possibly 
in  part  through  the  aldehyde  group,  so  that  analyses  of  total  tissue  levels 
do  not  necessarily  reflect  changes  in  coenzyme  concentration.  The  fact  that 
enzyme  activity  is  often  depressed  without  significant  changes  in  total  vi- 
tamin Bg  suggests  that  this  can  be  an  explanation.  That  bound  to  nonen- 
zyme protein  may  not  be  displaced  by  deoxypyridoxol  since  this  analog 
contains  no  aldehyde  group.  In  the  second  place,  analyses  have  been  made 
only  in  the  liver  and  it  is  unlikely  that  changes  in  liver  pyridoxal  function 
are  responsible  for  any  of  the  common  toxic  symptoms  of  deficiency.  It  is 
even  possible  that  during  treatment  with  deoxypyridoxol  there  is  a  transfer 
of  vitamin  Bg  substances  from  one  tissue  to  another.  It  is  conceivable  that 
deoxypyridoxol  has  actions  other  than  interference  with  pyridoxal  function, 
but  the  fact  that  its  toxicity  can  be  reduced  by  pyridoxine  administration 
points  to  a  close  relationship  between  its  effects  and  pyridoxal.  I  think 
that  more  emphasis  must  be  placed  on  the  changes  in  enzyme  activity 
rather  than  on  the  tissue  levels  of  vitamin  Bg  for  the  reasons  given  above. 

The  most  complete  investigation  of  the  effects  of  deoxypyridoxol  on  the 
concentrations  of  tissue  Bg  vitamers  is  that  of  Bain  and  Williams  (1960), 
who  utilized  chromatographic  separation.  The  results  are  summarized  in 
Table  2-36.  The  extreme  fall  in  brain  pyridoxal-P  they  believe  must  in  some 
way  be  related  to  the  convulsions.  The  rapidity  with  which  the  analog  can 
deplete  the  brain  coenzyme  is  surprising.  Is  the  pyridoxal-P  replaced  on 
the  apoenzymes  by  deoxypyridoxol-P  ?  It  seems  unlikely  that  interference 
with  transport  or  metabolism  of  pyridoxol  could  produce  such  marked  ef- 
fects so  soon.  Also  the  displaced  pyridoxal-P  must  be  metabolized  or  leave 
the  tissue.  It  is  not  known  why  pyridoxamine-P  does  not  fall  comparably. 
Dietary  vitamin  Bg  deficiency  for  37-51  days  does  not  cause  such  marked 
losses  of  pyridoxal-P  or  total  Bg  vitamers  from  the  brain  as  the  single  dose 
of  deoxypyridoxol.  The  severe  drop  in  pyridoxal-P  in  brain  following  deoxy- 


568 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


O) 


CO 

Pn 

to 

o 

^ 

K! 

w 
a 

o 

pq 

H 

< 

<i1 

H 

OS 

H 

^ 

H 

U 

;z; 

o 

o 

cq 


Q 


Ph 


>>     O 

Ph  'TS 


COOM  OOi-O-^  OfM>0  C^iO 

1— IGO-*  OifC-*  5<lTt<C<l  lOi— I 

00^1  locoi  r~>0|  (Mfo 


o 

o 

„H 

05 

O 

c- 

o 

Tt< 

M 

C<l 

1— 1 

1 

<M 

i—i 

fc    c^         rt 


O 


X 

o 

IS 

'u 

>> 

.-^ 

a 

o 

>> 

t-l 

X 

c 

o 

o 

« 

o 

fi 

M 


+ 


+ 


+ 


00    t^    ■* 


lO    o    o 

fS 

(M     lO     iT^I 

3 

t— 1     1— 1     _j_ 

> 

+ 


A  r-.  ft         3 

O       2    >>  a 


d  Q 


X    ^ 
o 


d  Q 


« 

CD 

ft 

05 

cS 

r-( 

u 

■i^ 

a 

CO 

s 

W) 

.2 

J4 

■— ] 

s 

^ 

o 

73 

pp 


:2  &H 

X  >. 

o  ^ 

T3  t3 


ANALOGS    OF    PYRIDOXAL  569 

pyridoxol  must  be  reflected  in  depression  of  the  activity  of  enzymes  requir- 
ing this  coenzyme,  but  not  necessarily  in  all  equally.  The  failure  of  pyridox- 
amine-P  to  fall  as  much  as  pyridoxal-P  might  indicate  that  transaminases 
are  not  depleted  as  readily  as  other  pyridoxal-P  enzymes,-^  but  interpreta- 
tion is  made  difficult  by  the  fact  that  we  do  not  know  what  fractions  of 
these  substances  are  bound  to  apoenzymes  and  to  nonenzyme  protein.  The 
effects  on  the  tumors  are  similar  but  of  less  magnitude,  and  the  changes  in 
pyridoxamine-P  are  again  variable,  even  increasing  in  ascites  tumor.  The 
rise  in  pyridoxal-P  and  total  vitamers  in  ascites  fluid  may  reflect  the  loss 
of  these  substances  from  the  cells. 

Effects  of  Deoxypyridoxol  on  Pyridoxal-P-Dependent  Enzymes  in  Vivo 

The  results  reported  on  transaminase  activity  during  administration  of 
deoxypyridoxol  are  variable.  Transamination  in  hamster  hearts  is  reduced 
30-40%  in  animals  with  a  dietary  deficiency,  but  injecting  deoxypyridoxol 
at  50  //g  per  animal  3  times  a  week  does  not  lower  the  activity  further 
(Shwartzman  and  Hift,  1951).  However,  there  is  some  growth  inhibition 
beyond  that  shown  in  the  deficient  animals,  although  no  specific  symptoms 
were  noted.  Deoxypyridoxol  at  100  //g/day  in  rats  does  not  alter  the  aspar- 
tate-glutamate  transaminase  and  actually  seems  to  increase  the  alanine- 
glutamate  transaminase  activity  in  liver  compared  to  animals  on  a  deficient 
diet  (Caldwell  and  McHenry,  1953).  Since  the  animals  receiving  deoxypyri- 
doxol had  severe  dermatitis,  it  was  justifiably  concluded  that  the  production 
of  dermatitis  is  unrelated  to  liver  transaminase.  On  the  other  hand,  Dietrich 
and  Shapiro  (1953  a)  found  a  greater  fall  in  liver  transaminase  when  mice 
were  injected  with  deoxypyridoxol  at  150  mg/kg/day  than  in  simple  dietary 
deficiency  (  —  49%  and  —37%,  respectively).  Indeed,  transaminase  levels 
in  several  tissues  fall  very  markedly  in  mice  on  175  mg/kg/day  of  deoxy- 
pyridoxol (Shapiro  et  al.,  1953).  There  is  not  much  difference  in  the  rates 
of  decrease  in  the  various  tissues  (Fig.  2-20).  It  is  difficult  to  reconcile  all 
of  these  results  unless  it  is  a  matter  of  species  variation,  which  is  unlikely. 
Transaminases  are  not  directly  inhibited  very  potently  by  deoxypyridoxol; 
the  K^  is  0.12  m.M  for  the  alanine-pyruvate  transaminase  of  Pseudomonas, 
for  example  (Dempsey  and  Snell,  1963).  The  very  high  inhibitory  activity 
of  pyridoxyl-L-alanine  {K,  =  0.00018  mM)  is  surprising. 

The  results  on  decarboxylases  are  similar.  Dietary  pyridoxine  deficiency 
causes  a  50%  fall  in  rat  brain  glutamate  decarboxylase,  but  administering 
deoxypyridoxol  in  addition  produces  no  further  depletion  (Roberts  et  al., 
1951).  More  recent  studies,  summarized  in  Table  2-37,  clearly  indicate  a 
lack  of  correlation  between  the  brain  decarboxylase  levels  and  the  occur- 
rence of  convulsions;  e.g.,  3-deoxy pyridoxol  convulses  without  a  significant 
change  in  enzyme  activity,  whereas  c/j-methylpyridoxol  lowers  the  enzyme 
level  without  producing  convulsions.  Liver  dopa  decarboxylase  is  decreased 


570 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


moderately  in  mice  from  both  deficiency  and  injections  of  deoxypyridoxol 
(Dietrich  and  Shapiro,  1953  a).  Despite  the  lowered  levels  of  transaminase 
and  decarboxylase,  cysteine  desulfhydrase  (also  dependent  on  pyridoxal-P) 
is  not  affected  by  deoxypyridoxol,  due  to  the  fact  that  the  apodesulfhydrase 
has  little  affinity  for  deoxypyridoxol-P,  no  inhibition  of  the  binding  of  py- 
ridoxal-P being  observed  (Dietrich  and  Borries,  1956). 


600 


500 


400 


300 


200 


^     100 


LIVER 
KIDNEY 


TIME  (DAYS) 


Fig.  2-20.   Effects    of   deoxypyridoxol    given    intraperitoneally    at 

150  mg/kg/day  on  the  aspartate:  a-ketoglutarate  transaminase  of 

mouse  tissues.  (From  Shapiro  et  al.,  1953.) 


Serine  biosynthesis  from  formate  and  glycine,  involving  serine  transhy- 
droxymethylase,  is  dependent  on  pyridoxal-P.  The  incorporation  of  for- 
mate-C^*  into  serine  in  chick  liver  extracts  is  much  reduced  when  these 
are  obtained  from  deoxypyridoxol-treated  animals,  the  depression  being 
around  50%  and  reversible  with  pyridoxal-P  in  vitro  (Sakami,  1955).  Renal 


ANALOGS    OF    PYRIDOXAL 


571 


Table  2-37 
Effects  of  Pyridoxine  Analogs  on  Rat  Brain  Gltjtamate  Decarboxylase  ° 


Glutamate  decarboxylase 

Treatment 

Dose 

(mg/kg) 

Convulsions 

(//moles/g/hr) 

Endogenous 

+Pyridoxal-P 

Pjrridoxine-deficient 

Control 

36 

218 

Toxopyrimidine 

15 

4- 

40  (  +  11%) 

202  (-   7%) 

Control 

54 

235 

Toxopyrimidine 

25 

+ 

56  (+  4%) 

233  (-   1%) 

Normal  diet 

Control 

88 

274 

Toxopyrimidine 

700 

+ 

50  (-43%) 

266  (-  3%) 

3-Deoxypyridoxol 

150 

- 

56  (-36%) 

245  (-11%) 

cy-Methylpyridoxol 

200 

- 

31  (-65%) 

261  (-  5%) 

Pyridoxine-supplemented  * 

(100  //g%) 

Control 

70 

276 

3-DeoxypyridoxoI 

100 

+ 

72  (+  4%) 

252  (-  9%) 

4-Deoxypyridoxol 

100x7 

- 

71  (+  2%) 

296  (+   7%) 

5-Deoxypyridoxol 

100x7 

- 

82  (  +  18%) 

283  (+  3%) 

co-Methylpyridoxol 

100x7 

— 

30  (-57%) 

233  (-16%) 

°  From  Rosen  et  al.  (1960). 

''  The  analogs  were  injected  intraperitoneally  for  7  days,  except  for  3-deoxyp>Ti- 
doxol,  which  convulsed  the  animals  after  a  single  dose. 


glutaminase  is  also  reduced  by  deoxypyridoxol  at  21  days,  at  which  time 
dietary  deficiency  produces  no  change  (Beaton  and  Goodwin,  1955). 

It  is  impossible  to  evaluate  the  importance  of  these  changes  in  enzyme 
levels  for  the  toxic  effects  of  deoxypyridoxol.  The  toxic  effects  are  mani- 
fested principally  in  the  central  nervous  system,  skin,  and  hematopoietic 
system,  and  no  enzyme  determinations  during  deoxypyridoxol  treatment 
have  been  reported  in  any  of  these  tissues,  except  for  brain  glutamate  de- 
carboxylase. There  is  no  reason  to  attribute  the  toxic  effects  to  reduction 
of  transaminase  rather  than  to  the  possible  reduction  of  many  other  en- 
zymes dependent  on  pyridoxal-P,  most  of  which  have  never  been  examined 
in  this  connection.  It  is  more  likely  that  the  growth  inhibition  observed 


572  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

with  both  deficiency  and  deoxypyridoxol  is  related  to  lowered  transami- 
nase activity  and  protein  synthesis.  Analyses  of  total  brain  do  not,  of 
course,  necessarily  reflect  local  changes  in  amino  acid  metabolism,  and  the 
site  of  origin  of  the  convulsions  has  yet  to  be  determined.  It  must  be  re- 
membered that  transaminase,  glutamate  decarboxylase,  and  y-aminobu- 
tyrate  levels  vary  in  different  regions  of  the  central  nervous  system  and, 
furthermore,  the  functional  dependence  of  various  regions  on  pyridoxal- 
P-dependent  metabolism  must  also  vary.  The  displacement  of  pyridoxal-P 
from  apoenzymes  by  deoxypyridoxol-P  in  vivo  depends  on  several  factors: 
(1)  rate  of  penetration  of  the  analog  into  the  cells,  (2)  ability  of  the  tissue 
to  phosphorylate  the  analog,  and  (3)  relative  affinities  of  the  apoenzyme 
for  the  coenzyme  and  the  phosphorylated  analog.  Thus  one  might  for  sev- 
eral reasons  expect  the  pattern  of  enzyme  depression  from  deoxypyridoxol 
to  be  different  from  that  produced  by  simple  dietary  deficiency. 

Effects  on  Metabolism 

Kynurenate  is  a  normal  metabolite  of  tryptophan  in  the  rat,  but  in  pyri- 
doxine-deficient  animals  one  finds  kynurenine  and  xanthurenate  also.  The 
metabolic  pathway  involved  here  may  be  summarized  as  follows: 

kynurenate 

y 

Tryptophan  ->  formylkynurenine  ->  kynurenine  ->  3-OH-kynurenine  ->  xanthurenate 

\  " 

anthranilate 

This  is  an  interesting  situation  since  the  formation  of  all  three  of  these 
products  involves  pyridoxal-P  enzymes,  namely,  kynurenine  transaminase 
for  the  formation  of  kynurenate  and  xanthurenate,  and  kynureninase  for 
the  formation  of  anthranilate.  The  administration  of  deoxypyridoxal  to 
otherwise  normal  rats  produces  no  particular  effect,  but  if  tryptophan  is 
given  to  deoxypyridoxol-treated  animals  there  is  an  increase  in  the  appear- 
ance of  kynurenine  and  xanthurenate,  just  as  in  dietary  deficiency  (Porter 
et  al.,  1947).  The  rise  in  kynurenine  would  be  expected  because  two  of  its 
degradative  pathways  are  depressed  (including  the  normally  most  impor- 
tant one),  and  the  increase  in  xanthurenate  excretion  must  be  due  to  a 
diversion  of  the  metabolic  flow  through  the  remaining  pathway.  However, 
it  is  difficult  to  understand  why  xanthurenate  excretion  should  increase 
relative  to  kynurenate,  since  both  are  presumably  formed  with  the  same 
enzyme,  unless  a  rise  in  kynurenine  concentration  increases  relatively  more 
the  rate  of  the  xanthurenate  pathway.  It  would  be  interesting  to  know 
what  happens  to  the  level  of  3-hydroxykynurenine  during  deoxypyridoxol 
administration. 

Pyridoxine-deficient  rats  have  higher  blood  urea  than  normal  animals 
and  this  has  been  attributed  to  an  impaired  utilization  of  amino  acids, 


ANALOGS    OF    PYRIDOXAL  573 

since  it  is  not  of  renal  origin  and  hence  due  to  an  increased  urea  formation. 
Administration  of  100  //g/rat/day  of  deoxypyridoxol  for  28  days  apparently 
increases  urea  formation  in  liver  slices  in  both  deficient  and  pyridoxine-fed 
animals  (see  accompanying  tabulation),  although  the  authors  stated  that 


Pyridoxine 

Deoxypyridoxol 

Vurea 

4.32 

+ 

— 

3.38 

— 

+ 

4.76 

+ 

+ 

4.00 

deoxypyridoxol  lessens  the  deficiency  abnormalities  rather  than  accentuat- 
ing them  (Beaton  et  al.,  1954).  The  cycle  of  urea  formation  does  not  require 
pyridoxal  directly,  but  the  aspartate  to  condense  with  citrulline  must  be 
formed  by  transamination  reactions,  so  that  one  might  expect  impaired 
pyridoxal  function  to  depress  urea  formation  by  this  mechanism,  and  per- 
haps it  counteracts  to  some  extent  the  efl"ect  of  an  increased  supply  of  amino 
acids  for  catabolism. 

Convulsive  seizures  in  mice  are  produced  by  the  injection  of  4-methoxy- 
methylpyridoxol,  and  are  completely  prevented  by  pyridoxol  in  a  dose  3 
times  that  of  the  analog  (Gammon  et  al.,  1960).  The  convulsions  can  also 
be  prevented  by  prior  administration  of  y-aminobutyrate  but  not  by  any 
of  the  other  related  amino  acids  or  products  of  glutamate  metabolism  test- 
ed. Since  glutamate  decarboxylase  requires  pyridoxal-P,  the  most  obvious 
explanation  would  be  that  the  analog  reduces  brain  y-aminobutyrate,  there- 
by initiating  convulsions,  and  that  the  administration  of  y-aminobutyrate 
simply  restores  the  normal  level  or  prevents  depletion.  Others  have  suggest- 
ed that  certain  carbonyl  reagents,  such  as  thiosemicarbazide,  produce  sei- 
zures by  reducing  brain  y-aminobutyrate,  and  many  now  believe  that  ceU' 
tral  motor  activity  is  controlled  bj^  the  levels  of  such  amino  acids  and  their 
corresponding  amines.  Analyses  of  mouse  brains  obtained  during  convul-' 
sions  from  4-methoxymethylpyridoxol  and  other  analeptics  were  thus  made, 
and  it  was  found  that  y-aminobutyrate  drops  50-70%  when  the  analog  is 
used  but  shows  no  significant  change  when  the  convulsions  are  due  to  Me- 
trazol,  picrotoxin,  or  electroshock  (Gammon  et  al.,  1960;  Kamrin  and  Kam- 
rin,  1961).  Despite  the  coherence  of  these  observations  in  supporting  the 
role  of  y-aminobutyrate  in  antipyridoxine  convulsions,  one  additional  fact 
is  difficult  to  fit  in:  the  administration  of  pyridoxol,  which  blocks  the  sei- 
zures, does  not  alter  the  fall  in  brain  y-aminobutyrate.  It  has  also  been 
shown  that  administration  of  y-aminobutyrate  to  animals  with  seizures  pro- 
duced by  the  analog  does  not  reduce  the  seizures  although  the  y-amino- 


574  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

butyrate  content  of  the  brain  increases  (Purpura  et  al.,  1960).  These  dis- 
crepancies might  be  removed  if  one  could  determine  local  changes  in  y-ami- 
nobutyrate  in  the  central  nervous  system.  On  the  other  hand,  it  is  possible 
that  other  disturbances  are  more  pertinent  to  the  convulsive  state,  and,  as 
Rosen  et  al.  (1960)  have  pointed  out,  perhaps  more  thought  should  be  given 
to  the  interference  with  the  transport  of  amino  acids  into  the  brain  cells. 
Variations  in  the  metabolism  of  amino  acids  other  than  glutamate  and  the 
levels  of  physiologically  active  amines  have  not  been  studied.  In  this  con- 
nection it  is  interesting  that  Schrodt  et  al.  (1960)  administered  deoxypyri- 
doxol  to  two  patients  with  malignant  carcinoid  syndrome  at  doses  of  100- 
200  mg/day,  and  found  in  one  patient  a  fall  in  the  urinary  excretion  of  5- 
hydroxyindoleacetate,  which  is  the  primary  product  of  serotonin  metab- 
olism, along  with  symptoms  of  vitamin  Bg  deficiency. 

The  role  of  pyridoxal  in  lipid  metabolism  is  not  yet  clear  but  there  is 
some  evidence  in  animals  for  a  requirement  in  fatty  acid  synthesis.  When 
deoxypyridoxol  was  administered  to  seven  subjects  at  300  mg/day,  six  de- 
veloped symptoms  of  pyridoxine  deficiency  (Mueller  et  al.,  1959).  There  was 
a  general  decrease  in  the  polyunsaturated  fatty  acids  of  the  blood,  but  no 
change  is  phospholipids  or  cholesterol.  This  was  believed  to  be  evidence 
for  the  involvement  of  pyridoxine  in  maintaining  blood  fatty  acids  through 
participation  in  the  synthetic  reactions,  but  the  data  do  not  indicate  a  role 
in  the  interconversion  of  the  unsaturated  fatty  acids. 

Effects  on  Active  Transport 

Although  relatively  Httle  has  been  done  with  respect  to  the  actions  of 
pyridoxine  analogs,  there  is  accumulating  evidence  that  amino  acid  trans- 
port is  often  related  to  pyridoxal-P  function,  and  it  is  likely  that  some  of 
the  toxic  effects  of  the  analogs  will  be  explained  on  this  basis.  Part  of  the 
transport  of  amino  acids  across  the  intestinal  wall  is  active  and  is  inhibited 
by  deoxypyridoxol  (Fridhandler  and  Quastel,  1955).  A  41%  inhibition  of 
L-alanine  transport  was  observed  with  10  mM  deoxypyridoxol,  which  is 
certainly  a  very  high  concentration;  however,  pyridoxol  at  the  same  con- 
centration has  no  effect.  The  inhibition  is  not  antagonized  by  pyridoxol, 
pyridoxal,  or  pyridoxal-P,  which  may  indicate  that  phosphorylation  is  not 
rapid  enough  intracellularly  and  that  the  coenzyme  itself  cannot  penetrate, 
or  that  the  inhibition  is  not  an  antagonism  of  pyridoxine.  A  rather  disturb- 
ing fact  is  that  glucose  and  fructose  transport  is  also  inhibited  by  deoxy- 
pyridoxol to  about  the  same  extent  as  alanine  absorption,  so  that  this  is 
not  a  specific  effect  on  amino  acid  transport. 

The  transport  of  d-  and  L-methionine  across  rat  intestine  is  depressed 
by  deoxypyridoxol  injected  at  200-400  //g/day  (Jacobs,  1958;  Jacobs  and 
Hillman,  1958).  The  effect  appears  within  an  hour  after  intraperitoneal 
injection  and  can  be  abolished  by  injection  of  pyridoxol  (see  accompanying 


ANALOGS    OF    PYRIDOXAL  575 

tabulation)  (Jacobs  et  al.,  1960).  These  results  definitely  implicate  pyridoxal 
in  amino  acid  transport  but  do  not  prove  that  it  functions  directly  in  the 
transport  mechanism,  since  the  effect  could  be  an  indirect  one. 

Pyridoxol  Deoxypyridoxol  %  Change  in  transport 


-  -  —56 

+  —  +25 

+  -35 

+  +  +6 


The  uptake  of  glycine  by  ascites  carcinoma  cells  is  inhibited  by  5-25  mM 
deoxypyridoxol  (Christensen  et  al.,  1954).  Since  pyridoxine  deficiency  re- 
duces the  accumulating  ability  and  this  is  restored  by  pyridoxal  in  vitro, 
it  would  appear  that  pyridoxal  functions  here  in  some  manner,  although 
again  not  necessarily  in  the  membrane  transport  system. 

Effects  on   Growth 

Deoxypyridoxol  suppresses  the  growth  of  a  variety  of  microorganisms. 
Rabinowitz  and  Snell  (1953  a)  emphasized  that  sensitive  organisms  are 
those  requiring  an  exogenous  source  of  pyridoxine,  and  in  these  the  inhibi- 
tion can  be  counteracted  by  pyridoxine;  those  synthesizing  their  own  p\Ti- 
doxal  can  effectively  resist  the  analog.  The  situation  is  quite  complex,  how- 
ever, and  when  different  analogs  are  examined  a  marked  variability  in  sus- 
ceptibility is  found  (Rabinowitz  and  Snell,  1953  b).  For  example,  w-me- 
thylpyridoxol  is  inhibitory  to  yeast  but  not  at  all  to  Streptococcus  faecalis  or 
Lactobacillus  helveticus,  the  5-deoxypyridoxol  derivatives  being  the  most  ef- 
fective in  these  latter  organisms;  in  L.  helveticus,  only  5-deoxypyridoxol  is 
inhibitory,  4-deoxypyridoxol,  5-deoxypyridoxol,  and  5-deoxypyridoxamine 
being  without  action.  A  factor  that  is  very  important  in  determining  the 
susceptibility  of  bacteria  to  these  analogs  is  the  nature  of  the  exogenous 
amino  acids  supplied.  Streptococcus  faecalis  grows  well  if  all  amino  acids  are 
provided  even  though  pyridoxine  is  absent,  but  a  requirement  for  pyrido- 
xine and  a  sensitivity  to  analogs  are  created  by  restriction  of  the  amino 
acids  in  the  medium  (Olivard  and  Snell,  1955).  Under  certain  circumstances 
the  growth  can  be  limited  by  conversion  of  l-  to  D-alanine  by  alanine  race- 
mase,  which  involves  pyridoxal-P  and  is  quite  sensitive  to  5-deoxypyridoxol 
{K,  =  0.089  mM)  and  w-methylpyridoxol  {K^  =  0.53  mM).  The  inhibition 
of  growth  by  these  analogs  can  be  explained  on  the  basis  of  the  inhibition 
of  this  enzyme  under  these  conditions.  On  the  other  hand,  in  most  circum- 
stances the  inhibition  must  be  on  amino  acid  metabolism,  as  in  Vibrio 


576  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

cholera  where  alanine  and  aspartate  accumulate  when  growth  is  suppressed 
55%  by  0.53  vciM  deoxypyridoxol  (Chatterjee  and  Haider,  1960).  Perhaps 
the  first  pyridoxine  analogs  to  be  recognized  as  inhibitors  of  bacterial  growth 
were  the  tellurium  compounds  studied  by  Morgan,  Cooper,  and  their  col- 
leagues between  1923  and  1926.  Gulland  and  Farrar  (1944)  postulated  that 
2,4-dimethyl-c^c^o-telluropentane-3,5-dione  is  toxic  because  of  its  structural 
similarity  to  pyridoxine,  but  no  further  work  or  attempts  to  counteract  the 
inhibition  with  pyridoxine  have  come  to  my  attention.  An  in  vivo  effect  on 
bacteria  has  been  demonstrated  in  at  least  one  case  for  deoxypyridoxol, 
which  prolonged  the  survival  time  of  mice  infected  with  Toxoplasma  gondii 
from  5  days  to  12.1  days  when  it  was  incorported  into  the  diet  at  0.1%, 
although  some  toxic  symptoms  were  noted  (Summers,  1957).  Pyridoxine, 
can  counteract  both  the  beneficial  and  toxic  effects. 

Chick  embryogenesis  is  disturbed  by  deoxypyridoxol  and  other  analogs. 
When  1  mg  of  deoxypyridoxol  is  injected  into  eggs,  there  is  100%  mortality 
of  the  embryos  and  this  can  be  prevented  by  injecting  pyridoxine  (Cravens 
and  Snell,  1949).  However,  after  4  or  more  days  the  embryos  become  less 
sensitive,  and  although  higher  doses  are  toxic  they  are  not  counteracted 
by  pyridoxine.  Similar  effects  are  noted  for  4-methoxymethylpyridoxol  but 
it  is  at  least  25  times  more  toxic  than  deoxypyridoxol  to  the  early  chick 
embryo  (Karnofsky  et  al.,  1950).  Mammalian  fetal  development  is  also  dis- 
turbed by  deoxypyridoxol  —  fetuses  resorbed,  still-births,  and  abnormal 
young  —  but  administration  of  estrone  and  progesterone  together  prevents 
these  effects  and  pregnancy  is  maintained  in  the  majority  of  the  animals, 
indicating  that  the  action  of  the  analog  is  primarily  on  the  maternal  tissues 
rather  than  the  embryo  (Nelson,  1955). 

Some  of  the  observations  relative  to  the  inhibition  of  tumor  growth  by 
deoxypyridoxol  will  be  summarized  since  these  effects  are  interesting  in 
light  of  the  fairly  rapid  amino  acid  metabolism  in  tumors  and  the  generally 
low  levels  of  the  Bg  vitamers  in  solid  tumors.  Regression  of  mouse  lympho- 
sarcoma implants  with  deoxypyridoxol  was  achieved  by  Stoerk  (1947,  1950) 
when  the  animals  were  on  a  low-pyridoxine  diet,  but  there  was  also  loss 
of  body  weight,  suggesting  an  insufficiently  specific  inhibition.  The  fre- 
quency of  successful  fibrosarcoma  implants  in  rats  is  increased  by  pyri- 
doxol  and  decreased  by  deoxypyridoxol,  even  though  in  the  latter  case  no 
severe  deficiency  symptoms  are  observed  (Loefer,  1951).  On  the  basis  of 
these  findings,  Gellhorn  and  Jones  (1949)  gave  deoxypyridoxol  to  patients 
with  disseminated  lymphosarcoma  and  acute  leukemia  in  combination  with 
a  pyridoxine-deficient  diet.  Although  there  was  some  weight  loss  and  weak- 
ness, there  were  no  specific  signs  of  deficiency,  no  changes  in  tryptophan 
metabolism,  no  hematopoietic  depression,  and  no  retardation  in  the  growth 
of  lymphoid  tissue,  the  results  being  clinically  insignificant.  Deoxypyridoxol 
appears  to  be  reasonably  effective  in  suppressing  mammary  carcinoma  in 


ANALOGS    OF    PYEIDOXAL  577 

mice,  with  a  specific  regression  of  the  tumor  and  no  weight  loss  (Shapiro 
and  Gellhorn,  1951).  Human  carcinoma  cells  (Eagle's  KB  strain)  are  quite 
sensitive  to  deoxypyridoxol  in  tissue  culture,  0.08  mM  inhibiting  the  growth 
50%  (Smith  etal.,  1959).  It  appears  doubtful  that  deoxypyridoxol  is  suffi- 
ciently specific  as  a  carcinostatic  agent  but  its  use  in  conjunction  with  other 
inhibitors  remains  a  possibility,  especially  as  Doctor  (1959)  has  shown  that 
deoxypyridoxol  at  20  mg/kg/day  in  the  rat  has  no  effect  on  the  leucocyte 
count,  but  combined  with  a  moderately  effective  dose  of  aminopterin  exerts 
a  very  marked  suppression  of  the  leucocytes,  and  also  potentiates  the  action 
of  oxythiamine. 

Toxic  Effects  in  Whole  Animals 

Some  of  the  evidence  that  deoxypyridoxol  can  produce  rather  typical 
vitamin  Bg-deficiency  states  will  be  summarized  to  emphasize  that,  what- 
ever the  basic  biochemical  disturbances,  the  effects  are  primarily  related  to 
an  interference  with  formation  or  function  of  pyridoxal-P.  It  is  first  of  all 
quite  clear  that  the  doses  of  deoxypyridoxol  to  induce  toxic  reactions  must 
be  much  higher  when  the  animals  are  adequately  supplied  with  the  Bg  vi- 
tamers than  when  the  animals  are  subjected  to  a  dietary  deficiency,  and 
that  administration  of  pjTidoxol  can  overcome  the  toxic  reactions  produced 
by  the  analog.  In  general  the  responses  to  deoxypyridoxol  are  the  same 
as  in  pyridoxine  deficiency,  except  that  they  appear  earlier,  producing  an 
acute  deficiency  syndrome.  Thus  in  rats  and  mice  there  is  a  dermatitis 
characterized  by  acanthosis,  parakeratosis,  and  hyperkeratosis,  sometimes 
with  a  superimposed  infection,  which  is  similar  to  deficiency  dermatitis 
(Stoerk,  1950).  There  is  atrophy  and  degeneration  of  the  hematopoietic 
organs,  evidenced  by  decreases  in  thymus  and  spleen  weights,  and  these  are 
reflected  in  the  peripheral  blood  picture  (Mushett  et  al.,  1947).  The  nervous 
system  hyperirritability  leading  eventually  to  convulsions  has  been  men- 
tioned in  connection  with  the  metabolic  changes  in  the  brain. 

A  state  resembling  pyridoxine  deficiency  has  been  produced  in  man  by 
Mueller  and  Vilter  (1950).  Eight  individuals  on  a  pyridoxine-poor  diet  were 
injected  intramuscularly  with  60-150  mg/day.  Within  2-3  weeks  a  sebor- 
rheic dermatitis  appeared  around  the  eyes,  nose,  and  mouth,  with  simulta- 
neous glossitis  and  stomatitis.  These  symptoms  disappeared  in  2-3  days 
upon  administration  of  pyridoxol.  The  total  white  count  did  not  fall,  nor 
was  there  evidence  of  anemia,  but  the  lymphocytes  dropped  to  around  half 
the  initial  level.  Schrodt  et  al.  (1960)  in  their  two  carcinoid  patients  also 
observed  seborrheic  dermatitis  and  glossitis.  The  general  pharmacology  of 
analogs  and  inhibitors  of  pyridoxal  function  has  been  reviewed  by  Holtz 
and  Palm  (1964). 

Although  several  workers  have  stated  that  reactions  may  be  seen  in  acute 
deoxypyridoxol-treated  animals  which  are  not  seen  in  simple  dietary  de- 


578  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

ficiency,  there  appears  to  be  no  really  good  evidence  for  any  of  these  reac- 
tions being  unassociated  with  pyridoxal  function.  We  have  discussed  (p.  567) 
various  possible  reasons  for  different  effects  of  deoxypyridoxol  and  dietary 
deficiency  on  tissue  Bg  vitamer  levels  and,  hence,  on  metabolic  disturbances 
in  the  tissues. 

Toxopyrimidine 

This  substance  (4-amino-5-hydroxymethyl-2-methylpyrimidine),  which  is 
essentially  the  pyrimidine  portion  of  thiamine,  has  been  known  for  many 
years  to  produce  abnormal  motor  behavior  and  convulsions,  and  a  search 
for  antidotes  led  to  the  discovery  that  the  pyridoxine  group  is  specific  in 
preventing  the  toxic  reactions.  It  was  then  realized  that  toxopyrimidine 
bears  a  structural  resemblance  to  pyridoxal.  Makino  and  Koike  (1954  a,b) 
believed  that  toxopyrimidine  acts  in  the  phosphorylated  form  since  tyrosine 
decarboxylase  is  not  inhibited  by  toxopyrimidine  up  to  1  mM,  whereas  to- 
xopyrimidine-P  exerted  some  inhibition  at  0.001  mM  and  almost  complete 
inhibition  at  0.1  mM.  The  inhibition  is  competitive  with  respect  to  pyridox- 
al-P.  Haughton  and  King  (1958)  confirmed  this  inhibition  but  stated  that 
it  required  an  (I)/(C)  ratio  of  1000  to  get  50%  inhibition,  whereas  Makino 
and  Koike  found  around  50%  inhibition  with  a  ratio  near  3.  No  inhibition 
of  tryptophanase,  transaminase,  glutamate  decarboxylase,  or  arginine  de- 
carboxylase was  observed,  and  they  concluded  that  toxopyrimidine  is  not 
of  much  value  in  the  study  of  pyridoxal-P  enzymes.  The  failure  to  inhibit 
significantly  the  tryptophanase  of  E.  coli  was  also  reported  by  Wada  et 
al.  (1958).  McCormick  and  Snell  (1961)  found  no  inhibition  of  pyridoxal 
kinase  at  concentrations  of  0.01-0.1  mM  toxopyrimidine.  Rindi  and  Fer- 
rari (1959)  found  that  90-120  min  after  the  intraperitoneal  injection  of 
125  mg/kg  of  toxopyrimidine  in  pyridoxine-deficient  rats,  convulsions  hav- 
ing been  produced,  the  y-aminobutyrate  levels  in  the  brain  have  fallen 
some  23%,  although  glutamate  is  unchanged.  Administration  of  pyridoxa- 
mine  stops  the  convulsions  and  increases  brain  y-aminobutyrate.  This  dose 
of  toxopyrimidine  reduces  brain  glutamate  decarboxylase  20%  but  does 
not  significantly  alter  transaminase  activity  (Rindi  et  al.,  1959).  Again  pyri- 
doxamine  restores  activity.  We  have  already  noted  (Table  2-37)  that  toxo- 
pyrimidine can  reduce  brain  glutamate  decarboxylase  at  high  doses,  but  at 
lower  convulsive  doses  in  pyridoxine-deficient  animals  it  does  not.  It  would 
seem  that  if  toxopyrimidine  causes  convulsions  by  interfering  with  pyri- 
doxal function,  it  is  not  mediated  through  a  general  fall  in  y-aminobutyrate 
or  transaminase  activity,  and  the  status  of  the  mechanism  is  much  the 
same  as  for  deoxypyridoxol,  namely,  uncertain. 


ANALOGS    OF    PTEROYLGLUTAMATE    (FOLATE) 


579 


ANALOGS  OF  PTEROYLGLUTAMATE   (FOLATE) 

Tetrahydrofolate  functions  metabolically  in  the  transfer  of  C^  units  at 
the  oxidation  level  of  formaldehyde  or  formate,  and  thus  is  important  in 
the  biosynthesis  of  purines,  pyrimidines  (thymine),  certain  amino  acids  (ser- 
ine, histidine,  methionine),  choline,  and  other  biologically  important  sub- 
stances. Interference  with  its  function  leads  secondarily  to  a  depression  of 
nucleic  acid  and  protein  synthesis  and,  because  of  this,  to  a  general  sup- 
pression of  cellular  growth  and  multiplication.  The  possible  sites  of  block 
for  folate  analogs  may  be  summarized  as:  (1)  pathway  for  the  synthesis 
of  folate,  (2)  reduction  of  folate  to  tetrahydrofolate,  (3)  reactions  of  C^  unit 
transfer,  and  (4)  degradative  reactions  of  folate  and  its  derivatives.  The 
transport  of  folate  into  cells  should  probably  also  be  considered  as  a  pos- 
sible site  of  inhibition  but  little  about  this  process  is  known  at  the  present 


Folate     (F) 


(1) 


dihydrofolate  (FH,) 


(3) 


(2) 


S-formyl-FH^ 

(5) 
(8)  5-formimino-FHj 

(9)  (10) 

5,  10-methenyl-FH, 

(1)  folate  reductase 

(2)  dihydrofolate  reductase 

(3)  glutamate-FH4  transformylase 

(4)  serine  hydroxymethylase 

(5)  formimino  transferase 

(6)  10-formyl-FH4  deacylase 


—  tetrahydrofolate  (FHJ  — 


(4) 


-•^  lO-hydroxymethyl-  FH^ 


(6y\(7) 

10-formyl-FH4 


(11) 


5,  10- methylene- FH4 


(7)  tetrahydrofolate  formylase 

(8)  (cyclohydrolase) 

(9)  formimino- FH4  cyclodeaminase 

(10)  cyclohydrolase 

(11)  (cyclohydrolase) 

(12)  5, 10-methylene-FH4  dehydrogenase 


time.  A  number  of  substances  can  inhibit  the  synthesis  of  folate  in  micro- 
organisms —  e.g.,  the  sulfonamides  and  certain  pteridines  —  but  they  will 
not  be  treated  in  this  chapter.  There  is  no  evidence  that  the  important 
actions  of  folate  analogs  are  related  to  inhibition  of  degradative  reactions. 
We  shall  therefore  limit  the  subject  in  this  section  to  inhibitions  of  folate 
reduction  and  transformylation  reactions,  namely,  those  pathways  shown 
in  the  accompanying  diagram. 


580 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


H,N 


-NH 


v\     // 


COO 

I 

CH, 

I  " 
CHo 

I  "" 
CH— COO" 

I 


CONH 


Folate 


H,N 


OH       CHO 


5-Formyltetrahydrofolate 
(folinate,   citrovorum  factor) 


H,N^   /N.    /N^  ^CH, 


OH 

7-Methylfolate 


H,Nv.     M^      M 


H,N 


NH, 


Amethopterin 
(methotrexate) 


Pyrimethamine 
(Daraprim) 


2,  4,  7   Triamino-6- 
o-methylphenylpteridine 


ANALOGS   OF   PTEROYLGLUTAMATE    (fOLATE)  581 

The  most  thoroughly  studied  folate  analogs  are  aminopterin  (4-hydroxy 
group  replaced  by  amino  group)  and  amethopterin  (10-methylaminopterin) 
because  of  their  importance  in  the  chemotherapy  of  cancer.  These  substances 
produce  states  of  folate  deficiency  in  all  types  of  organism  and,  in  most 
instances,  the  symptoms  of  this  deficiency  can  be  prevented  by  providing 
tetrahydrofolate  or  5-formyltetrahydrofolate,  but  not  with  folate,  indicating 
that  the  site  of  block  lies  somewhere  in  the  pathway  of  the  transformation 
of  folate  to  its  coenzymically  active  forms.  Although  this  block  produces  a 
general  disturbance  in  tetrahydrofolate  function  and,  it  appears,  nucleic 
acid  and  protein  synthesis,  there  are  two  reasons  for  some  degree  of  speci- 
ficity. In  the  first  place,  the  syntheses  of  the  various  substances  requiring 
Ci  units  from  folate  coenzymes  are  not  necessarily  all  depressed  equally, 
since  it  is  a  general  rule  that  a  lowering  of  the  concentration  of  some  sub- 
stance from  which  several  pathways  lead  will  produce  varying  effects  on 
these  pathways,  depending  mainly  on  the  nature  of  the  enzymes  involved 
and  the  supply  of  reactants  for  each  pathway  (in  this  case  C^  unit  accep- 
tors). In  the  second  place,  those  cells  or  tissues  with  the  highest  rates  of 
synthesis  and  dependent  on  these  synthetic  reactions  for  growth  or  multi- 
plication wiU  be  most  adversely  affected  by  the  folate  analogs.  Here  it  is 
not  a  matter  of  degree  of  functional  activity  but  of  proliferation;  the  heart 
is  not  readily  affected  by  these  analogs  whereas  the  hematopoietic  system  is. 
We  shall  not  discuss  the  more  biological  aspects  of  the  actions  of  these 
analogs,  since  this  is  covered  adequately  in  a  number  of  books  and  reviews 
(e.g.  Holland,  1961;  Delmonte  and  Jukes,  1962),  but  confine  attention  to 
the  basic  enzyme  and  metabolic  effects. 

Inhibition   of  the    Reduction   of  Folate   to   Tetrahydrofolate 

This  reduction  occurs  in  two  steps  and  in  most  instances  it  appears  that 
each  step  is  catalyzed  by  a  specific  enzyme,  but  possibly  in  other  cases  a 
single  enzyme  is  responsible.  Most  assays  of  folate  reduction  for  analog 
inhibition  have  involved  determination  of  the  formation  of  either  tetrahy- 
drofolate or  folinate,  or  the  disappearance  of  folate,  and  it  is  difficult  to 
differentiate  between  the  two  steps  with  regard  to  inhibition.  In  some  re- 
ports the  term  "folate  reductase"  is  applied  to  the  over-all  reaction.  One 
thing  is  certain:  Dihydrofolate  reductase,  which  has  been  better  purified 
and  more  thorougly  studied  than  folate  reductase,  is  very  potently  inhibited 
by  aminopterin  and  amethopterin.  The  enzyme  from  chicken  liver  is  inhib- 
ited 74%  by  0.000053  mM  aminopterin  (Futterman,  1957)  and  from  hu- 
man leukemic  leucocytes  64%  by  0.00001  mM  (Bertino  et  al,  1960),  in 
both  cases  the  substrate  being  10,000-fold  or  more  in  excess  of  the  analog. 
Osborn  et  al.  (1958)  calculated  the  K's  for  aminopterin  and  amethopterin 
to  be  0.000001  mM  and  0.0000023  mM,  respectively,  using  the  chicken  liver 
enzyme,  Blakley  and  McDougall  (1961)  reported  a  value  of  0.0000024  mM 


582  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

for  aminopterin  and  the  enzyme  from  Streptococcus  faecalis,  and  Nath  and 
Greenberg  (1962)  gave  K,  as  0.0000023  mM  for  amethopterin  and  the  calf 
thymus  enzyme.  The  inhibitions  of  the  full  reduction  of  folate  to  tetrahy- 
drofolate  by  aminopterin  and  amethopterin  are  very  similar  (Futterman, 
1957;  Futterman  and  Silverman,  1957;  Zakrzewski  and  Nichol,  1958;  Silber 
et  al.,  1962),  and  point  to  the  dihydrofolate  reductase  as  being  the  more 
sensitive  enzyme.  Furthermore,  the  formation  of  folinate  (citrovorum  fac- 
tor) from  folate  is  very  potently  inhibited  in  rat  liver  slices  (Nichol  and 
Welch,  1950),  Lactobacillus  casei  and  Streptococcus  faecalis  (Hendlin  et  al., 
1953),  mouse  leukemic  cells  (Nichol,  1954),  and  chicken  liver  extracts 
(Doctor,  1958).  The  most  potent  pteridine  analog  was  found  by  Doctor 
(1958)  to  be  2,4,7-triamino-6-o-methylphenylpteridine,  although  it  is  not 
as  potent  as  aminopterin,  and  he  showed  that  the  site  of  inhibition  is  pre- 
vious to  tetrahydrofolate.  There  is  thus  much  evidence  that  folate  reduc- 
tion is  blocked  by  low  concentrations  of  these  analogs,  and  it  is  generally 
agreed  that  this  must  be  the  primary  mechanism  by  which  folate  defi- 
ciency and  growth  depression  are  produced. 

The  inhibitions  of  folate  reduction  by  aminopterin  and  amethopterin  have 
been  reported  to  be  noncompetitive  by  several  investigators,  but  it  is  very 
likely  that  the  inhibitions  are  truly  competitive,  this  being  obscured  by  the 
much  greater  affinity  of  the  enzyme  for  the  analog  than  for  dihydrofolate. 
In  no  case  has  the  rate  of  inhibition  in  the  presence  of  varying  concentra- 
tions of  substrate  been  determined,  but  one  might  predict  that  the  competi- 
tive nature  of  the  inhibition  would  be  demonstrated  in  this  way.  Once  the 
enzyme  is  inhibited,  it  is  very  difficult  to  recover  the  activity  because  the 
rate  of  dissociation  of  the  analog  from  the  enzyme  is  extremely  slow.  In 
other  words,  this  is  an  example  of  pseudoirreversible  inhibition  obeying  mu- 
tual depletion  kinetics.  This  was  shown  by  Peters  and  Greenberg  (1959)  on  a 
sheep  liver  folate  reductase,  the  inhibition  at  constant  analog  concentration 
being  dependent  on  the  enzyme  concentration.  It  is  thus  possible  to  titrate 
this  enzyme  in  tissues  or  extracts.  This  has  been  well  discussed  by  Werk- 
heiser  (1961),  who  also  showed  that  the  amount  of  analog  bound  by  rat 
liver  supernates  is  equivalent  to  the  amount  required  to  inhibit  folate  re- 
duction; in  other  words,  the  tightly  bound  analog  seems  to  be  combined 
only  with  dihydrofolate  reductase.  If  rats  are  injected  with  amethopterin, 
the  supernatant  fraction  of  the  liver  contains  most  of  the  analog  and  only 
10-15%  of  this  is  lost  during  dialysis  for  6  days  (Werkheiser,  1959).  The 
amount  of  amethopterin  or  aminopterin  to  inhibit  completely  the  reductase 
in  liver  extracts  in  0.56  //g/g  of  liver,  and  the  supernates  of  livers  from 
analog-treated  rats  contain  0.52  //g/g  of  tissue.  The  reductase  is  the  only 
protein  binding  these  analogs  significantly  in  chicken  liver  homogenates 
(Schrecker  and  Huennekens,  1964).  Fountain  et  al.  (1953)  had  found  that 
there  is  a  remarkable  retention  of  amethopterin  in  the  tissues  of  mice,  the 


ANALOGS    OF    PTEROYLGLUTAMATE    (fOLATE)  583 

concentration  in  the  liver  remaining  approximately  constant  for  at  least  3 
weeks  after  a  single  intravenous  dose.  It  is  noteworthy  that  some  tissues, 
such  as  the  lung  and  spleen,  do  not  pick  up  much  of  the  analog,  and  that 
the  kidney  loses  the  analog  relatively  more  rapidly  than  the  liver.  Werk- 
heiser  (1960)  likewise  found  that  the  liver  retains  aminopterin  for  long  pe- 
riods, while  the  intestine  does  not.  In  mice  given  a  lethal  dose  of  aminop- 
terin 1  hr  after  a  protective  injection  of  folate,  the  liver  folate  reductase 
activity  is  depressed  96%  after  24  hr  and  remains  at  this  level  for  7  days, 
following  which  there  is  a  slow  recovery  (Werkheiser,  1962).  The  intestinal 
enzyme  is  similarly  inhibited  but  recovers  faster.  The  loss  of  aminopterin 
from  the  intestine  is  characterized  by  a  half-life  of  60  hr,  but  the  liver  shows 
two  components  with  half-lives  of  60  hr  and  90  days,  respectively.  It  was 
suggested  that  rapidly  proliferating  cells  are  dependent  on  folate  reductase 
activity  and  that  the  disappearance  of  the  inhibition  is  faster  in  such  tissue 
because  of  the  more  rapid  turnover  of  cells;  in  other  words,  the  60  hr  com- 
ponent would  arise  from  proliferating  tissue  while  the  90  day  component 
would  relate  to  nonproliferating  tissue.  If  this  is  the  case,  the  binding  of 
aminopterin  to  the  enzyme  in  vivo  must  be  essentially  irreversible.  These 
results  all  point  to  a  very  high  degree  of  specificity  in  the  binding  and  inhi- 
bition, and  confirm  the  major  site  of  attack  as  being  on  folate  reduction. 
Further  evidence  comes  from  the  reduced  urinary  folinate  levels  in  rats  on 
25  //g/day  of  aminopterin  (Nichol  and  Welch,  1950).  Less  direct  evidence 
is  provided  by  Nichol  (1954)  and  Broquist  et  al.  (1953),  who  showed  that 
resistant  streptococci  or  leukemic  cells  have  a  much  greater  ability  to  pro- 
duce folinate  from  folate  than  do  normal  cells.  It  is  possible  that  this  in- 
creased activity  allows  enough  active  folinate  to  be  formed  to  enable  the 
cells  to  grow  and  multiply  in  the  presence  of  the  analogs,  but  it  is  probable 
that  this  is  not  the  only  mechanism  of  resistance  to  these  agents. 

The  nature  of  the  binding  of  these  folate  analogs  to  the  reductase  has 
not  been  fully  elucidated,  but  Zakrzewski  (1963)  has  determined  the  ther- 
modynamic characteristics  for  the  dissociation  of  the  EI  complexes  (see  ac- 
companying tabulation).  The  inhibitions  by  the  substituted  pteridines  are 
competitive.  The  K^  values  for  aminopterin  were  taken  from  Werkheiser 


Changes  for 

EI  dissociation 

Inhibitor 

Ki  (mM)          AH 

AFo 

AS 

2,6-Diaminopurine 

0.0018            +6.0 

+  7.9 

-6.4 

2,4-Diamino-6-methylpteridine 

0.0018           +5.0 

+  7.9 

-9.5 

2,4-Diamino-6-formylpteridine 

0.0081            +4.1 

+6.9 

-9.5 

2,4-Diamino-6-hydroxypteridine 

0.37                +1.8 

+4.7 

-9.6 

Aminopterin 

10  '-10-8       +11.6 

+  13.7-15.1     - 

-7.0-11.7 

584  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

(1961)  and  are  uncertain  because  it  is  not  known  whether  competitive  or 
noncompetitive  inhibition  occurs;  it  may  be  noted  that  the  values  are  one- 
hundredth  to  one-tenth  those  given  by  previous  workers.  The  pyrimidine 
amino  groups  are  essential  for  tight  binding  and  the  pyrazine  ring  presum- 
ably does  not  participate  in  the  binding.  A  binding  mechanism  was  proposed 
in  which  emphasis  is  placed  on  the  tautomeric  state  of  the  analogs  and  the 
formation  of  hydrogen  bonds  between  the  ring  nitrogens  and  the  amino 
groups  with  the  enzyme,  the  replacement  of  the  4-OH  group  of  folate  with 
an  amino  group  favoring  greater  hydrogen  bonding.  It  may  also  be  ob- 
served that  the  benzoylglutamate  portion  of  the  molecule  must  contribute 
around  6-7  kcal/mole  binding  energy. 

Another  analog  with  less  obvious  structural  similarity  to  folate  is  pyri- 
methamine (Daraprim),  an  antimalarial  drug  that  in  chronic  dosage  pro- 
duces folate  deficiency  in  bacteria  and  animals  (Wood  and  Hitchings,  1959  a; 
Hitchings,  1960).  Pyrimethamine,  like  the  analogs  previously  discussed, 
inhibits  the  reduction  of  folate,  and  does  this  at  a  concentration  equivalent 
to  that  required  for  growth  inhibition.  A  35%  inhibition  of  folinate  forma- 
tion is  caused  in  extracts  of  S.  faecalis  by  0.000012  mM  pyrimethamine, 
so  that  its  potency  is  comparable  to  that  of  aminopterin.  There  are  no  ef- 
fects on  the  biosynthesis  or  assimilation  of  folate.  It  is  believed  that  the 
antimalarial  action  is  due  to  the  tighter  binding  of  the  drug  to  the  plasmo- 
dial  folate  reductase  than  to  the  host  enzyme.  However,  the  uptake  of 
pyrimethamine  by  bacterial  cells  is  unique  inasmuch  as  it  is  inhibited 
strongly  by  glucose,  whereas  the  uptake  of  aminopterin  is  augmented  by 
glucose  (see  accompanying  tabulation)  (Wood  and  Hitchings,  1959  b).  Fur- 


Analog  taken 

up   (cpm) 

Analog 

No  glucose 

Glucose 

Pyrimethamine 
Aminopterin 

51.6 
55 

0.7 
523 

thermore,  aminopterin  uptake  is  increased  by  a  rise  in  temperature,  whereas 
less  pyrimethamine  appears  in  the  cells  at  higher  temperatures.  Despite 
the  apparent  similarity  of  site  of  action  of  these  two  analogs,  there  is  some 
basic  difference  in  the  movement  or  disposition  of  the  materials  in  the  cells. 

Effects   on    Synthetic   Processes    Mediated    by  Tetrahydrofolate 

A  block  in  the  reduction  of  folate  would  be  expected  to  depress  the  C^ 
unit  transfers  and  the  synthesis  of  nucleic  acids  and  proteins  as  long  as 
there  is  no  supply  of  tetrahydrofolate  or  folinate.  The  analogs  in  addition 


ANALOGS   OF   PTEROYLGLUTAMATE    (FOLATE)  585 

might  inhibit  directly  the  reactions  in  which  tetrahydrofolate  functions. 
There  is  essentially  no  information  on  this  second  possibility.  Cyclohydro- 
lase  is  inhibited  weakly  by  amethopterin  (43%  at  0.5  mM)  and  aminopterin 
(69%  at  0.5  mM)  (Tabor  and  Wyngarden,  1959),  and  formyltetrahydro- 
folate  synthetase  is  even  more  weakly  inhibited  by  several  analogs  ( Jaenicke 
and  Erode,  1961;  Whiteley  et  al.,  1959).  The  many  other  reactions  involved 
have  never  been  examined  for  inhibition  by  analogs.  All  one  can  say  at  the 
present  time  is  that  the  known  inhibitions  on  folate  reduction  are  sufficient 
to  explain  most  or  all  of  the  effects  of  these  analogs. 

Some  examples  of  the  inhibition  of  syntheses  will  be  mentioned  to  illus- 
trate the  nature  of  the  metabolic  actions  of  these  analogs.  Aminopterin  and 
amethopterin  invariably  depress  the  incorporation  of  formate-C^*  into  pu- 
rines and  nucleic  acids;  this  has  been  shown  in  rabbit  bone  marrow  (Totter 
and  Best,  1955),  leukemic  spleen  extracts  (Balis  and  Dancis,  1955),  and  the 
whole  animal  (Skipper  et  al.,  1950).  It  appears  that  thymine  synthesis  is 
more  sensitive  than  purine  synthesis  to  inhibition  by  these  analogs.  In  bone 
marrow  0.0021  mM  aminopterin  inhibits  incorporation  into  thymine  72% 
but  into  adenine  or  guanine  only  22%.  It  would  be  interesting  to  know 
what  the  effect  on  adenine  nucleotides  is,  but  little  is  known.  Aminopterin 
elevates  ATP  in  liver,  has  no  effect  on  spleen  or  muscle  ATP,  and  reduces 
tumor  ATP  (Zahl  and  Albaum,  1955).  Indeed,  the  total  adenine  nucleotides 
in  the  sarcoma  decrease.  One  might  expect  the  effects  to  depend  on  the 
relative  rates  of  adenine  synthesis  and  turnover  in  the  tissues.  Aminopterin 
lowers  liver  NAD  levels  —  39%  fall  at  60  //g/day  and  56%  fall  at  100  //g/day 
—  and  if  this  occurs  throughout  the  body  it  could  be  an  important  conse- 
quence of  interference  with  folate  metabolism  (Strength  et  al.,  1954).  The 
in  vitro  depression  of  respiration  by  aminopterin  is  not  completely  reversed 
by  added  NAD,  and  it  is  very  possible  that  other  coenzymes  (e.g.  NADP, 
coenzyme  A.  or  FAD)  are  decreased. 

The  interconversion  of  glycine  and  serine  is  inhibited  by  aminopterin 
when  only  folate  is  supplied,  but  the  activity  is  restored  with  folinate  (Blak- 
ley,  1954).  Tetrahydroaminopterin  does  not  inhibit  serine  synthesis  when 
folinate  is  provided,  but  2-deaminofolate  inhibits  some  55%  at  0.75  mM  in 
rabbit  liver  extracts  (Blakley,  1957).  The  incorporation  of  formate-C^^  into 
lymphoma  proteins  is  inhibited  72%  by  0.0073  mM  amethopterin,  but  little 
effect  is  observed  in  normal  liver  (Williams  et  al.,  1955).  In  general  these 
analogs  block  nucleic  acid  synthesis  more  than  protein  synthesis,  but  this 
may  vary  a  good  deal  from  one  tissue  to  another,  or  one  organism  to  an- 
other. Most  of  the  inhibitions  in  nucleic  acid  and  protein  synthesis  have  been 
attributed  solely  to  defects  in  the  formation  of  the  constituent  units,  and 
little  consideration  has  been  given  to  other  possible  contributing  factors, 
such  as  lowered  levels  of  various  coenzjTnes  with  impairment  of  oxidative 
and  phosphorylative  reactions. 


586  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Effects  on  Certain  Enzymes  Unrelated  to  Q  Transfer 

Brief  mention  should  be  made  of  the  early  demonstrations  that  7-methyI- 
folate  is  inhibitory  to  dopa  decarboxylase  and  that  this  is  reversible  with 
folate  at  10-100  times  the  analog  concentration  (Martin  and  Beiler,  1947). 
Pteroylaspartate  and  7-methylpteroate  are  less  potent  inhibitors,  but  the 
difference  is  not  great,  so  that  the  glutamate  portion  is  not  very  important 
for  the  binding  (Martin  and  Beiler,  1948).  Tyrosine  decarboxylase  is  not 
inhibited  by  0.67  mM  7-methylfolate  whereas  0.067  mM  inhibits  dopa  de- 
carboxylase 25%.  The  mechanism  of  this  inhibition,  the  role  of  folate  in 
decarboxylase  activity,  and  the  effects  of  the  more  commonly  used  folate 
analogs  are  all  unknown.  The  decarboxylase  inhibition  prompted  a  study 
of  the  effects  of  7-methylfolate  on  blood  pressure  (Martin  et  al.,  1947),  and 
it  was  found  that  5  mg/kg  in  the  dog  depresses  the  blood  pressure  quite 
significantly  and  for  an  extended  period  of  time,  but  whether  this  is  related 
to  decarboxylase  inhibition  is  problematical. 

The  inhibition  of  acetyl  transfer  by  amethopterin  (K,  =  0.032  mM)  in 
pigeon  liver  extracts  is  interesting  because  it  represents  another  possible 
site  of  action  for  this  analog,  even  though  it  is  obviously  much  less  potent 
here  than  on  folate  reduction  (Jacobson,  1960).  The  inhibition  is  not  coun- 
teracted by  folate,  tetrahydrofolate,  or  folinate,  and  is  competitive  with 
the  acetyl  donor  (p-nitroacetanilide)  but  noncompetitive  with  the  acetyl 
acceptor  (aniline  or  sulfanilamide).  These  results  do  not  implicate  a  folate 
compound  in  the  acetylation  reaction  —  indeed,  folate  and  folinate  are 
weak  inhibitors,  and  the  mechanism  is  more  likely  a  simple  binding  to  the 
substrate  site.  The  10-methyl  group  is  important  since  aminopterin  is  only 
one-tenth,  or  less,  as  inhibitory  as  amethopterin.  Evidence  for  inhibition 
of  acetylation  in  vivo  is  the  higher  sulfanilamide  level  and  the  lower  acetyl- 
sulfanilamide  level  in  rabbit  plasma  of  animals  treated  with  amethopterin 
(Johnson  et  al.,  1958). 

ANALOGS   OF   OTHER  VITAMINS,  COENZYMES, 
AND  THEIR   COMPONENTS 

There  has  been  a  great  deal  of  work  on  the  growth  inhibitions  produced 
by  numerous  analogs  of  pantothenate,  biotin,  cobalamin,  and  other  meta- 
bolically  necessary  cofactors,  but  relatively  few  reports  on  enzyme  inhibi- 
tions are  available.  However,  some  of  this  isolated  work  on  enzyme  systems 
is  interesting  in  itself  and  perhaps  some  reference  to  it  will  stimulate  further 
study. 

Pantothenate  is  required  by  many  microorganisms  and  animals  because 
it  is  a  component  of  coenzyme  A,  the  biosynthetic  pathway  being: 

Pantothenate  ->  pantothenylcysteine  ->•  pantetheine  ->•  CoA 


ANALOGS    OF    OTHER    VITAMINS  587 

Mcllwain  (1945)  found  that  pantothenate  analogs,  such  as  pantoyltaurine, 
do  not  displace  bound  pantothenate  from  bacterial  cells,  and  concluded 
that  these  analogs  are  bacteriostatic  because  they  block  the  formation  of 
the  metabolically  active  form  of  pantothenate  (which  was  not  known  at 
that  time).  Furthermore,  as  Martin  et  al.  (1950)  showed,  the  analogs  in  gen- 
eral do  not  interfere  in  the  reactions  involving  CoA.  Pantoyltaurine  and 
other  analogs  do  not  inhibit  brain  choline  acetylase,  even  at  concentrations 
around  5  vaM.  One  analog,  salicyloyl-/?-alanide,  does  inhibit  this  enzyme 
(20%  at  0.47  mM  and  100%  at  4.7  mM),  but  since  neither  pantothenate 
nor  CoA  reverses  this  inhibition  it  is  doubtful  if  it  is  specific.  Most  effective 
analogs  thus  seem  to  block  the  pathway  of  pantothenate  -^  CoA,  and  no 
known  direct  antagonists  of  CoA  are  known.  Pantoylaminoethanethiol  in- 
hibits the  synthesis  of  CoA  from  pantetheine  (50%  at  a  ratio  of  analog  to 
pantetheine  of  13)  and  thus  inhibits  sulfonamide  acetylation  in  liver  ex- 
tracts provided  with  pantetheine  (Boxer  et  al.,  1955).  One  cannot  help  but 
wonder  in  some  of  these  instances  if  abnormal  CoA  analogs  are  formed, 

CH, 
I 
HOCH,—  C—  CHOH—  CONH— CH,CH,—  COO' 

"■      I 
CH3 

Pantothenate 


CH3 

I 

HOCH2—  C— CHOH— CONH  -  CHoCH,—  SO: 

I  -       2  3 

CH, 


Pantoyltaurine 


CH3 
I 
HOCH2— C-CHOH-CONH-CH2CHi-CONH-CH2CH2SH 

CH3 

Pantetheine 


CH, 

I 


OH 


HOCH2—C  — CHOH— CONH- CH2CH2SH                      /         A— CONH-CH2CH2— COO" 
CH3  \ / 

Pantoylaminoethanethiol  Salicyloyl-/3-alanide 


588 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


rather  than  a  simple  inhibition  of  the  biosynthetic  pathway.  It  is  possible 
to  go  back  farther  and  inhibit  the  synthesis  of  pantothenate  by  analogs  of 
pantoate  or /5-alanine.  For  example,  2,3-dichloroisobutyrate  blocks  the  cou- 
pling of  these  two  components  of  pantothenate,  competitive  with  pantoate 
and  uncompetitive  with  /5-alanine  (K^  =  1.4-6.4  mM)  (Hilton,  1958). 

Biotin  functions  in  several  metabolic  pathways  (synthesis  of  aspartate 
and  higher  fatty  acids,  and  COg  fixation)  and  in  certain  organisms  can  be 
formed  from  desthiobiotin.  2-Oxo-4-imidazolidinecaproate  (desmethyldes- 
thiobiotin)  inhibits  the  growth  of  E.  coli  by  competing  with  desthiobiotin 
for  an  enzyme  involved  in  biotin  synthesis,  and  this  inhibition  is  competi- 
tive (Rogers  and  Shive,  1947).  Biotin  by  an  unknown  mechanism  stimulates 
fermentation  in  biotin-deficient  yeast  and  this  is  inhibited  by  homooxy- 
biotin,  oxybiotinsulfonate  (COO-  group  replaced  by  SOg"  group  in  oxybio- 
tin),  and  7-(3,4-ureylenecyclohexyl)butyrate  when  the  analogs  are  added 

n  O 


HN" 


■NH 


HN' 


■^NH 


Biotin 


■(CH2)^COO" 


H3C         CH.— (CH2)4— COO 
Desthiobiotin 


O 

II 

c 


CH.r-(CH2)  — COO' 


-(CH2)5— COO' 


Desmethyldes- 
thiobiotin 


Homobiotin 


HN"^      NH 


HN^       NH 


(CH,)  — COO' 


/         V(CH2)3— COO' 


Oxybiotin 


y-(3,  4-Ureylenecyclo- 
hexyl)butyrate 


ANALOGS    OF    OTHER    VITAMINS  589 

before  the  biotin,  but  not  afterward  (Axelrod  et  al.,  1948).  This  could  mean 
that  these  analogs  cannot  displace  biotin  once  it  is  bound  or  that  they  in- 
hibit in  some  way  the  formation  of  an  active  form  of  biotin.  Biotin  is  inac- 
tivated by  kidney  slices,  possibly  by  removal  of  fragments  from  the  side 
chain,  by  biotin  oxidase,  and  this  enzyme  is  inhibited  by  several  analogs 
competitively  (see  accompanying  tabulation)  (Baxter  and  Quastel,  1953). 
There  is  thus  the  possibility  that  some  analogs  can  conserve  biotin  in  the 
tissues  as  well  as  inhibit  its  svnthesis  or  function. 


4      1  Concentration      ,  .      ,     , ,  ,  .     . 

Analog  (Analog)/ (biotin)       %  Inhibition 

{raM ) 


DL-Homobiotin 

1.65 

10 

96 

DL-Desthiobiotin 

1.63 

10 

96 

L-Biotin 

0.4 

10 

82 

DL-Biotindiaminecarboxylate 

0.4 

10 

82 

D-Biotinoldiamine 

0.4 

10 

45 

D-Biotinol 

0.41 

10 

24 

2.1 

50 

61 

D-Biotinsulfone 

0.41 

10 

16 

Analogs  of  cyanocobalamin  (vitamin  B^g)  have  not  been  extensively  stud- 
ied because  of  the  complexity  of  the  structure.  5,6-Dimethylbenzimidazole 
is  a  component  of  cyanocobalamin  and  l,2-dimethyl-4,5-diaminobenzene  is 
a  precursor  in  the  synthesis.  Analogs  of  these  substances  are  often  inhibi- 
tory to  bacterial  growth  and  the  biosynthesis  of  vitamin  B^g.  1,2-Dichloro- 
4,5-diaminobenzene  inhibits  the  synthesis  of  vitamin  B^g  in  bacteria  and  the 


HaC^^^^^^X^  ^  Clv^/^^^NHa 


,iU, 


H,C^    ^^^    ^N^  Cr     ^^    ^NHa 


5,  6-Dimethylbenzimi-  1,  2-Dichloro-4,  5- 

dazole  diaminobenzene 

growth  of  those  bacteria  requiring  vitamin  B^g  (WooUey  and  Pringle,  1951). 
Vitamin  B^g  is  unable  to  counteract  these  inhibitions.  Although  5,6-dimeth- 
ylbenzimidazole  can  be  used  by  the  rat  to  form  vitamin  B^g,  this  sub- 
stance is  inhibitory  to  Lactobacillus  lactis,  and  like  a  number  of  analogs,  is 
able  to  inhibit  the  synthesis  of  vitamin  B^g  in  these  bacteria  (Hendlin  and 
Soars,  1951).  These  inhibitions  do  not  appear  to  be  competitive  with  vi- 


590  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

tamin  B^g,  but  the  inhibition  by  l,2-diamino-4,5-dimethylbenzene  is  com- 
petitive. The  only  instance  of  the  inhibition  of  vitamin  B^g  function  is  in 
the  synthesis  of  methionine  from  serine  in  E.  coli  where  the  methylamide, 
ethylamide,  and  aniHde  analogs  of  cyanocobalamin  inhibit  competitively 
with  respect  to  vitamin  B^j  (Guest,  1960).  These  analogs  also  inhibit  the 
growth  of  organisms  that  require  methionine  or  vitamin  B^g,  but  do  not 
when  methionine  is  supplied.  The  inhibitions  are  reasonably  potent,  61% 
depression  being  given  by  0.029  mM  of  the  anilide  derivative  when  cyano- 
cobalamin is  0.000032  mM,  but  the  analogs  are  obviously  bound  less  tightly 
than  the  cyanocobalamin  to  the  enzyme  involved.  Hydroxocobalamin  and 
cyanocobalamin,  which  are  analogs  of  cobalamin  coenzyme,  inhibit  potent- 
ly and  competitively  the  diol  dehydrase  from  Aerobacter,  but  once  inhibition 
occurs  it  cannot  be  reversed  by  either  dialysis  or  the  coenzyme  (Lee  and 
Abeles,  1963). 

Lipoate  functions  in  acyl  transfer  during  the  oxidation  of  keto  acids  and 
this  is  inhibited  by  6-ethyl-8-mercaptooctanoate,  an  analog  of  6-acetyl-6,8- 
dimercaptooctanoate  (a  functional  form  of  lipoate)  (Albrecht,  1957).  This 
analog  does  not  inhibit  the  anaerobic  decarboxylation  of  pyruvate  in  ex- 
tracts from  E.  coli  but  inhibits  pyruvate  oxidation.  The  phosphotransacetyl- 
ase  reaction  and  the  formation  of  acetyllipoate  are  inhibited. 


MrSCELLANEOUS  ANALOG  INHIBfTiONS 

There  are  a  number  of  reports  of  inhibitions  by  analogs  that  do  not  readily 
fall  into  any  general  classification.  Some  of  these  have  been  put  into  Table 
2-38  in  order  to  illustrate  further  the  various  types  of  analog,  although  in 
some  cases  it  is  not  quite  certain  whether  the  inhibitor  should  be  considered 
as  an  analog  or  not.  In  most  instances  the  inhibitions  are  competitive,  but 
in  others  it  may  be  noncompetitive  or  mixed,  indicating  that  the  inhibitions 
do  not  all  involve  a  simple  competition  between  the  analog  and  the  sub- 
strate for  an  enzyme  site.  Unfortunately,  most  of  these  inhibitions  have  not 
been  adequately  studied  with  respect  to  mechanism.  Some  more  important 
examples  that  could  not  be  easily  summarized  in  the  table  will  be  discussed 
briefly. 

Inhibition  of  Morphine  A/-Demethylase 

The  antagonistic  actions  of  nalorphine  (A'^-allylnormorphine)  to  the  phar- 
macological responses  to  morphine  have  been  extended  to  the  enzyme  sys- 
tem for  morphine  inactivation,  and  it  appears  that  the  configurations  of 
the  tissue  receptor  groups  and  the  enzyme  active  site  are  very  similar 
(Axelrod  and  Cochin,  1957).  Several  normorphine  analogs  were  tested  on 
the  A-demethylation  of  morphine  by  rat  liver  enzyme  (see  accompanying 
tabulation)  and  the  role  of  the  alkyl  substituent  in  the  binding  is  evident. 


MISCELLANEOUS   ANALOG   INHIBITIONS 


591 


tf 


«-^ 


0,    o  ^ 

G       rt^       c^ 


a 


•I      -S" 


o 


CO    o 

o  o 
o  o 


o  o 
V 


>. 

Cl 

■D 

<B 

+3 

> 
3 

rC 

ft 

>■. 

u 

X 

>-. 

"o    2 

ft 
5 

OJ 

-t^     -C 

o 

c3   -j; 

< 

o  Q 

_ 

>■. 

X 

«^ 

o 

-4^ 

lO    M    O    lO 


t^    IC    rt    CC    00    OO    (M 
5<l    Tt    lO    -^    ^    00    C<I 


O    O  (M    (M    lO    (M    (M 


•2  ^  ^  i  ^' 

S    P    o    p 


-C   iS   -p  -O  73    *    "5 

'JT*       cS      G       ri       C     I.  m 


£  hS  hS  ^  ^  6  f5 


o 


S 

>> 

!«1 

c« 

>> 

o 

42 

T3 

« 

O 

ert 

>> 

Tl 

<D 

© 

X 

-f< 

T3 

^«4 

o 

w 


S      O 


Q 


P^ 


o 


CO  fC 

-^  fo  fc  CO  eo 

O  I— I  CC  CO  CO 

o  d  -^  — '  ^ 


O  .S 


-5  '^ 

p     eS 
S     P 


o  W 


s 

^— ^ 

f> 

eo 

» 

CO 

o 

f-H 

2 

a> 

c 

crt 

<11 

C 

CO 

1 

C 

so 

3 

4) 

n 

(Tt 

t3 

o 

c 

X) 

>. 

>. 

j3 

^ 

0) 

S 
<1 

0) 

'P 

tS 

592 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


D5 


'^-^ 


-2 

e 


®  .2  ;^ 

s  ^  s 

O         tH       i 


s 

SI 


i 


c 


c  .S 


X    X 

o  o 


o 

o 

r-; 

r; 

« 

-c 

bi 

"a 

^ 

c 

S 

O^ 

'^. 

■>< 

<£ 

2 

2 

c 

j:3 

.g 

-tJ 

5 

CO 

4^ 

o 

3 

o 

Q 

=0 

(m' 

o 

CM 


s  :s 

cS      O 


4)    43 


M 


o  -n 


WS 
(% 


W 


OS         2 


aj   «•■ 


O 


tf 


S  ©     5 

S  3    to 

t-  fl      cS 

S  O 

^  S 
Cm 


tj) 


c  CO 

Q  - 


>» 

(a 

"o 

2 

, 

c 

fl 

o 

'x 

o 

CO 

J3 

o 

o 

tJO 

cS 

» 

« 

-»^ 

o 

(^ 

a 

1 

O 

-til 

CO 

1 

j3 

s 

P4 

P-l 

W 

o 

O 

C 

ft 

T- 

51- 

S 

'E 

(» 

bl 

© 

o 

^ 

c 

OJ 

-k^ 

u 

ft  '-; 

■ft 

r2 

i 

^- 

c  o 

<» 

© 

o 

X 

"^  2- 

o 

Is 
> 

-^ 

>> 

"o 

W 

o 

O 

ffi 


>> 
o 


MISCELLANEOUS   ANALOG    INHIBITIONS  593 


> 

c 
a 

^5 


pq  ^  O  o  Q 


o  c;  t^  t^  05  o 
<6  S  S  (6  '^  '^ 


3 

?• 

05 

J2 

-c 

c? 

s 

TS 

05 

e3 

-c 

^ 

c 

c3 

c 

ft 

c3 

53 

o 

o 

O 

b 

<D 

05 

3 

05 

>Ci 

^ 

c? 

JH 

"o 

LO 

ft 

•>* 

03 

-* 

3 

■^ 

s 

05 

t^ 

Oi 

0) 

05 

c 

^ 

03 

eg 

eS 

3 

» 

S 

fq 

"-^ 

K 

^-^ 

-5 

'    ' 

^ 

w 

00   o  oo   >o 

O    i-i    CO       O    -<    CO 
'^  XX  OOOOOO 

0>Ci0^iO>00       -^  -^  OOO       OOO^O       ^(N 


c 

(ft 

^ 

>> 

« 

OJ 

n 

-4-> 

ti 

rt 

T-l 

o 

>■. 

o 

tU 

p 

_  — ,    o 

"o    o  s      ^ 

r«     o     eS       .3 


^     3     >i 

>  6  i,  pl|  d  c5  CL, 


CD 

11 

6 

s 

4S 

IS 

i 

O 

'ft 
>> 

o 

ft 

ft 

03 

o 

2 

.a 

§ 

p 

ft 

r5 

-3 

Co 

>> 

o 
ft 

o 

T3 
>> 

^ 


<     £       -3 


^ 

X 

ft 

^.^ 

o 

(U 

*r^ 

c8 

c3 

0) 
S 

P 

ce 


tn 

O 

crt 

-fi 

T! 

o 

>> 

3 

-C 

TJ 

(1) 

0) 

Q 

S     >,  S  >>  3    -3        « 

'oj         «  s       o  I  ill 

.Stj  frtT^  r<0  Oft^ 


3     S  ^^     >. 


-3  ^   S     J  S 

>>  O    ^  3      X 


"o 

o 

>> 

3 

O 

M 

>> 

o 

3 

-►i 

rs 

« 

3 

ft 

:^ 

C 

P 

CQ 

o 

w  _  ^  -  « 

o  ::    .  ^         §  o 

a 

:s  :s^|s  o  ss9^^ 


>> 

> 

IT) 

3 

3 

T1 

■^ 

O 

-ii 

fit  Ecj  t^  ttj  t^-mg-^fiim 

-2 


O-TJ  O^  'ft— 'O  — ''^ 


594 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


tf 


^ 


c    c  ^ 

s  =«  s 


W 


o      -s 


r-H  t^ 


=Q. 


_:,    b'   i=i    a    c 
>>   S    "    S  ^ 

r-»        TO        p*-^      TO     ^ 


CS 


W2  ^§ 

S    ^     O    ^ 


S       <0 


^  aj      QJ 


a. 

o 


-fl      r— I 


O    CO 

d  d 


^  ^  ^  S 

cfl  O  cd  c^ 

CO  K^  03  TO 

O  O  <^  O 


■r  -t-  t:  ■s       >> 


X 


§  ^  §  o     o 


>H 


i2 

X 

Q 

"cS 

>. 

ci3 

S 

o 


CO      ^ 

(B  ri 


c« 

CO 

t>>  T3 

X 
O 

'x 

O 

^ 

W 

^ 

ec 

o 


'^  +^ 


3 


(Ih 


o3 

>. 

X 

o 

o 

(-1 

-5 

>> 

j3 
t3 

oi. 

MISCELLANEOUS    ANALOG   INHIBITIONS  595 


c3 
J3 


05 


^2    |2  1=    |2  ^  I  |2  I 


A  -^  A 


X 

CO 

o 

OO        OOO  0*0000        O  O  v;0?OCDCOCD        O-h 


O    O      O  ^ 

(MlOOOo*-'^  Tj<C<l  0<M 


c 

<D 

JH 

X 

a> 

2 

S 

c3 

© 

-o 

J> 

-►^ 

>. 

3 

c3 

W 

_o 

^ 

' 

ft 

2 

^ci 

^ 

Q 

O 

">. 

a 

0) 

C<j 

« 

ft 

^ 

X 

^ 

o 

c3 

O 

^ 

> 

^^ 

3 

^ 

K 

>> 

c 

O 

^ 

o 

">> 

m 

e8 

72 

^ 

O 

® 

ft 

I      O  ^4  -^   •-<       I      e     eg 


«.Sg1^  ^ft  g^ 


_S     -ti 


t<    «    a> 

'^  ^  s 

iJft3c3iJ        ^-^  C  'O'^cS^.S 


5 

>> 

o 

u 
ft 
>> 

3 

+3 
o 

X 

g 

1 

3 

s 

3~ 

X 

o 

S 

S 

O 
09 

s 

^  t^      £2      lj    —        k>    "^     !-< 

\4  S=6^i^M 


^ 

^^ 

cS 

- — - 

>, 

"-H 

>> 

® 

s 

(D 

s 

-4^ 

;h 

c 

Sh 

^O 

'S 

o 

'S 

a 


•s'a  s^  S  -S  US©  ^ 


cS 

o8 

-i-i 

55 

O       (D 

'S 

c3      CO 

(U 

IB       tS 

(h 

-o  :« 

3 

'T3     X 

^ 

G     O 

k> 

.3    ^3 

'2    S 

(0     c 

<»   S 

JH       cs 

^          Wl 

§      1 

§    § 

>>   o 

>■.        JH 

w  '^ 

W        ^ 

596 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


-2 
c 


0)    o   ^ 


-§ 

p 

>" 

lO 

4) 

Ifl 

(D 

lO 

O 

> 

O) 

-^ 

<1 

1— 1 

TS 

2 

^4-< 

c    ^^ 

" 

« 
tf 

13 

Tl* 

c3    ^ 

tH     CO 

y3 

o 

lO 

«    •* 

3 

Oi 

bC  05 

t4 

Sri    '^ 

p 

«    ^^ 

o 

cc 

pq 

M 

_,^ 

-i 

OS   »o 

(M 

O    CO 

o  00 


c<i    r-    CO 

CO     CO    <M 


o  -'  q 


O 


■rj<     lO     O     t^     CO     lO 
CO    GO    00    CO    (>i    o 


CO    iC    CO    lO    CO    »o 

o  -H  q  -H  o  ^ 
o  o  o  O  o  o 


c 

05 

>. 

0) 

^^ 

3h 

£ 

c 

"E 

^ 

S 

> 

>> 

-C 


H    o 


S  H  O 


O      43 


^        so-Ps*.    -Ts       t;       -73        ==: 


cS  l>i   GO 

O  -fi    ^^ 


(K      Cj 


^        ^       o 


P5 


Ph 


^ 


05 


eq 


MISCELLANEOUS   ANALOG    INHIBITIONS  597 


._    c 


73    ^         73    —I 


|-^  <1  H  O  O  > 


OiO-Hl        I05C0         OOOOiO 


CO    O  CD    CD         O    O 

O    O      I  O    -H         ^    '7' 

(MOO  O    O         O    -^ 


ft 


=s  -S  j2  -2    -*^ 

i  o    2    §    c     y 


ci  ^ 


—2  T3      °  "S  Ci      S      «« 


ce 

c 

3 

0) 

CO 

O 

Is 

^ ^ 

'2 

to 

c8 

t^ 

< 

O 

^ 

9  ^  ^ 

-*-'>;  *^  "^  ^  -t^  'X  a  c8 

0§^  S  ^t^Og  SO  O 


_  _  S        GO  o      — ^      "" 

O  O  ,,    TJ  O     4>       S 


^  -2  b    -s 


3 

00 

_o 

o 

s 

o 

'S 

o 

'zi 

o 

M 

C 

^-^ 

<» 

'-3 

s 

">. 

X 

c 

o 

« 

;-i 

o 

Is 

T3 

S 

•I   S  8  - 


m 

«> 

_cS 

Ph 

C 

'>^, 

OJ 

«.  ^ 

o 

X 

-S 

1   ^ 

o 

O 

cS 

3 

^ 

O 

S       (u 

03 

;h 

•*i 

Is 

eS 

o 

-2     ^ 

« 

Ph 

■       tiC 

X 

T3 

QQ. :- 

O 

ij 

598 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


tf 


'^   S 


c 


ai    o  ^ 

C         Hj         ? 


a 

c 


;s  -s 


S  ■* 


i5  -JS 


oooooo    oooooo 


OOOO'-^CO   OO^Tt^COOO 

oooooo  oooooo 
<6  <6  S  (6  <6  '^    dododd 


.i 

.^ 

o 

^ 

c 

c 

o 

ft 

o 

6 

Q 

2 

<D 

2 

C 

o 

& 
o 

ft 
o 

0) 

s 

ft 

2 
o 

43 

a 
o 

o 

Q 
a 

a' 

c 
_o 

2 

p 

H 

3 
2 

o 

'ft 

2 

ft 

a 

'o 

X 
O 

O 

o 

'o 

'ft 

2 

'S 

c 

Si 

'o 

"S 
c 
Ph 

'o 

X 
O 

O 

'S 

X 

O 

'S 

.S   ."  .S  '^  ^ 

^  :j3  S  ^  :S  ^ 


43  £; 


3 
T.  — 


m 


ft 

>. 

^^ 

N 

r^ 

C 

P3 

Ji 

1^ 


p-l 


MISCELLANEOUS   ANALOG   INHIBITIONS  599 


5  ^  CO  £ 

.ti  O  'I^    -S  C  S  *  o 

O  «<  OJ  O  O 


•M    00    ■>*    .— ( 

■*     fC      L'T     fC 


O    rt    rM    O    O    O 
O    O    O    O    O    O 


Q    s 


1:0       N 

fl 

(S 

'0 

ft 

O    O    O    O    O    O    o 


o 


>    .    .        .    ^    .    . 

_  j^     O"    q"    J     J     ^     Q     Q     J     J 


^  ^  3  |te 


-2 

,4 

ft 

eS 

>> 

C 

X 

m 

ft 
5 

2 

0 

CO 

0 

0 

?i, 

&>  '3  ■«  s  Q 


>. 

X 

r^ 

p 

r- 

-5 

>-, 

>i 

<o 

H 

ffi 

ft 

i. 

a 

*5  ft 


W  P4  P4  i.                                          Q 

"*  •*  » 

s  s  §  _ 

53  «  5=  "-g 

S  §  §  »                                        g 

-§  -2  g"  S                                                                     S^                                                            e3 

a  «i  «  a                                                       fc!                                                « 

cq  Q^  CQ  «                                               M                                         m 


s     o 


600 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


c 

c 

^^ 

8 

.2 

^ 

o 

^ 

a 

4^ 

s 

Pi 


.'^ 


<N    (M    (M    <M 


0)      Ph 

c    o 


ri  -^  i  -^    2 


_fl 

IB 

c 

'a 

& 

.2 

o 

>•; 

T3 

^ 

'?. 

■^ 

■>> 

<» 

c 

^ 

0) 

B 

CM 

fi  .s 


P,    Oi   -2     q 


eS 

03 

1-1 

§ 

B    ^ 

Ti 

G      -H 

c 

a 

o    a 


O 

o 

-c 

o 


o 

M 
ft 

O 


fi 

4) 

^ 

73 
(Si 

-a 

c: 

n 

_C 

H 

ft 

n1 

i1> 

PLi 

T3 

■^   •*   o   «c 

— I  00  lo  ^ 


O    CO    M    05    CO    CO    ffC 


15      ft 


«  is        cs   :» 


cS 


o3 


2  -is  ft 

S  O  O      g  ^ 

Q,  -1^  -k^        M  CS 

--  "  »  ,«^  -Ji  ft 

o  <t:  w  w  ^  ^5 

j    hj  a  VX.O  -< 


'—  CO 
tJH    CO 

O    CO 


I 
13 

03 


•§-2   2 


eS 


a  .a 


a;     S 


X 

O  cS  K 

r"  ™  S 

f:-  X  g 


P^ 


P 

ft  ^ 

r'.  OS 

O  >« 


iJ 

TJ 

V 

>> 

>> 

^ 

X 

v 

-2 

2 

x) 

'eg 

3 

s 

<o 

Ih 

o 

o 

''I 

p^ 

I   - 


MISCELLANEOUS    ANALOG    INHIBITIONS  601 


•r   -<  o 


s 


V 

cS 

Id 

02 

!> 
O 

T3 

T3 

C 

cS 

c3 

M 

02 

M 

iC 

b 

O 

0) 

1^ 

h-) 

^ 

A         tf  P5 


c 

4) 

G 

2 

>> 

o 

"S 

(^ 

'x 

o 

(U 

c6 

3D 

o 

3 

-0 

00    O    O    O    <M 

Tt<    to    (M    O    O 

M    lO 

■*    ■'l^    (M 

ic  ic  lO  '-f  c^ 

lo  o  c- 

CI    fO 

t-      I-      Tj< 

O  rt^^^cq         ^^^^^ 


»0    .-    T3     (B 
>1    S    •?    t3 


t>v 


^ 

-2 

Id 

5 

S 

5 

C 

o 

03 

o3 

H 

H 
o 

2 
(-1 

H 

a 

til 

H 

-2 

o3 

C 

2 

u 

H 

p 

ij 

hj 

s 

Q 

0 

Q 

H 

I— I  fa,  <N  oCOCD55,?i, 


<D      fa  ^  «      ^ 


o  ^  O   ,?3  00 


c8 


«>  _2   02  45   03     S   03 

02  "ti  oS  -t^  c5    t-  =« 


0) 

02 

T3 

2 

>> 

1 

"eS 

'^ 

>. 

s 

<U 

"« 

602 


2.  ANALOGS  OF  ENZYME  KEACTION  COMPONENTS 


-Q 

m 

CI 

o 

t4 

c 

o 

£ 

w 

'^ 

t3 

OJ 

c 

oi 

cS 

O 

+J 

o 

O 

.-^ 

^  s 

s  s 


s 


O   O   O   o   o   o   o 


■2  ^   * 


»>1 


^'  3    X 


O  O  H  O  Ph    j  o 


t3 

J3 


T) 

>. 

^ 

4) 

o 

« 

to 

2 

eS 

c 

+s 

o 

'rt 

o 

u 

•^ 

3 

(-1 
as 

C 

73 

aj 

(D 

e  2 


f<i    ^  lO 


fc 

n 

43 

s 

^. 

O 

a 

0) 

t3 
o 

a 
n 

H 

S 

H  o 


b 


3  S 


■* 

CO 

00 

lO 

(M 

(M 

O 

o 

<^ 

O 

o 

o 

o 

.5      X 


.s  -^  -s 


>> 

-C 


O      >>  T3     (U 


.2  H 


a3   .ii    Is   .ii    ti;  . 

<5      '        "      ■        — 


O  ■<  H 


.<U     eS    ,>»     ft 


•a 

O 
43 


tf 


V 


« 


H    4J 


MISCELLANEOUS   ANALOG   INHIBITIONS 


603 


rt  ^  o 


.S    o 


<D 


P  Q 


M     ;>. 


o 
o  .2 
"-  9 


CO  H 


tj   c 


-t!         o 


01 

'53 

c 

»o 

§■1 


"ft    § 
ft  TS 


c 

ft 
o 

ft 


!0 

05 

-0 

c 

^ — 

c« 

"S 

03 

CO 

> 
ci3 

OS 

tf 

T) 

o 

^^ 

c 

.5 

^ 

eS 

o3 

^ 

IB 

J= 

o3 

W 

«} 

GO 

IC    O     "*    >c 


fO    w 

o 

o 

o 

o 

00  00 

o 

o 

o 

o 

IC 

>o 

C<l 

Tj< 

eS     •- 


a;   .i;    -tJ    .i; 
_     .,     o    J2     0)    ^ 

g  o  <:  H  S  H 


P 


t3    P 


»-i  o 


;ir  ;=r     ph 


<r> 

CO 

T3 

05 

05 

C 

■ — ' 

^-' 

cS 

^^ 

C 

^ 

_> 

_> 

Oi 

'3 

O 

O 

" 

(M 

T(< 

■* 

s^ 

Tt< 

O 

o 

O 

o 

o 

C5 

lo 

o 

o 

o 

o  o 


O 

CO 

o 

o 

o 

3 

_2 

M 

"m 

CM 

Pm 

Q 

P 

ID 

t) 

P^ 


s 

J5 

ft 

CO 

O 

"m 

ft 

t3 

o 
ft 

o  o  o 


a>  i  .S   -S   .2 

(B  OJ  ^     ^     ,£5 

-S  2  ^  ^  ^ 

_2  cs  000 


>. 
S  S 


^   ^        c 


>^ 


.a  -c    c 


-i 

^ 

773 

0) 

ce 

> 

^ 

0 

>i    "1^ 


c 

^ 

<o 

2 

_c 

_o 

0 

■-3 

00 

0) 

CO 
<» 

-0 

e 

c3 

■*^ 

c 

01 

fi 

_o 

t4 

'-3 

•^ 

03 

^ 

'pi 

ft 

^ 

0) 

ft 

■— ' 

s 

^ 

0 

0 
0 

B 

t< 

SB 

c 

2 

c 

^.^ 

3 

? 

•^ 

"5 

3_ 

CO 

bh 

t3 

c« 

c 

-P 

0 

fl 

'.+:> 

73 

0 
0 

5 

^ 

cS 

« 

_M 

"3 

1 

CO 

0 

T3 

"S 

M 

0 

^ 

« 

0 

-C 

■w 

CO 

H 

!^ 

604  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


Analog  (0.2  mM)" 

%  Inhibition 

Relative  —AF  of  binding 
(kcal/mole) 

iV^-Methallylnormorphine 

74 

4.90 

iV-Isobutylnormorphine 

69 

4.75 

iV^-AUylnormorphine 

64 

4.61 

i\7^-Hexylnormorphine 

44 

4.13 

iV-Butylnormorphine 

35 

3.88 

2\r-Propylnormorphine 

20 

3.40 

iV-Ethylnormorphine 

13 

3.09 

iV-Isopropylnormorphine 

11 

2.97 

Normorphine 

0 

<2.45 

"  Morphine  =  1  mM. 

Some  generalizations  may  be  made:  (1)  increase  in  chain  length  increases 
the  binding  energy,  (2)  unsaturation  increases  the  binding  by  1.0-1.2  kcal/ 
mole,  (3)  each  additional  methylene  group  augments  binding  by  approxi- 
mately 0.3  kcal/mole,  and  (4)  the  inhibition  by  nalorphine  is  noncompetitive 
(it  was  stated  that  this  may  be  a  slow  pseudoirreversible  inhibition  but  in- 
cubation with  morphine  and  inhibitors  was  for  2  hr. 

Dehydroshikimate  Reductase 

The  inhibition  of  this  enzyme  by  various  phenolic  compounds  points  to 
the  manner  in  which  the  substrate  is  bound  and  the  configuration  of  the 
active  site.  Relative  binding  energies  are  given  in  Table  2-39.  These  inhibi- 
tions are  all  strictly  competitive.  It  is  seen  that  all  effective  inhibitors  have 
a  p-OH  group,  and  Balinsky  and  Davies  (1961  b)  postulated  from  the  pos- 
sible ring  configurations  of  shikimate  that  this  group  must  lie  approximately 
in  the  equatorial  plane.  Additional  OH  groups  increase  the  binding  slightly 
or  not  at  all,  so  that  m-OH  groups  seem  to  participate  little  in  the  binding. 
One  might  expect  the  carboxylate  group  to  be  bound  to  an  enzyme  cationic 
group,  but  this  does  not  appear  likely;  e.g.,  the  addition  of  a  C00~  group 
to  catechol  increases  the  binding  very  little,  and  the  benzoates  without  a 
p-OH  are  bound  very  poorly.  The  stronger  reaction  of  the  aldehyde  group 
in  vanillin  also  indicates  that  the  forces  here  are  not  merely  electrostatic. 
Substitution  of  benzoate  in  the  o-position  is  detrimental  to  the  binding  and 
this  may  be  due  to  steric  hindrance,  as  shown  in  the  diagram  of  the  active 
site  presented  by  Balinsky  and  Davies  (Fig.  2-21).  The  energy  of  binding 
of  the  p-OH  group  is  greater  than  2.3  kcal/mole  and  thus  hydrogen  bond- 
ing may  be  involved.  It  is  worth  noting  that  the  experiments  were  run  at 


MISCELLANEOUS   ANALOG   INHIBITIONS 


605 


Table  2-39 
Analog  Inhibition  of  Dehydroshikimate  Reductase" 


Analog 


Structure 


Relative  -  AF 

of  binding 

(kcal/mole) 


Vanillin 


CH3O 

HO— (\       V-CHO 


5.72 


(Shikimate) 


Gallate 


HO 
HO— (        />— COO' 
HO 
HO 


HO 


COO 


HO 


(5.29)' 


4.85 


/»-Hydroxybenzoate 


^0—{\        ,)—  COO" 


4.13 


Protocatechuate 


HO 


HO— <x        /)— COO 


4.12 


Catechol 


HO 


HO 


3.86 


Guaiacol 


CH3O 


2.78 


606  2,  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Table  2-39  (continued) 


Analog 


Structure 


Relative  -  AF 

of  binding 

(kcal/mole) 


Phenol 


HO 


1.84' 


2, 4-Dihydroxybenzoate 


HO 


COO 


OH 


1.84' 


Benzoate 


COO 


<  1.84' 


w -Hydroxybenzoate 


HO 

<Q>-coo- 


<  1.84^^ 


HO 


3,  5-Dihydroxybenzoate 


COO' 


<  1.84' 


Salicylate 


<Q^coo- 

OH 


<  1.84' 


«From  Balinsky  and  Davies  (1961b). 

f>  The  value  for  shikimate  was  obtained  from  K,,,  assuming  that  this  is  a  true  substrate 

constant. 
'"Values  calculated  on  the  basis  of  inhibitions  reported  and  are  very  approximate. 

pH  9  where  the  phenolic  groups  are  partially  ionized,  and  this  may  account 
for  some  of  the  differences  in  inhibitory  activity  between  the  compounds. 
Variation  of  the  inhibition  with  pH  might  provide  some  interesting  infor- 
mation. 


MISCELLANEOUS   ANALOG   INHIBITIONS  607 

Kynurenine   Transaminase 

The  transamination  between  kynurenine  and  a-ketoglutarate  catalyzed 
by  an  enzyme  from  rat  kidney  is  inhibited  by  a  variety  of  mono-  and  di- 
carboxylates  (Mason,  1959).  The  inhibitions  are  competitive  with  respect 
to  kynurenine  and  are  reversible.  The  results  are  interpreted  in  terms  of 


NAOP 

Fig.  2-21.   Topographical    scheme   of 
the  active  site  of  dehydroshikimate 
reductase.   (From  BaHnsky  and   Da- 
vies,   1961  b.) 


two  types  of  interaction:  (1)  electrostatic  binding  of  carboxylate  groups  to 
cationic  groups  on  the  enzyme,  and  (2)  van  der  Waals'  forces  between  the 
hydrocarbon  portions  of  the  inhibitors  and  the  enzyme  surface.  The  varia- 
tion of  inhibition  in  the  dicarboxylate  series  (Table  2-40)  indicates  that  two 
cationic  groups  interact  maximally  with  adipate.  The  distance  between 
these  groups  was  given  by  Mason  as  11  A  on  the  basis  of  an  extended  adipate 
molecule;  the  intercarboxylate  distance  for  adipate  is  given  as  6.87  A  in 
Table  1-1.  However,  the  cationic  groups  need  not  be  the  same  distance 
apart  as  the  carboxylate  groups,  and  not  only  is  the  distance  important 
but  the  allowed  configuration  of  the  methylene  chain  to  interact  maximally 
with  the  enzyme.  As  Mason  points  out,  the  data  from  the  phthalates  do 
not  support  this  distance  entirely,  since  terephthalate  inhibits  least  of  the 
three  isomers  and  its  intercarboxylate  distance  is  the  closest  to  that  of 
adipate.  The  intercarboxylate  distance  in  isophthalate  is  5.84  A  (Table  1-1) 
and  this  might  indicate  that  the  cationic  groups  are  closer  than  might  be 
expected  from  the  data  on  the  flexible  dicarboxylates,  but  again  there  is 
the  problem  of  the  orientation  of  the  benzene  ring.  The  importance  of  van 
der  Waals'  interactions  is  shown  by  the  increasing  inhibition  given  by  the 
higher  fatty  acids,  and  the  greater  inhibition  by  the  alkyl-substituted  glu- 


608 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


Table  2-40 
Inhibition  of  Kynurenine  Transaminase  by  Mono-  and  Dicarboxylates  " 


Relative  —AF 

Inhibitor 

Concentration 
(mM) 

%  Inhibition 

of  binding 

(kcal/mole) 

Alkyl  monocarboxylates 

Formate 

6 

0 

<1.43 

Acetate 

6 

0 

<1.43 

Propionate 

6 

0 

<1.43 

Butyrate 

6 

0 

<1.43 

Valerate 

6 

0 

<1.43 

Caproate 

6 

1 

1.43 

Heptanoate 

6 

8 

2.75 

Caprylate 

6 

15 

3.19 

Nonanoate 

6 

24 

3.55 

Caprate 

6 

41 

4.04 

Straight-chain  dicarboxylates 

Oxalate 

6 

0 

<1.43 

Malonate 

6 

0 

<1.43 

Succinate 

6 

0 

<1.43 

Glutarate 

6 

8 

2.75 

Adipate 

6 

65 

4.64 

Pimelate 

6 

30 

3.74 

Suberate 

6 

2 

1.86 

Azelaeate 

6 

35 

3.88 

Sebacate 

6 

65 

4.64 

1,10-Decanedicarboxylate 

3 

74 

5.33 

1,11  -Undecanedicarboxylate 

3 

78 

5.47 

1 ,  14-Tetradecanedicarboxylate 

3 

87 

5.86 

Cyclic  monocarboxylates 

Benzoate 

6 

0 

<1.43 

y-Phenylbutyrate 

6 

9 

2.91 

Cyclohexanecarboxylate 

6 

0 

<1.43 

y-Cyclohexanebutyrate 

6 

64 

4.62 

Cyclic  dicarboxylates 

o-Phthalate 

12 

14 

2.72 

Isophthalate 

12 

35 

3.45 

Terephthalate 

12 

9 

2.48 

Cyclohexane- 1 ,2-dicarboxyiate 

6 

9 

2.91 

Glutarate  derivatives 

2-Methylglutarate 

6 

15 

3.19 

2,2-Diniethylglutarate 

6 

35 

3.88 

2,4-Dimethylglutarate 

6 

33 

3.82 

3-Methylglutarate 

6 

47 

4.18 

3,3-Dimethylglutarate 

6 

56 

4.41 

3-Methyl-3-ethylglutarate 

6 

39 

3.98 

3,3-Diethylglutarate 

6 

20 

3.40 

/J-Ketoglutarate 

6 

0 

<1.43 

2,2-Dimethylsuccinate 

fi 

2 

1.86 

"  Kynurenine   was   3.7    mM   and   a-ketoglutarate   was   6   mM;   pH 
Mason,  1959.) 


6.3.   (From 


MISCELLANEOUS   ANALOG   INHIBITIONS 


609 


tarates  compared  to  glutarate.  One  of  the  cationic  groups  on  the  enzyme 
seems  to  have  a  pK,,  around  6.7,  since  the  mhibition  by  the  dicarboxylates 
decreases  from  pH  5.5  to  8.5  and  approaches  that  of  the  monocarboxylates 
(Fig.  2-22).  The  increase  in  inhibition  of  the  dicarboxylates  with  longer 
chain  lengths  than  suberate  is  explained  by  the  ability  of  the  flexible  hy- 
drocarbon portions  to  orient  for  effective  interaction  with  the  enzyme  sur- 
face between  or  around  the  cationic  groups.  The  contribution  of  a  methyl- 
ene group  to  the  binding  is  around  0.2-0.4  kcal/mole.  It  is  likely  that  a 


100 


VALERATE 


5.5 


6.0 


6.5 


7.0 


7.5 


8.0 


8.5 


pH 


Fig.  2-22.  Effects  of  pH  on  the  inhibitions  of  kynurenine  transaminase  by 
various  fatty  acid  anions  at  6  mM.  (From  Mason,   1959.) 


hydrophobic  region  of  the  enzyme  lies  at  some  distance  from  the  cationic 
groups  since  the  fatty  acids  do  not  begin  to  inhibit  until  the  chain  length 
reaches  5  or  6  carbon  atoms,  and  the  marked  differences  between  the  bind- 
ing energies  of  benzoate  and  y-phenylbutyrate,  and  cyclohexanecarboxylate 
and  y-cyclohexanebutyrate,  indicate  that  the  ring  interaction  is  effective 
when  the  ring  is  separated  from  the  carboxylate  group  by  several  angstroms. 
However,  these  differences  may  be  due  more  to  steric  factors,  a  ring  close 
to  the  carboxylate  group  interfering  with  its  electrostatic  interaction.  In- 
deed, there  seems  to  be  a  region  for  van  der  Waals'  interactions  between  the 


610  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

two  cationic  groups  on  the  enzyme,  since  the  3-substituted  gkitarates  are 
bound  more  tightly  than  the  2-substituted  isomers. 

The  administration  of  nicotinylalanine  to  rats  leads  to  a  4-fold  increase 
in  the  urinary  level  of  A"-methylnicotinamide  (Decker  et  al.,  1963).  Nicotin- 
ylalanine is  possibly  formed  from  tryptophan  through  3-hydroxykynure- 
nine  but  studies  with  tryptophan-C^^  indicate  it  not  to  be  a  metabolite  but 
actually  a  strong  inhibitor  of  kynureninase  and  kynurenine  hydroxylase. 
This  inhibition  presumably  occurs  in  vivo,  resulting  in  a  sequential  block 
in  the  major  route  of  kynurenine  degradation.  The  effect  of  this  analog  on 
the  transaminase  is  not  known. 

Inhibition   of  Urease   by   Methylurea  and   Thiourea 

There  is  a  good  deal  of  disagreement  on  the  analog  inhibitions  of  urease 
and  the  kinetics  are  certainly  not  simple.  Takeuchi  (1933)  originally  found 
very  little  inhibition  (perhaps  5%)  by  83  mM  methylurea  (although  mark- 
ed inhibition  by  oxyurea  —  which  may  be  HgN — CO — NHgOH  —  occurs, 
this  is  probably  not  competitive).  Sophianopoulos  and  Corley  (1959)  ob- 
tained competitive  inhibition  by  methylurea  at  lower  concentrations  (un- 
specified), but  higher  concentrations  (300-1000  mM)  increase  the  substrate 
inhibition  produced  by  urea;  these  latter  effects  may  be  related  to  enzyme 
denaturation.  The  inhibition  was  found  to  depend  on  the  pH  by  Shaw  and 
Raval  (1961),  it  being  noncompetitive  between  pH  7  and  8.9,  and  competi- 
tive below  pH  7.  Furthermore,  the  kinetics  correspond  to  the  reaction  of  2 
molecules  of  methylurea  with  the  active  center,  which  may  relate  this  type 
of  inhibition  to  substrate  inhibition. 

Thiourea  was  claimed  to  stimulate  urease  at  5  mM  (Sizer  and  Tytell, 
1941),  to  have  no  effect  below  50  mM,  and  to  inhibit  35%  at  500  mM 
(Kistiakowsky  and  Shaw,  1953).  This  inhibition  is  completely  reversible 
and  occurs  rapidly.  At  pH  6  the  inhibition  is  competitive  but  as  the  pH  is 
raised,  deviation  occurs.  The  kinetics  again  point  to  2  molecules  of  thiourea 
reacting  with  the  enzyme.  The  reaction 

E  -f  2  I  ->  EI2 

is  not  affected  by  change  of  pH,  whereas  the  reaction 

ES  +  2  I  ->  ESl2 

is  sensitive  to  pH,  which  serves  to  explain  the  change  in  inhibition  type 
with  the  pH.  Lister  (1956)  reported  that  of  the  17  urea  analogs  tested, 
only  thiourea  is  inhibitory  —  35%  at  200  mM,  70%  at  400  mM,  and  85% 
at  1000  mM,  when  urea  is  500  mM.  The  inhibition  is  prevented  by  cysteine, 
which  brings  up  the  possibility  of  disulfide  bond  formation  by  thiourea  at 
high  concentrations. 


MISCELLANEOUS   ANALOG   INHIBITIONS  611 

Inhibition    of  Catechol-0-methyitransferase    by    Pyrogallol 

This  inhibition  is  of  interest  because  of  the  bearing  it  has  on  the  metab- 
olism of  epinephrine  and  norepinephrine.  Bacq  (1936)  observed  that  pyro- 
gallol increases  the  responses  of  tissues  to  sympathetic  nerve  stimulation  and 
to  epinephrine.  However,  he  then  attributed  this  action  to  the  antioxidant 
properties  of  pyrogallol.  Lembeck  and  Resch  (1960)  and  Vanov  (1962)  have 
recently  confirmed  this  by  showing  that  the  pressor  response  to  epineph- 
rine is  prolonged  by  pyrogallol.  The  inhibition  of  the  catechol-0-methyl- 
transferase  was  reported  by  Bacq  et  al.  (1959),  who  believed  that  this  could 
explain  the  sensitization  of  smooth  muscles  to  the  catecholamines  by  pyro- 
gallol and  other  phenolic  compounds.  Axelrod  and  Laroche  (1959)  also 
found  a  potent  inhibition  of  this  enzyme  (50%  when  pyrogallol  =  epineph- 
rine =  0.01  niM),  which  decreases  with  increasing  substrate  concentration, 
indicating  a  competitive  action.  Furthermore,  about  70%  of  intravenously 
injected  epinephrine-H^  is  metabolized  in  10  min  in  mice,  but  pretreatment 
with  100  mg/kg  pyrogallol  reduces  the  amount  metabolized  to  22%.  The 
half-life  of  norepinephrine  in  mice  is  increased  from  22  to  42  min  by  10  mg 
pyrogallol,  while  at  the  same  time  0-methylation  is  inhibited  99%,  indi- 
cating other  pathways  for  norepinephrine  metabolism  (Udenfriend  et  al., 
1959).  Probably  the  monoamine  oxidase  pathway  is  also  important.  The 
administration  of  pyrogallol  to  rats  does  not  by  itself  increase  brain  nor- 
epinephrine levels,  but  in  conjunction  with  iproniazid  (which  inhibits  mo- 
noamine oxidase)  it  does,  in  this  case  the  two  major  degradative  pathways 
being  blocked  (Jaattela  and  Paasonen,  1961).  This  is  a  good  example  of 
the  action  of  two  inhibitors  on  a  divergent  multienzyme  system  and,  in 
addition,  has  interesting  possibilities  for  clinical  application. 

Repeated  administration  of  pyrogallol  causes  a  rise  in  the  blood  pressure 
but  this  is  soon  followed  by  a  loss  of  response  or  tachyphylaxis  (Wylie  et 
al.,  1960).  The  rate  of  urinary  excretion  of  0-methylated  derivatives  of  the 
catecholamines  is  briefly  decreased  by  pyrogallol,  but  if  the  administration 
is  continued  the  rate  returns  to  normal  (Nukada  et  al.,  1962).  Long-term 
treatment  with  pyrogallol  leads  to  an  increase  in  0-methyltransferase  and 
monoamine  oxidase  in  the  liver  of  rats,  so  it  may  well  be  that  these  enzymes 
are  adaptively  altered.  The  urinary  excretion  changes  of  the  catecholamines 
and  their  0-methylated  products  are  shown  in  Fig.  2-23. 

The  kinetics  of  the  in  vivo  inhibition  have  been  studied  by  Crout  (1961). 
Inhibition  of  0-methyltransferase  occurs  very  rapidly  in  liver,  heart,  and 
brain  even  when  the  pyrogallol  is  injected  intraperitoneally,  and  by  30  min 
has  developed  appreciably.  The  inhibition  of  the  enzyme  obtained  from  rat 
tissues,  however,  is  only  partly  competitive  (actually  the  curves  appear  to 
indicate  pure  noncompetitive  inhibition)  despite  the  fact  that  pyrogallol 
is  a  substrate  for  the  enzyme.  The  K,  of  0.008  vs\M  for  pyrogallol  indicates 
the  high  potency  of  the  inhibition  (^,„  for  norepinephrine  is  0.3  vaM).  Fur- 


612 


2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 


ther  work  on  the  exact^mechanism  of  this  inhibition  might  provide  interest- 
ing information.  Wylie  et  al.  (1960)  found  pyrogallol  to  be  the  most  potent 
inhibitor  of  a  series  of  polyphenols  and  epinephrine  analogs,  inhibiting  50% 
at  0.03  mM  when  epinephrine  is  0.3  mM.  However,  gallate,  adrenalone,  and 
arterenone  are  almost  as  active.  A  new  inhibitor  of  0-methyltransferase  was 
studied  briefly  by  D'lorio  and  Mavrides  (19^3).  This  is  3.5-diiodo-4-hy- 
droxybenzoate  and  it  inhibits  the  rat  liver  enzyme  competitively  with 
Ki  =  0.013  mM. 

Inhibition  of  the  Oxidation  of  Aromatic  Compounds  by  Bacteria 

The  metabolism  of  o-nitro-  and  p-nitrobenzoate  by  Nocardia  is  quite 
strongly  inhibited  by  m-nitrobenzoate,  o-nitrophenol,  and  p-nitrophenol 
(Cain,  1958).  The  competitive  nature  of  this  interference  was  demonstrated 
by  reciprocal  plotting.  The  enzymes  involved  here  are  not  well  characterized 
and,  indeed,  the  action  could  be  on  a  transport  mechanism  at  the  mem- 


+  100 


+  50 


% 

CHANGE 


-50 


-100 


METANEPHRINE 

AND 

NORMETANEPHRINE 


30 

DAYS 


Fig.  2-23.  Effects  of  pyrogallol  injected  subcutaneously 

at  a  dose  of  20  mg/kg  on  the  urinary  excretion  of  free 

and     methylated     catecholamines    by    rabbits.     (From 

Nukada   et  al,    1962.) 


MISCELLANEOUS   ANALOG   INHIBITIONS  613 

brane.  Durham  and  Hubbard  (1959,  1960)  favor  competition  for  a  transport 
system  in  the  inhibition  by  p- aminosalicylate  of  the  oxidative  assimilation 
of  p-aminobenzoate  in  Flavobacterium.  In  the  presence  of  ^-aminosalicylate 
there  is  more  p-aminobenzoate  remaining  in  the  medium  and  almost  com- 
plete inhibition  of  uptake  is  seen  at  a  p-AS/p-AB  ratio  of  10.  p-Aminoben- 
zoate  may  be  not  only  a  necessary  metabolite  for  folate  synthesis,  but  also 
the  principal  source  of  energy  for  growth.  It  is  very  difficult  in  such  cases 
to  determine  whether  the  inhibition  is  on  a  surface  transport  or  an  intra- 
cellular enzyme  until  the  enzymes  responsible  for  the  metabolism  have  been 
isolated  and  examined. 

Inhibition  of  Acetate  and  Fat  Metabolism  by  Propionate 

The  original  work  in  this  field  was  done  by  Jowett  and  Quastel  (1935  a,b). 
The  transformation  of  butyrate  into  acetoacetate  in  guinea  pig  liver  slices 
is  inhibited  strongly  by  benzoate,  /9-phenylpropionate,  and  cinnamate.  Pro- 
pionate also  inhibits  but  more  weakly  (59%  at  10  mM).  Much  later  Felts 
et  al.  (1956)  reported  that  4  mM  propionate  almost  completely  depresses 
the  incorporation  of  acetate- 1-C^*  into  fatty  acids  in  rat  liver  slices.  The 
formation  of  C^Og  is  also  suppressed.  Propionate  is  known  to  be  inhibitory 
to  the  growth  of  many  bacteria  and  fungi,  so  the  question  of  the  mechanism 
of  its  action  is  of  some  importance.  It  has  often  been  attributed  to  a  com- 
bination with  and  depletion  of  coenzyme  A.  This  inhibition  has  been  studied 
most  thoroughly  by  Pennington  (1956,  1957),  who  found  that  the  reaction 
acetate- 1-C^*  -^  C^^Og  in  rat  liver  can  be  inhibited  readily  and  almost  com- 
pletely, while  simultaneously  the  total  amount  of  acetate  disappearing  is 
reduced.  This  also  occurs  in  kidney,  heart,  and  diaphragm,  but  to  a  lesser 
extent.  Even  concentrations  as  low  as  0.5  mM  are  40%  inhibitory  in  the 
liver.  It  was  felt  that  propionate  blocks  both  the  uptake  of  acetate  and  the 
formation  of  acetyl-CoA.  The  oxidation  of  pyruvate  is  inhibited  much  less 
and  that  of  butyrate  not  at  all.  However,  most  of  the  action  must  be  on 
the  intracellular  metabolism  inasmuch  as  marked  inhibition  is  seen  in  liver 
homogenates  (Pennington  and  Appleton,  1958).  Addition  of  coenzyme  A 
in  the  presence  of  propionate  increases  the  amount  of  COg  formed  from 
acetate  slightly  but  does  not  reverse  the  inhibition,  indicating  that  a  simple 
depletion  of  coenzyme  A  is  not  the  mechanism.  It  was  postulated  that 
perhaps  propionate  inhibits  after  being  metabolically  altered,  possibly  to 
propionyl-CoA,  or  directly  inhibits  acetyl-CoA  synthetase.  This  is  an  in- 
teresting and  metabolically  important  inhibition  so  that  one  looks  forward 
to  studies  on  the  enzymes  involved  in  acetate  metabolism. 

A  few  instances  of  the  inhibition  of  acyl-CoA  metabolism  by  analogs  have 
been  reported.  Tetrolyl-CoA  and  propiolyl-CoA,  the  acetylenic  analogs  of 
butyryl-CoA  and  propionyl-CoA,  respectively,  are  potent  noncompetitive 
inhibitors  of  fatty  acid  synthesis  in  brain  and  liver  extracts  (Brady,  1963; 


614  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

Robinson  et  al.,  1963).  The  mechanism  of  inhibition  here,  however,  appears 
to  be  by  reaction  with  enzyme  SH  groups.  The  coenzyme  A  is  necessary 
since  free  tetrolate  does  not  inhibit,  and  is  split  off  during  the  reaction. 
Palmityl-CoA  noncompetitively  inhibits  the  condensing  enzyme  with  respect 
to  acetyl-CoA,  and  perhaps  competitively  with  respect  to  oxalacetate  (Wiel- 
and  and  Weiss,  1963).  The  inhibition  develops  slowly  and  the  authors  sug- 
gest configurational  changes  in  the  enzyme,  although  there  is  no  direct 
evidence  for  this.  Such  an  inhibition  might  be  regulatory  with  regard  to 
the  operation  of  the  cycle  and  the  formation  of  acetoacetate  in  the  liver  by 
controlling  the  rate  of  oxidation  of  acetyl-CoA  through  the  cycle.  Higher 
acyl-CoA's  inhibit  rat  liver  acetyl-CoA  carboxylase  very  strongly,  oleyl- 
CoA  and  stearyl-CoA  being  competitive  with  K/s  of  0.0013  milf  and  0.00071 
mM,  respectively,  this  possibly  playing  an  important  role  in  the  homeo- 
static  control  of  fatty  acid  synthesis  (Bortz  and  Lynen,  1963). 

Inhibition  of  cholesterol  biosynthesis  will  be  discussed  in  a  subsequent 
chapter,  but  it  is  worthwhile  mentioning  at  this  point  that  a-phenyl-n-bu- 
tyrate  not  only  lowers  serum  cholesterol  but  also  inhibits  the  incorporation 
of  acetate  into  fatty  acids,  whereas  the  oxidation  of  acetate  is  only  weakly 
inhibited  (Steinberg  and  Fredrickson,  1955).  The  evidence  points  to  an  ac- 
tion early  in  acetate  metabolism,  possibly  the  acetylation  of  coenzyme  A 
or  transacetylations  from  acetyl-CoA. 

Antagonism  between  Tungstate  and  Molybdate 

Molybdate  is  a  necessary  cof actor  in  the  growth  of  many  organisms  and 
has  been  found  to  participate  in  certain  enzyme  reactions,  such  as  those 
catalyzed  by  xanthine  oxidase  and  nitrate  reductase.  If  chicks  are  fed  on 
a  low-Mo  diet  containing  4.5-9.4%  mg  sodium  tungstate,  the  growth  rates 
are  depressed  and  signs  of  molybdenum  deficiency  appear  (Higgins  et  al., 
1956  a).  The  levels  of  molybdenum  in  the  tissues  fall  to  less  than  10%  of 
the  normal  and  xanthine  oxidase  activity  is  severely  depressed,  leading  to 
an  alteration  in  the  excretory  pattern  of  purines.  These  changes  are  re- 
versed by  adding  2-6  mg%  sodium  molybdate  to  the  diet.  Similar  falls  in 
xanthine  oxidase  were  observed  in  rats. 

Aspergillus  niger  requires  molybdate  especially  when  nitrate  is  the  sole 
source  of  nitrogen,  since  the  enzymic  reduction  of  nitrate  by  nitrate  reduc- 
tase involves  molybdate  as  a  prosthetic  group  (Higgins  et  al.,  1956  b).  Tung- 
state is  able  to  compete  with  molybdate  and  inhibition  of  growth  occurs 
when  the  (tungstate)/ (molybdate)  ratio  is  20.  Azotobacter  vinelandii  is  like- 
wise inhibited  by  tungstate  when  nitrogen  or  nitrate  is  the  source  of  amino 
acids  and  proteins,  but  not  when  ammonia  is  provided.  The  uptake  of  Mo^^ 
by  the  cells  is  also  depressed  by  tungstate.  The  ability  of  tungstate  to  inhibit 
growth  is  dependent  on  the  level  of  molybdate  in  the  medium  and  it  requires 
rather  high  ratios  of  (tungstate)/ (molybdate)  to  inhibit  well  (Bulen,  1961). 


MISCELLANEOUS   ANALOG   INHIBITIONS  615 

The  question  of  the  site,  or  sites,  of  tungstate  inhibition  is  not  yet  settled. 
Bulen  believes  that  the  primary  effect  is  a  depression  of  molybdate  uptake 
and  has  provided  evidence  that  that  there  is  no  antagonism  of  the  enzymi- 
caUy  functioning  molybdate.  The  growth  of  the  crown-gall  organism  Agro- 
bacterium  tumefaciens  is  also  inhibited  in  nitrate  medium  (about  50%  at 
0.05  milf),  but  when  ammonia  is  added  there  is  only  slight  inhibition  at 
1  ToM  (Kurup  and  Vaidyanathan,  1963).  Molybdate  antagonizes  the  growth 
depression.  A  decrease  in  nitrate  reductase  activity  during  the  inhibition 
was  observed.  Before  the  problem  of  the  site  of  inhibition  can  be  finally 
settled,  more  work  must  be  done  on  isolated  molybdenum-dependent  en- 
zymes. The  NADH-dependent  nitrate  reductase  from  wheat  is  not  inhibited 
by  1  mM  tungstate  (Spencer,  1959).  In  any  event,  this  represents  a  unique 
type  of  competitive  inhibition  which  is  basically  due  to  the  similar  structures 
and  properties  of  tungstate  and  molybdate. 

Inhibition   of  Penicillinases 

Some  of  the  data  are  summarized  in  Table  2-38,  and  it  is  evident  that  the 
penicillinases  from  different  bacteria  exhibit  various  patterns  of  sensitivity. 
In  particular,  the  gram-negative  and  gram-positive  organisms  possess  dif- 
ferent types  of  enzyme,  the  former  usually  being  more  sensitive  to  penicillin 
analogs.  The  inhibitions  are  generally  competitive.  The  exact  mechanism  of 
the  inhibition,  however,  is  not  clear,  since  there  is  some  evidence  that  the 
analogs  alter  the  configuration  of  the  enzyme  (Garber  and  Citri,  1962;  Citri 
and  Garber,  1963).  In  the  first  place,  the  analogs  accelerate  the  tempera- 
ture inactivation  of  penicillinase  and,  in  the  second  place,  the  inhibition  by 
6-(2,6-dimethoxybenzamido)penicillanate  is  accompanied  by  the  appear- 
ance of  groups  sensitive  to  iodine.  These  penicillinase  inhibitors  are  of  use 
in  determining  the  mechanism  of  penicillin  resistance  in  bacteria;  if  the 
resistance  is  due  to  the  increased  synthesis  of  penicillinase,  the  inhibitor 
wiU  abolish  the  resistance.  (Hamilton-Miller  et  al.,  1964). 

Inhibition  of  Cycle  Enzymes  by  y-Hydroxy-a-ketoglutarate 

Glyoxylate  and  oxalacetate  condense  under  physiological  conditions  to 
form  a  product  which  inhibits  certain  steps  in  the  tricarboxylate  cycle. 
The  reaction  is  quite  rapid  at  20°  and  pH  7.4,  and  since  both  glyoxylate 
and  oxalacetate  are  produced  normally  in  cells,  it  was  of  some  interest  to 
study  the  properties  of  this  condensation  product,  which  was  considered  to 
be  «-hydroxy-/?-oxalosuccinate  by  Ruffo  et  al.  (1962  a).  It  was  found  to  be 
competitive  with  respect  to  citrate  and  cis-aconitate,  and  to  inhibit  50% 
at  concentrations  around  0.12  mM.  When  glyoxylate  is  added  to  mito- 
chondria there  is  respiratory  inhibition  and  some  accumulation  of  citrate 
if  oxalacetate  is  present,  which  is  due  to  the  inhibition  of  aconitase  and 


616  2.  ANALOGS  OF  ENZYME  REACTION  COMPONENTS 

isocitrate  dehydrogenase  (Ruffo  et  al.,  1962  b;  Ruffo  and  Adinolfi,  1963). 
If  no  oxalacetate  is  present,  glyoxylate  directly  inhibits  the  oxidations  of 
a-ketoglutarate  and  succinate,  so  that  two  sites  of  cycle  inhibition  can  occur, 
one  by  glyoxylate  alone  and  one  by  its  condensation  product  with  oxalace- 
tate. Payes  and  Laties  (1963)  have  claimed  that  the  of-hydroxy-/?-oxalosuc- 
cinate  (oxalomalate)  initially  formed  in  the  condensation  reaction  is  rapid- 
ly decarboxylated  to  y-hydroxy-or-ketoglutarate,  which  is  the  actual  in- 
hibitor. y-Hydroxy-a-ketoglutarate  competitively  inhibits  yeast  aconitase 
[K^  =  0.14  mM),  potato  a-ketoglutarate  dehydrogenase  {K^  =  0.7  mM), 
and  isocitrate  dehydrogenase.  It  is  also  moderately  inhibitory  to  the  res- 
piration of  potato  slices,  35%  depression  being  observed  with  2  mM  and 
71%  with  5  mM.  It  is  rather  surprising  that  it  does  not  serve  as  a  sub- 
strate for  a-ketoglutarate  dehydrogenase  and  enter  into  the  sequence  of 
reactions  involving  coenzyme  A,  as  does  y-methyl- y-hydroxy-or-ketogluta- 
rate (parapyruvate). 


CHAPTER  3 


DEHYDROACETATE 


Since  this  substance  inhibits  succinate  dehydrogenase  and  may  exert  some 
of  its  effects  because  of  a  structure  analogous  to  certain  cellular  metabolites, 
it  is  appropriate  to  discuss  it  at  this  time,  although  little  of  its  basic  meta- 
bolic effects  is  understood.  In  a  study  of  the  formation  of  ethyl  acetoacetate, 
Geuther  (1866)  found  that  distillation  of  this  substance  yielded  a  crystalline 
material,  to  which  he  gave  the  name  dehydroacetic  acid.  It  was  soon  shown 
to  contain  a  pyran  ring  structure  (Feist,  1890)  and  has  been  used  extensively 
in  organic  preparative  procedures.  The  effects  on  biological  systems  were  not 
studied  until  Brodersen  and  Kjaer  (1946)  in  Copenhagen  investigated  a  se- 
ries of  unsaturated  lactones  for  antibacterial  activity.  They  reasoned  that 
several  such  compounds  are  antimicrobial  —  anemonin  from  species  of  Ra- 
nunculaceae,  parasorbic  acid  from  the  mountain  ash  (Sorbus),  Dicumarol 
from  sweet  clover,  penicillic  acid  from  certain  species  of  Penicillium,  kojic 
acid  from  Aspergillus,  patulin  (clavacin)  from  various  fungi,  and  others  — 
and  that  a  correlation  between  activity  and  the  — 0 — CO — C=C —  struc- 
ture might  exist: 


Q  0 

^O       o 


OH 


H,C 


.CH3 
^CH, 


OCH, 


Anemonin 


Parasorbic  acid 


Penicillic  acid 


O^     .OH 


O^    ^CH,OH 


Patulin 


Kojic  acid 


617 


618  3.  DEHYDROACETATE 

Dehydroacetic  acid  was  found  to  be  rather  inactive  against  most  of  the  25 
types  of  bacteria  studied  and  no  further  mention  was  made  of  it.  McGowan 
et  al.  (1948)  investigated  80  compounds  with  ethylenic  linkages  for  the  pur- 
pose of  correlating  fungistatic  activity  with  the  abihty  of  the  substituents 
to  withdraw  electrons  from  these  double  bonds,  and  found  dehydroacetate 
to  exert  very  little  effect.  An  investigation  stimulated  by  the  antibacterial 
effects  of  usnic  acid  from  lichens  led  Ukita  et  al.  (1949)  to  examine  dehydro- 
acetate, which  they  found  to  inhibit  staphylococci  and  mycobacteria  slight- 
ly; however,  other  substances  were  more  potent  and  of  greater  interest. 
The  Dow  Chemical  Company  meanwhile  had  been  studying  the  antimicro- 
bial action  of  dehydroacetate  and  on  June  28,  1949  issued  three  patents  for 
its  use  in  food  preservation.*  These  were  based  on  work  started  in  1946  in 
cooperation  with  the  Department  of  Pharmacology  of  the  University  of 
Michigan  Medical  School,  the  results  of  which  were  published  in  a  series  of 
papers  in  1950.  Almost  all  of  our  present  basic  knowledge  of  dehydroacetate 
stems  from  this  work  and  essentially  no  fundamental  biochemical  reports 
have  been  made  since,  although  a  great  many  papers  on  its  practical  use 
appear  annually. 

CHEMICAL  PROPERTIES 

The  structure  of  dehydroacetic  acid  was  debated  for  many  years  until 
Rassweiler  and  Adams  (1924)  proved  that  the  formula  suggested  by  Feist 
(1890)  is  basically  correct.  The  dipole  moment  of  2.83  was  claimed  by  Le- 


Dehydroacetic  acid 

Fevre  and  LeFevre  (1937)  to  be  consistent  with  this  structure  if  restrictions 
were  imposed  on  the  rotation  of  the  acetyl  group.  The  formula  is  commonly 
written  in  the  keto  form,  but  it  is  likely  that  a  tautomeric  equilibrium  with 
the  following  enolic  forms  occurs: 


H,c.       o      ^o  H,C 


(a)  (b)  (c) 

*  Dehydroacetate  is  actually  a  rather  weak  inhibitor  of  microbial  growth  but  can 
be  used  to  ^reserve  food  because  it  is  so  little  toxic  to  humans. 


CHEMICAL   PROPERTIES  619 

Hydrogen  bonding  of  the  type  suggested  by  Forsen  and  Nilsson  (1961)  is 
represented;  this  undoubtedly  occurs  in  the  pure  form  but  may  not  in 
aqueous  solution.  Proton  magnetic  resonance  and  infrared  spectra  suggest 
structure  (b),  but  the  other  forms  are  not  excluded.  The  restriction  of  the 
acetyl  group  rotation  may  derive  from  such  hydrogen  bonding. 

Ionization 

It  is  important  with  regard  to  the  action  of  dehydroacetate  on  enzymes 
and  its  penetration  into  cells  to  determine  the  predominant  forms  in  aqueous 
solution  at  physiological  pH.  It  is  usually  stated  that  dehydroacetic  acid  is 
a  very  weak  acid,  and  Wolf  and  Westveer  (1950)  remarked  that  it  would 
exist  primarily  in  the  ionized  state  at  pH  9.  The  acidic  property  is  the  result 
of  enolization  and  each  of  the  enolic  forms  shown  above  could  lose  a  proton 
to  form  the  corresponding  anion.  However,  once  this  occurs,  resonance  be- 
tween the  three  forms  is  possible,  stabilizing  the  anion  and  increasing  the 
acidity  and,  furthermore,  producing  a  more  diffuse  negative  charge.  In  order 
to  determine  the  state  of  the  inhibitor  under  physiological  conditions,  titra- 
tions were  done,  starting  at  pH  3.64  (the  pH  of  a  saturated  solution).  It 
was  found  that  two  equivalents  of  base  are  taken  up  between  pH  4.2  and 
6.4,  with  a  mean  pK^  of  5.20,  so  that  around  pH  7  dianions  of  the  a-pyran 
and  y-pyrone  type  must  be  present. 


H,C    ^O^     ^O  H,C 


O 


-CH, 


General   Properties 

Dehydroacetic  acid  has  an  absorption  peak  around  313  m//  (Calvin  et  al., 
1941),  sublimes  at  109°,  is  fairly  soluble  in  organic  solvents  but  poorly  sol- 
uble in  water  (around  0.25%  at  37°).  The  sodium  salt,  however,  is  quite 
soluble  in  water  (33%)  (Wolf,  1950)  and  sufficiently  stable  in  solution  for 
most  purposes.  It  may  be  catalytically  hydrogenated  with  PtOa  to  give  the 
corresponding  3-ethyl  compound  (Malachowski  and  Wanczura,  1933),  or 
with  Ni  under  pressure  to  yield  more  completely  hydrogenated  forms,  cor- 
responding to  the  uptake  of  3-5  moles  of  Hg  (Adkins  et  al.,  1931). 

Synthesis 

Dehydroacetic  acid  was  first  obtained  by  Geuther  (1866)  by  distilHng  the 
ethyl  ester  of  acetoacetate,  and  Conrad  (1874)  found  that  a  reasonable  yield 
could  be  obtained  by  heating  this  substance  under  pressure.  The  present 


620  3.  DEHYDROACETATE 

method  of  synthesis  involves  heating  ethyl  acetoacetate  with  a  small  amount 
of  NaHCOg  at  200o-210o  for  7-8  hr  with  subsequent  distillation  at  1280-140° 
in  vacuo,  the  yield  being  over  50%  (Arndt,  1955).  It  is  also  formed  by  the 
tetramerization  of  ketene,  and  by  suitable  catalytic  means  may  be  almost 
quantitatively  obtained  from  diketene  (Steele  et  al.,  1949),  the  reaction 
presumably  proceeding  through  the  enolized  forms  of  the  diketene.  The 
simplest  method  of  purification  of  dehydroacetic  acid  is  probably  recrys- 
tallization  from  ethanol. 

Estimation 

Two  methods  suitable  for  tissue  analyses  were  developed  by  Woods  et  al. 
(1950).  A  colorimetric  test,  based  on  the  reaction  of  the  acetyl  group  with 
salicylaldehyde  in  alkaline  solution  to  give  a  red-orange  color,  is  sensitive 
in  the  range  10-200  //g.  The  spectrophotometric  test,  based  on  absorption 
at  312  m//,  is  approximately  10-fold  more  sensitive.  Both  tests  depend  on 
the  proper  pre  treatment  and  extraction  of  the  tissue  samples,  since  neither 
test  is  particularly  specific.  The  spectrophotometric  tests  give  the  best  re- 
coveries and  are  preferable  for  most  analyses. 


INHIBITION  OF  ENZYMES 

Dehydroacetate  at  concentrations  between  2.3  and  93  vaM  progressively 
depresses  the  oxygen  uptake  of  slices  and  minces  of  cerebral  cortex  and 
kidney  respiring  endogenously,  but  the  excess  oxygen  uptake  following  ad- 
dition of  various  substrates  is  inhibited  only  in  the  case  of  succinate  (Seevers 
et  al.,  1950).  This  suggested  that  dehydroacetate  might  inhibit  succinate 
oxidase  and  thus  this  enzyme  was  examined  in  some  detail. 

Succinate  Oxidase 

Inhibition  of  succinate  oxidase  is  proportional  to  the  logarithm  of  the 
dehydroacetate  concentration  from  27  to  77%  inhibition  (Seevers  et  al., 
1950).  Although  no  data  were  given,  it  was  stated  that  the  substrate  con- 
centration has  little  or  no  effect  on  the  degree  of  inhibition,  indicating  it  is 
not  competitive.  If  this  is  true  (see  page  621),  if  ^  would  be  around  11.6  xaM 
indicating  much  less  affinity  of  the  enzyme  for  dehydroacetate  than  for  mal- 
onate.  The  site  of  inhibition  in  the  succinate  oxidase  sequence  was  determin- 
ed in  two  types  of  experiment.  In  the  first,  the  cytochrome  system  was  block- 
ed by  cyanide  and  cresyl  blue  added  as  a  hydrogen  carrier;  dehydroacetate 
inhibited  this  system  to  the  same  degree  as  the  normal  one,  indicating  the 
action  not  to  be  on  the  cytochrome  system  (see  accompanying  tabulation). 
In  the  second,  the  cytochrome  system  was  studied  directly  and  no  inhibition 


INHIBITION   OF   ENZYMES  621 


Cyanide 

Cresyl  blue 

Dehydroacetate 

Oj  Uptake 

°/o  Inhibition  by 

(mM) 

(mM) 

(mJ/) 

{/a/30  min) 

dehydroacetate 

165 

2 

— 

— 

0 

— 

2 

6 

— 

88.5 

— 

— 

— 

9.3 

84 

49 

2 

6 

9.3 

43.5 

51 

by  dehydroacetate  was  observed.  It  was  concluded  that  dehydroacetate  acts 
on  succinate  dehydrogenase.  Although  this  is  probably  true  it  is  evident 
that,  according  to  the  modern  elaboration  of  succinate  oxidase,  other  sites 
are  possible. 

Evidence  that  dehydroacetate  does  not  inhibit  by  reacting  with  the  SH 
groups  of  succinate  dehydrogenase  was  obtained.  First,  the  inhibition  is  al- 
most instantaneous  and  readily  reversible,  unlike  inhibitions  with  most  SH 
reagents.  Second,  cysteine,  glutathione,  and  dimercaprol  are  unable  to  pro- 
tect the  enzyme  against  dehydroacetate.*  Third,  no  inhibition  of  urease 
was  observed,  this  enzyme  being  sensitive  to  most  SH  reagents;  indeed 
stimulation  was  observed.  All  in  all,  one  must  conclude  that  the  possibility 
of  the  reaction  of  dehydroacetate  with  SH  groups  has  not  been  eliminated, 
although  there  is  little  positive  evidence  for  such  a  mechanism. 

The  possibility  of  competitive  inhibition  of  succinate  dehydrogenase  can- 
not be  eliminated  since  no  data  were  given  for  the  statement,  "Increasing 
the  substrate  concentration  does  not  appreciably  alter  the  degree  of  inhi- 
bition." If,  for  example,  a  succinate  concentration  of  50  mM  was  used 
(which  is  the  only  concentration  mentioned  in  the  paper)  with  dehydro- 
acetate at  9.3  mM,  increasing  the  succinate  to  100  or  200  m.M  would  not 
be  expected  to  reduce  the  inhibition  markedly.  It  is  interesting  to  speculate 
that  the  dianionic  forms  of  dehydroacetate  have  a  basically  similar  charge 
distribution  to  malonate.  However,  due  to  resonance  the  charge  magnitude 

H        H 
^   ^C^      CH,  O^  X  ^^ 

c       c  c       c 

o~      o"  o'      o" 

Dehydroacetate  Malonate 

*  Unfortunately  the  cysteine  and  glutathione  were  used  at  only  about  one  fifth  the 
dehydroacetate  concentration,  so  that  even  total  reaction  of  the  inhibitor  would  have 
reduced  the  inhibition  relatively  little,  actually  about  5%.  Cysteine  reduced  the  inhi- 
bition around  5%  from  the  predicted  value  but  glutathione  did  not.  Cavallito  and 
Haskell  (1945)  mentioned  that  dehydroacetate  does  not  react  with  cysteine. 


622  3.  DEHYDROACETATE 

on  dehydroacetate.  might  be  lower  than  on  malonate.  The  succinate  dehy- 
drogenase from  calf  thymus  nuclei  is  inhibited  somewhat  more  potently  by 
dehydroacetate  at  pH  6.6  than  at  pH  7.6  (see  accompanying  tabulation) 


Yo  Inhibition 


I) 

pH6.6 

pH7.6 

1 

33 

25 

5 

56 

42 

10 

67 

47 

(McEwen  et  al.,  1963  a).  This  does  not  fit  the  dianion  inhibition  theory  very 
well,  since  at  pH  7.6  there  should  be  more  of  the  dianion  than  at  pH  6.6. 
One  cannot  attribute  the  change  in  inhibition  to  the  ionization  of  enzyme 
groups  because  malonate  inhibits  better  at  the  higher  pH  from  the  limited 
data  provided.  Unfortunately  the  succinate  concentration  was  unvaried 
from  20  mM  and  the  formal  nature  of  the  inhibition  remains  unknown. 
The  mechanism  of  the  inhibition  of  succinate  dehydrogenase  is  thus  at 
present  unsolved. 

One  further  experiment  deserves  brief  mention.  It  was  claimed  that  al- 
though malonate  protects  succinate  oxidase  from  SH  reagents,  it  does  not 
protect  against  dehydroacetate  (see  accompanying  tabulation).  First,  one 


Dehydroacetate  (milf )       Malonate  {mM) 


%  Inhibition 
of  O2  uptake 


9.3 
9.3 


— 

49 

0.33 

33 

0.33 

65 

would  not  expect  malonate  at  a  concentration  inhibiting  only  33%  to  pro- 
tect very  much.  Second,  the  inhibition  given  by  both  inhibitors  is  exactly 
what  would  be  predicted  if  both  acted  at  the  same  site  on  the  enzyme.  No 
conclusions  as  to  the  mechanism  of  inhibition  can  be  deduced  from  this  ex- 
periment. 

Other  Enzymes 

Dehydroacetate  has  no  effect  on  cholinesterase  up  to  20  mM  (Seevers  et 
al.,  1950),  or  on  pepsin,  amylase,  and  trypsin  at  8.5  mM  (Bauer  and  La  Sala, 
1956),  while  urease  is  stimulated  by  concentrations  up  to  93  mM  and  pan- 


EFFECTS    ON    RESPIRATION    AND    GLYCOLYSIS 


623 


creatic  lipase  is  stimulated  at  8.5  mM.  The  ATPase  of  pea  mitochondria  is 
also  stimulated  by  dehydroacetate,  phosphate  splitting  being  increased 
around  200%  by  1  mM  (Forti,  1957),  which  might  be  related  to  the  un- 
coupling action  reported  by  MarreeiaZ.  (1956).  Catalase  is  inhibited  weakly 
(K^  =  22  jnM)  by  dehydroacetate  compared  with  other  organic  acids  (Liick, 
1957).  The  enzymes  responsible  for  the  destruction  of  mitomycin  in  Strepto- 
myces  mycelia  are  not  affected  by  4.8  mM  dehydroacetate  (Gourevitch  et 
al.,  1961).  It  is  difficult  to  explain  some  of  the  actions  on  metabolism  with 
this  limited  amount  of  information. 


EFFECTS   ON    RESPIRATION   AND  GLYCOLYSIS 

The  effects  of  dehydroacetate  on  the  endogenous  respiration  of  minces  of 
various  rat  tissues  are  shown  in  Fig.  3-1.  Brain  and  kidney  respiration  is 
progressively  depressed  but  muscle  is  anomalous  in  that  marked  stimulation 
is  observed  at  high  dehydroacetate  concentrations,  while  liver  is  stimulated 


■3  -2 

LOG   (DEHYDROACETATE) ( Ml       


Fig.  3-1.   Effects  of  dehydroacetate  on  the  respi- 
ration of  rat  tissue  minces  measured  over  a  period 
of  2  hr.   (From  Seevers  et  al.,  1950.) 


moderately  at  all  concentrations  used.  Brain  slices  respond  as  do  the  minces, 
but  liver  slices  are  unaffected  by  dehydroacetate  up  to  9.3  mM  and  are 
then  depressed  at  higher  concentrations.  Mudge  (1951)  found  10  mM  dehy- 
droacetate to  depress  rabbit  kidney  slice  respiration  25%,  which  is  similar 
to  the  inhibition  reported  by  Seevers  et  al.  (1950)  in  rat  kidney.  The  stim- 
ulation of  respiration  in  muscle  and  liver  might  result  from  metabolism 


624  3.  DEHYDROACETATE 

of  dehydroacetate;  this  will  be  discussed  later  (page  629),  but  Shideman 
et  al.  (1950  b)  do  not  think  it  is  likely.  The  endogenous  respiration  in  such 
minces  and  slices  is  poorly  understood,  so  that  it  is  difficult  to  interpret 
these  results,  and  the  data  when  glucose  is  present  are  inconsistent.  It  is 
also  difficult  to  relate  these  changes  in  respiration  to  inhibition  of  succinate 
oxidase  (the  effects  of  malonate  and  dehydroacetate  are  quite  different), 
and  such  statements  as  "It  appears  probable  that  the  manifestations  of 
toxicity  result  largely,  if  not  exclusively,  from  a  specific  type  of  chemical 
(or  physicochemical)  action  involving  interference  with  oxidative  or  other 
enzyme  mechanisms  which  proceed  by  way  of  the  Krebs  cycle"  (Shideman 
et  al.,  1950  b)  appear  to  have  little  basis,  particularly  since  the  effects  of 
dehydroacetate  on  the  operation  of  the  cycle  (as  in  mitochondrial  prepara- 
tions) have  not  been  studied.  It  is  significant  that  dehydroacetate  at  4.7- 
9.3  TaM  stimulates  the  anaerobic  glycolysis  in  rat  brain  mince  40-50% 
(See vers  et  al.,  1950),  an  effect  greater  than  any  observed  on  respiration. 
On  the  other  hand,  50  mM  dehydroacetate  inhibits  the  formation  of  C^^Og 
from  glucose-6-C^*  87%  in  suspensions  of  isolated  thymus  nuclei,  simultane- 
ously the  O2  uptake  being  depressed  only  14%  and  the  ATP  level  falling 
27%  (McEwen  et  al,  1963  b).  Malonate  at  10  mM  has  very  little  effect  and 
this  was  attributed  to  a  failure  to  penetrate  into  the  nuclei;  dehydroacetate 
either  penetrates  better  than  malonate  or  exerts  an  effect  other  than  inhi- 
bition of  succinate  oxidation. 


EFFECTS  ON  TISSUE  FUNCTIONS 

Dehydroacetate  in  the  whole  animal  produces  changes  in  central  nervous 
system,  cardiovascular,  and  renal  functions.  Only  the  renal  effects  have  been 
investigated  in  detail.  In  addition,  the  actions  on  the  isolated  intestine  have 
been  studied  relative  to  the  metabolic  disturbances  produced. 

Intestine 

The  contractile  amplitude  of  isolated  rabbit  intestine  is  depressed  slightly 
by  1  mM  and  markedly  by  10  vclM  dehydroacetate  (see  tabulation  below) 


Dehydroacetate 
(mM) 

Substrate 

%   Inhibition 
of   amplitude 

1 

Acetate 

2 

1 

Glucose 

8 

10 

Acetate 

72 

10 

Glucose 

54 

EFFECTS   ON   TISSUE    FUNCTIONS  625 

(Weeks  et  al.,  1950).  When  the  intestine  is  allowed  to  contract  for  30-60  min 
in  the  absence  of  substrate  the  amplitude  is  reduced  to  15-35%  of  normal. 
The  addition  of  glucose,  acetate,  or  pyruvate  allows  recovery,  essentially 
complete  in  the  case  of  glucose.  Dehydroacetate  at  10  mM  effectively  pre- 
vents this  recovery  with  acetate  and  pyruvate,  but  only  partially  counter- 
acts the  effect  of  glucose.  This  was  taken  to  mean  that  dehydroacetate  blocks 
the  cycle  preferentially.  Malonate  neither  depresses  the  amplitude  in  the 
presence  of  substrates,  nor  prevents  the  recovery  of  substrate-depleted  in- 
testine upon  addition  of  substrates.  This  could  be  related  to  the  poor  pen- 
etration of  malonate  into  the  cells.  Dehydroacetate,  even  though  it  may 
have  a  double  negative  charge,  might  penetrate  better  than  malonate  be- 
cause the  lipophilic  fraction  of  the  molecule  is  greater  (sodium  dehydroace- 
tate is  fairly  soluble  in  a  number  of  organic  solvents). 

Heart 

Dehydroacetate  given  intravenously  to  dogs  causes  some  slowing  of  the  heart 
rate  at  a  dose  of  300  mg/kg  (Seevers  et  al,  1950),  but  in  general  the  effects  on 
the  cardiovascular  system  are  minimal.  Like  many  substances  with  the 
— CH  =  CH — CO —  grouping,  dehydroacetate  exerts  a  positive  inotropic  ac- 
tion on  hypodynamic  cat  papillary  (Bennett  et  al.,  1958).  Although  not  so 
potent  as  many  other  compounds,  it  is  active  at  5  mM  and  rated  the  same 
as  /3-angelicalactone.  Isolated  rat  atria  are  depressed  quite  markedly  by 
5-10  mM  dehydroacetate,  and  to  some  extent  even  by  1  mM,  and  this  is 
accompanied  by  rather  unique  effects  on  the  membrane  potentials  (Webb 
and  Hollander,  1959).  The  resting  and  action  potential  magnitudes  are  re- 
duced more  strikingly  than  with  most  metabolic  inhibitors.  However,  the 
action  potential  duration  is  actually  increased,  due  to  a  slowing  of  repolar- 
ization, an  effect  observed  with  no  other  inhibitor.  The  depolarization  rate 
is  unaffected,  so  that  the  moderate  slowing  of  conduction  noted  must  be 
related  to  the  reduced  action  potential.  Malonate  at  higher  concentration 
scarcely  alters  the  properties  of  such  atria,  again  indicating  a  difference 
either  in  penetration  or  in  action. 

Renal  Transport 

When  dehydroacetate  is  injected  into  dogs  in  amounts  sufficient  to  give 
plasma  concentrations  of  20-25  mg%  (1.2-1.5  mM),  the  renal  tubular  trans- 
port of  certain  substances  is  markedly  depressed  —  p-aminohippurate  81%, 
phenolsulfonphthalein  80%,  penicillin-G  60%,  and  iV-methylnicotinamide 
77%,  while  the  transport  of  glucose,  creatinine,  and  phosphate  is  unaltered 
(Shideman  et  al.,  1950  b).  Since  diuresis  is  observed  after  dehydroacetate, 
it  is  likely  that  water  and  ion  transport  is  also  affected  to  some  extent.  It 
is  thus  clear  that  certain  transport  systems  are  inhibited  and  others  are 


626  3.  DEHYDRO ACETATE 

untouched  by  dehydroacetate.  The  mechanism  is  entirely  tubular  and  no 
changes  in  femoral  flow,  renal  blood  flow,  or  glomerular  filtrate  rate  occur 
(Shideman  and  Rene,  1951  b).  The  action  is  similar  to  that  of  carinamide 
2)-(benzylsulfonamido)benzoate,  a  previously  used  blocker  of  penicillin  ex- 
cretion. Indeed,  dehydroacetate  at  0.5  g  every  6  hr  prolongs  penicillin  blood 
levels  in  patients  (Schimmel  et  at.,  1956). 

The  active  accumulation  of  phenolsulfonphthalein  (Rathbun  and  Shide- 
man, 1951),  phenol  red  (Shideman  and  Rene,  1951  b),  and  p-aminohippu- 
rate  (Shideman  and  Rene,  1951  b;  Farah  and  Rennick,  1956)  in  kidney 
slices  is  readily  inhibited  by  dehydroacetate.  The  effect  on  ^J-aminohippu- 
rate  uptake  is  shown  in  Fig.  1-18,  from  which  it  is  seen  that  50%  inhibition 
is  given  by  0.14  mM  in  dog  kidney  slices.  The  accumulation  of  tetraethylam- 
monium  ion  is  completely  resistant  to  dehydroacetate,  and  it  is  believed  that 
the  transport  of  this  ion  is  not  dependent  on  the  cycle  (Farah  and  Rennick, 
1956;  Farah,  1957). 

It  appears  that  certain  renal  transport  mechanisms  are  more  sensitive  to 
dehydroacetate  than  any  other  cell  functions  examined.  The  question  as  to 
the  relation  of  this  inhibition  to  succinate  oxidase  or  cycle  depression  is  dif- 
ficult to  resolve.  Dehydroacetate  might  block  transport  by  acting  directly 
on  the  carrier  system  or  by  reducing  the  energy  available  for  the  transport, 
Shideman  and  Rene  (1951  b)  incline  to  the  latter  view  and  attribute  the 
inhibition  to  an  action  on  the  cycle.  The  evidence  for  this  comes  partially 
from  the  observation  that  high  concentrations  of  acetate  are  able  to  coun- 
teract the  effects  of  dehydroacetate,  both  in  vivo  and  in  slices  (Stoneman 
et  al.,  1951;  Rathbun  and  Shideman,  1951;  Shideman  and  Rene,  1951  a). 
However,  the  ability  of  acetate  to  reverse  an  inhibition  is  not  evidence  for 
an  action  on  the  cyle,  much  less  on  succinate  oxidase;  indeed,  the  opposite 
might  be  justifiably  concluded.  Furthermore,  there  is  not  a  good  correlation 
between  the  activity  in  depressing  p-aminohippurate  transport  and  the  in- 
hibitory potency  on  succinate  oxidation,  especially  when  carinamide  is  con- 
sidered, this  substance  being  a  weak  succinate  oxidase  inhibitor  but  a  more 
potent  transport  inhibitor  than  dehydroacetate.  Of  course,  different  pene- 
trabilities into  the  renal  cells  may  account  in  part  for  this  lack  of  correlation. 

Ion  transport  in  kidney  slices  is  also  inhibited  by  dehydroacetate  (Mudge, 
1951).  Rabbit  kidney  slices  were  leached  in  0.15  M  NaCl  to  lower  the  in- 
tracellular K+,  and  then  incubated  in  10  mM  K+  with  10  vaM  acetate  as 
the  substrate,  during  which  period  the  lost  K+  is  regained  and  the  excess 
intracellular  Na+  is  pumped  out.  Dehydroacetate  at  10  mM  inhibits  this 
K+-Na+  exchange  81%,  while  simultaneously  the  respiration  is  inhibited 
only  25%.  The  action  of  dehydroacetate  on  the  pH-regulating  exchanges 
of  the  kidney  is  not  known,  but  may  be  important  in  contributing  to  the 
acidosis  observed  in  whole  animals,  and  indirectly  in  the  effects  on  certain 
tissues  such  as  the  central  nervous  system. 


EFFECTS    ON    THE    WHOLE    ANIMAL  627 

EFFECTS  ON  THE  WHOLE  ANIMAL 

The  potential  use  of  dehydroacetate  as  a  food  preservative  led  Spencer 
et  al.  (1950  a,b)  of  the  Dow  Chemical  Company,  and  Seevers  et  al.  (1950)  of 
the  University  of  Michigan,  to  investigate  the  effects  on  various  animals  and 
man  when  administered  by  different  routes  for  varying  durations  of  time. 
These  studies  were  very  thorough  and  our  knowledge  at  this  level  of  action 
is  better  than  for  most  inhibitors.  We  shall  summarize  only  the  more  im- 
portant aspects  relative  to  the  metabolic  disturbances  produced. 

Acute  Toxicity 

The  earliest  evidence  of  toxicity  in  rats,  dogs,  and  monkeys  when  de- 
hydroacetate is  given  by  any  route  is  loss  of  appetite.  As  the  dosage  is  in- 
creased, various  sjTnptoms  related  mainly  to  the  central  nervous  system 
appear:  ataxia,  salivation,  emesis,  incoordination,  weakness,  and  stupor, 
followed  by  muscle  twitching  and  a  gradual  increase  in  muscle  tone,  passing 
into  clonic  and  tonic  convulsions  which  persist  until  death,  which  is  attri- 
buted to  respiratory  i)aralysis.  During  the  later  phases,  convulsions  may  be 
initiated  by  excess  sensory  stimulation,  indicating  a  general  increase  in  the 
reflex  excitability.  The  effects  are  produced  rather  slowly  and  death  occurs 
usually  after  24-72  hr.  The  acute  oral  toxicity  in  rats  is  given  as:  LD^.i  = 
0.52  g/kg,  LDi6  =  0.80  g/kg,  LD50  =  1.0  g/kg,  LDg^  =  1.23  g/kg,  and 
LD99.9  =  1.92  g/kg. 

The  initial  effects  of  intravenous  injection  (0.3-0.4  g/kg)  are  probably 
related  to  the  temporary  disturbance  in  blood  pH  if  the  solutions  are  not 
neutralized.  The  alterations  in  nervous  system  function  may  not  be  specific 
and  due  to  the  direct  action  of  dehydroacetate  on  the  nerve  cells,  but  de- 
pendent on  the  developing  acidosis.  An  initial  respiratory  alkalosis  in  dogs 
is  soon  followed  by  a  shift  toward  metabolic  acidosis,  compensated  at  first 
but  later  becoming  uncompensated.  As  the  plasma  pH  suddenly  drops  to 
levels  approaching  7,  with  simultaneous  decreases  in  plasma  bicarbonate 
and  Pqq  ,  convulsions  occur.  We  have  noted  that  renal  transport  is  more 
sensitive  to  dehydroacetate  than  is  nerve  or  muscle  respiration,  especially 
in  the  presence  of  glucose.  The  plasma  levels  of  dehydroacetate  in  acute 
poisoning  are  probably  between  1  and  3  mM  in  most  cases,  which,  coupled 
with  the  possible  concentration  in  the  kidney,  could  easily  disturb  renal 
function  seriously.  The  concentration  in  the  central  nervous  system  during 
poisoning  appears  to  be  quite  low  (see  page  631).  All  of  this  points  to  an 
indirect  action  on  the  nervous  system  and  perhaps  a  primary  renal  site  for 
the  toxicity. 

Chronic  Toxicity 

Rats  fed  diets  containing  0.02-0.10%  dehydroacetic  acid  for  2  years 
showed  no  obvious  adverse  effects  on  growth,  mortality,  hematology,  organ 


628 


3.  DEHYDROACETATE 


weights,  or  tissue  cytology,  with  the  possible  exception  of  an  increase  in 
liver  fat  at  the  highest  dose  level.  Monkeys  maintained  on  oral  doses  of  50- 
100  mg/kg/day  for  a  year  likewise  showed  no  evidence  of  a  toxic  effect. 
Even  the  maximum  tolerated  dose  of  200  mg/kg/day,  producing  some  toxic 
effects  in  the  monkeys,  produced  no  pathological  changes  in  the  tissues.  One 
must  conclude  that  dehydroacetate  is  a  relatively  nontoxic  substance.  Hu- 
man subjects  can  tolerate  14-17  mg/kg/day  for  26-48  days  (approximately 
1.2  g/day  or  a  total  of  30-60  g),  maintaining  a  plasma  level  of  15-25  mg% 
(around  1  mM),  and  experience  only  occasional  anorexia  and  nausea.  Hu- 
man subjects  are  more  sensitive  to  dehydroacetate  than  are  experimental 
animals  on  a  dosage  basis,  but  the  plasma  concentrations  are  approximately 
the  same,  indicating  that  man  either  metabolizes  or  excretes  dehydroacetate 
less  readily  (see  accompanying  tabulation). 


Tolerated  levels 

Nontolerated 

levels 

Species 

Dose 
(mg/kg) 

Days 

Plasma 

cone. 

(mg%) 

Dose 

(mg/kg) 

Days 

Plasma 

cone. 

(mg%) 

Rat 

50 

730 

11- 

-17 

Dog 

50 

199 

12- 

13 

60-80 

73 

20-25 

Monkey 

50 

378 

12- 

-16 

100 

397 

20-30 

Man 

6-13 

173 

12- 

-17 

14-17 

26-48 

15-25 

Urinary   Excretion   of  Succinate 

If  dehydroacetate  is  able  to  inhibit  succinate  oxidase  in  the  whole  animal, 
one  would  expect  to  find  an  accumulation  of  succinate  in  the  body  and  an 
increased  excretion.  This  was  indeed  found,  providing  the  best  evidence 
that  such  enzyme  inhibition  actually  occurs  in  vivo.  Rats  given  600  mg/kg 
of  dehydroacetic  acid  orally  show  an  increase  in  urinary  succinate  during 
the  first  day  from  a  control  level  of  3.2  mg/day  to  24.8  mg/day  (Seevers  et 
al.,  1950).  Dogs  given  200  mg/kg  orally  for  3  days  produce  a  maximal  ex- 
cretion of  succinate  on  the  second  day  (more  than  175  mg/day  compared  to 
a  control  of  28  mg/day),  and  the  rise  is  maintained  after  cessation  of 
administration,  being  around  90  mg/day  4  days  after  the  last  dose.  One 
would  also  expect  other  acids  to  accumulate  if  the  cycle  is  depressed  and 
the  appearance  of  ketonemia,  but  this  is  not  mentioned. 


Antidotes 

Seevers  et  al.  (1950)  attempted  to  combat  the  toxic  effects  of  dehydroace- 
tate in  dogs  by  the  administration  of  glucose,  ammonium  lactate,  calcium 


DISTRIBUTION  AND  METABOLISM  629 

lactate,  ammonium  chloride,  magnesium  sulfate,  sodium  bicarbonate,  and 
dimercaprol  [ !  ],  but  no  protection  or  benefit  was  observed,  which  is  not 
very  surprising.  It  would  be  interesting  to  know  if  either  fumarate  or  malate, 
the  products  of  succinate  oxidation  and  possible  restorers  of  cycle  activity, 
is  effective.  Barbital  controls  the  convulsions  produced  by  dehydroacetate 
and  allows  recovery  from  an  otherwise  fatal  dose,  indicating  that  the  con- 
vulsions must  contribute  to  the  death  of  the  animals. 


DISTRIBUTION    AND    METABOLISM 

Whether  given  orally  or  parenterally,  during  acute  or  chronic  administra- 
tion, most  of  the  dehydroacetate  seems  to  be  metabolized  in  the  body,  since 
less  that  25%  is  found  in  the  urine  (monkeys  10%,  dogs  20%,  man  22%) 
and  only  around  5%  in  the  feces  (Shideman  et  at.,  1950  a,b).  Only  insignifi- 
cant amounts  of  conjugated  dehydroacetate  occur  in  the  urine  inasmuch  as 
2-4%  more  can  be  obtained  on  acid  hydrolysis.  It  was  not  possible  to  de- 
monstrate destruction  of  dehydroacetate  by  slices  of  rat  liver,  kidney,  or 
brain,  or  in  muscle  mince  in  incubations  up  to  4  hr,  although,  as  pointed  out, 
the  analyses  may  not  have  been  specific  enough  to  have  detected  certain 
chemical  modifications.  No  evidence  could  be  found  for  the  appearance  of 
2,6-dimethyl-l,4-pyrone  or  6-methyl-2^-pyran-2,4(3/f)-dione,  substances 
formed  in  the  chemical  degradation  of  dehydroacetate.  Nor  were  positive 
tests  for  acetomalonate  or  acetoacetate,  two  possible  metabolic  products, 
obtained  in  dog  or  human  urine.  The  stimulation  of  liver  and  muscle  respi- 
ration by  dehydroacetate  might  indicate  metabolism  of  dehydroacetate  in 
these  tissues,  but  this  is  not  at  all  certain.  However,  it  has  been  shown  that 
in  animals  with  carbon  tetrachloride  liver  damage  the  toxicity  of  dehydro- 
acetate is  increased  29%,  pointing  to  the  liver  as  at  least  one  site  for  the 
metabolism. 

The  oral  administration  to  rats  of  small  doses  (60  mg/kg)  of  dehydroace- 
tate, labeled  in  four  of  its  carbon  atoms,  leads  after  5  days  to  the  following 
distribution  of  the  label:  urine  23%,  feces  19%,  carcass  22%,  and  respira- 
tory CO2  12.4%  (Barman  et  al.,  1961).  The  urine  contains  five  labeled  sub- 
stances: unchanged  dehydroacetate  (4.7%),  hydroxydehydroacetate  (7%), 
triacetic  acid  lactone  (1.2%),  urea  (0.2%),  and  an  unknown  pyrone  meta- 
bolite (the  figures  in  parentheses  give  the  percentages  of  the  dose).  In  ad- 
dition, the  imino  derivatives  of  dehydroacetate  and  hydroxydehydroacetate 
are  found,  since  reaction  with  ammonia  occurs  in  the  urine.  The  major 
metabolic  pathway  was  postulated  to  be 

Dehydroacetate  -►  hydroxydehydroacetate  ->  triacetic  acid  lactone  -> 
acetoacetate  +  acetate 

The  hydroxylation  of  the  S-COCHg  group  occurs  in  slices  of  liver,  but  not 


630  3.  DEHYDROACETATE 

in  kidney,  muscle,  or  testis  (Barman  et  al.,  1963).  The  rabbit  apparently 
metabolizes  dehydroacetate  more  readily  than  the  rat.  The  absence  of  glu- 
curonides  or  ethereal  sulfates  in  the  urine  was  confirmed.  The  urinary  pyrone 
metabolite  is  probably  the  3-carboxylate  of  triacetic  acid  lactone. 

Dehydroacetate  is  excreted  by  the  kidney  quite  slowly.  This  could  be  due 
to  the  binding  of  a  major  fraction  in  the  plasma  to  protein,  or  to  active 
resorption  by  the  tubules;  both  actually  occur.  The  fraction  bound  to  plasma 
proteins  depends  on  the  species  and  the  dehydroacetate  concentration 
(Woods  et  al.,  1950)  (some  averages  are  shown  in  the  accompanying  tabu- 
lation). However,  even  when  the  renal  excretion  is  corrected  for  the  amount 


Plasma 

Total  plasma 

Bound  de 

hydroacetate 

Species 

protein 

dehydroacetate 

(%) 

(mg%) 

0/ 

/o 

mg/g 

Rat 

5.58 

5.7 

90.5 

0.93 

Dog 

4.28 

4.0 

58.5 

0.62 

Man 

5.15 

6.2 

98.5 

1.20 

bound,  it  is  evident  that  this  is  not  the  primary  factor  in  the  slow  excretion. 
Dehydroacetate  is  resorbed  through  the  tubules  to  about  the  same  extent 
as  water  (98-99%).  Since  the  ring  structure  of  dehydroacetate  is  identical 
to  the  pyranose  form  of  glucose,  it  was  felt  that  transport  by  the  glucose 
system  might  occur,  but  phlorizin,  at  a  concentration  that  markedly  in- 
hibits glucose  transport,  does  not  alter  dehydroacetate  resorption.  The  re- 
lationship between  the  transport  of  dehydroacetate  and  the  effects  of  dehy- 
droacetate on  other  transport  mechanisms,  if  any,  is  not  clear. 

Both  dehydroacetic  acid  and  its  sodium  salt  are  absorbed  rapidly  when 
given  orally,  peak  plasma  concentrations  occurring  in  90-120  min.  It  can 
be  detected  in  the  blood  3  4  days  after  single  doses,  and  when  administered 
chronically  many  days  are  required  for  the  plasma  level  to  drop  to  negli- 
gible concentrations.  The  distribution  of  dehydroacetate  in  the  tissues  (Ta- 
ble 3-1)  and  its  variation  with  time  illustrate  the  complexities  of  the  factors 
governing  penetration  and  binding.  The  low  concentration  in  the  central 
nervous  system  and  the  relatively  high  level  in  spinal  fluid  are  surprising. 
The  biliary  circulation  of  dehydroacetate  may  play  a  minor  role  in  retain- 
ing it  in  the  body. 

Dehydroacetate  is  secreted  in  the  saliva  at  a  reasonably  high  concentra- 
tion, an  injection  of  5  mg  intraperitoneally  in  the  rat  giving  salivary  levels 
of  0.25-0.33  mM  at  2-6  hr  (Zipkin  and  McClure,  1958),  roughly  about  half 
the  plasma  concentration.  Dehydroacetate  has  been  incorporated  into  tooth- 
pastes to  inhibit  bacterial  growth  and  caries.  However,  it  has  been  found 


ANTIMICROBIAL   ACTIVITY 


631 


Table  3-1 
Distribution  of  Dehydroacetate  in  Tissues  of  the  Dog" 


Tissue 


Dehydroacetate 

Intravenously 

(160  mg/kg) 

Oral] 

y  (80  mg/kg/day) 

At  1  hr 

(mg°o) 

At  5  hr 

(mg%) 

for  46  days 
(mg%) 

25 

21 

18 

15 

3.9 

2.4 

14 

11 

2.0 

11 

0.8 

3.1 

11 

0 

1.5 

9.7 

2.4 

0.8 

6.7 

5.5 

1.8 

6.6 

0 

0.9 

6.6 

13 

3.1 

1.6 

0 

0 

0 

— 

27 

— 

19 

16 

— 

4.4 

4.5 

Blood 

Kidney 

Intestine 

Heart 

Spleen 

Muscle 

Liver 

Cerebrum 

Lung 

Cerebellum  - 

Bile 

Spinal  fluid 

Colon 


medulla 


°  From  Woods  et  al.  (1950). 

that  dehydroacetate  in  the  diet  of  rats  markedly  potentiates  the  develop- 
ment of  caries  (Zipkin  and  McClure,  1957,  1958).  Dehydroacetate  at  0.1% 
in  a  cariogenic  diet  or  drinking  water  (corresponding  to  around  5  mg  up- 
take per  day)  increases  significantly  the  frequency  of  caries.  It  is  suspected 
that  the  cariogenic  action  of  dehydroacetate  may  be  related  to  its  secretion 
in  the  saliva,  but  the  mechanism  is  unknown. 


ANTIMICROBIAL  ACTIVITY 

Dehydroacetate  has  been  used  widely  the  past  few  years  as  a  food  pre- 
servative, especially  against  molds,  and  is  certainly  one  of  the  safest  and 
most  effective.  This  has  stimulated  extensive  work  to  determine  the  mini- 
mal growth  inhibitory  concentrations  for  various  microorganisms,  some  of 
the  results  of  which  are  summarized  in  Table  3-2.  Two  things  are  imme- 
diately evident  from  this  table.  Dehydroacetate  is  in  general  a  rather  weak 
antimicrobial  agent;  it  is  of  practical  value  because  of  its  low  toxicity.  It 


632 


3.  DEHYDROACETATE 


Table  3-2 
Antimicrobial  Activity  of  Dehydroacetate 


Organism 


Bacteria 
Aerobacter  aerogenes 
Alcaligenes  faecalis 
Bacillus  anthracis 
B.  cereus 
B.  megaterium 
B.  mesentericus 
B.  suhtilis 

Corynebacterium  diphtheriae 
Escherichia  coli 
Lactobacillus  acidophilus 
L.  brevis 
L.  casei 
L.  fermenti 
L.  plantarum 

Mycobacterium  tuberculosis 
Pseudomonas  aeruginosa 
Salmonella  pullorum 
Salmonella  typhosa 
Staphylococcus  aureus 


Vibrio  cholerae 
V.  metchnikowii 

Fungi 
Aspergillus  niger 
Botrytis  allii 
Fusarium  graminearum, 
Penicillium  digitalum 

P.  expansum 
Rhizopus  nigricans 
Trichophyton  interdigitale 
T.  mentagrophytes 

Yeasts 
Candida  albicans 
Saccharomyces  cerevisiae 


Minimal   inhibitory 

concentration 

Reference  " 

(mM) 

17.8 

(6,7) 

23.8 

(6) 

59.5 

(1) 

17.8 

(6) 

17.8 

(6) 

17.8 

(6) 

17.8 

(6) 

3.0 

(1) 

23.8 

(6) 

59.5 

(2) 

12-60 

(2) 

12-24 

(2) 

12-60 

(2) 

12 

(2) 

5.9 

(6) 

5.9 

(5) 

23.8 

(6) 

17.8 

(6) 

12 

(6) 

108 

(4) 

10 

(5) 

17.8 

(7) 

59.5 

(1) 

5.9 

(1) 

3.0 

(6) 

0.3 

(3) 

0.47 

(3) 

2.4 

(3) 

1.8 

(6,7) 

0.6 

(6) 

2.4 

(6,7) 

0.3 

(6) 

0.59 

(4) 

0.3 

(6) 

1.9 

(4) 

5.9 

(6,7) 

"  References:  (1)  Brodersen  and  Kjaer  (1946);  (2)  Fitzgerald  and  Jordan  (1953); 
(3)  McGowan  et  al.  (1948);  (4)  Stedman  et  al.  (1954);  (5)  Ukita  etal.  (1949);  (6)  Wolf 
(1950);  (7)  Wolf  and  Westveer  (1950). 


ANTIMICROBIAL   ACTIVITY  633 

is  clear  that  fungi  are  more  sensitive  to  dehydroacetate  than  are  bacteria. 
The  means  of  the  concentrations  in  the  table  are,  of  course,  not  quantita- 
tively significant,  but  show  well  the  difference:  26  mM  for  bacteria  and 
1.1  mM  for  fungi.  The  mechanism  of  growth  inhibition  is  completely  un- 
known. There  is  no  obvious  correlation  between  cycle  activity  in  these  organ- 
isms and  susceptibility  to  dehydroacetate.  The  fungi  behave  differently  than 
bacteria  with  regard  to  so  many  drugs  that  one  must  assume  basic  differ- 
ences in  metabolism  or  permeabilities,  and  it  would  be  impossible  at  this  time 
to  attribute  the  greater  sensitivity  to  dehydroacetate  to  any  one  factor. 
Permeability  seems  to  be  of  importance  in  the  action  of  dehydroacetate, 
as  indicated  by  the  effects  of  pH.  A  decrease  in  activity  with  increasing  pH 
has  been  generally  noted  (Shibasaki  and  Terui,  1953;  Bandelin,  1958),  with 
the  exception  of  Salmonella  and  Staphylococcus  (Wolf  and  Westveer,  1950). 
The  results  of  Bandelin  on  several  fungi  are  typical  (see  accompanying  ta- 
bulation). A  100-  to  200-fold  increase  in  activity  as  the  pH  is  raised  from 


Minimal  inhibitory  concentration  (mM) 

Organism 

pH  3 

pH  5 

pH  7 

pH  9 

AUernaria  solani 

0.059 

0.12 

1.18 

11.8 

Aspergillus  niger 

0.12 

0.30 

2.36 

11.8 

Chaetowium  globosum 

0.059 

0.30 

2.36 

5.9 

Penicillium  citrinum 

0.059 

0.30 

2.36 

11.8 

3  to  9  is  observed.  The  only  obvious  explanation  is  that  the  anionic  forms 
of  dehydroacetate  do  not  penetrate  well.  The  major  decrease  in  activity 
occurs  between  pH  5  and  7,  correlating  with  the  pK^'s  near  5.2, 


CHAPTER  4 

SULFHYDRYL  REAGENTS 


A  substance  which  can  react  with  sulfhydryl  groups  and  thus  alter  en- 
zymic,  metabolic,  or  functional  processes  is  generally  called  a  sulfhydryl 
reagent.  Such  substances  represent  a  very  important  group  of  inhibitors 
and  have  been  used  extensively  to  determine  if  enzymes  or  metabolic  reac- 
tions depend  in  any  way  on  intact  sulfhydryl  groups.  In  addition,  they  are 
often  used  to  estimate  the  number  and  reactivity  of  sulfhydryl  groups  on 
proteins,  or  to  histochemically  localize  the  sulfhydryl  groups  in  cells  or  tis- 
sues. The  next  few  chapters  will  be  concerned  with  sulfhydryl  reagents,  and 
in  this  chapter  we  shall  discuss  several  general  aspects  of  inhibition  result- 
ing from  modifications  of  sulfhydryl  groups  and  some  of  the  problems  en- 
countered in  work  with  these  substances.  This  is  one  phase  of  inhibition 
that  has  recently  received  considerable  attention,  and  several  reviews  cover- 
ing certain  aspects  of  the  problem  are  available.  The  articles  by  Boyer  (1959) 
and  Putnam  (1953),  and  the  books  "Glutathione"  (1954)  and  "Sulfur  in 
Proteins"  (1959),  are  particularly  recommended. 

The  terminology  to  be  adopted  attempts  to  follow  the  most  recent  usage. 
The  sulfhydryl  group  (=  mercapto  group)  will  for  brevity  be  designated 
as  an  SH  group.  Compounds  containing  SH  groups  will  be  designated  as 
thiols  (elsewhere  occasionally  called  sulfhydryl  compounds  or  mercaptans). 
A  sulfhydryl  reagent  wiU  be  termed  an  SH  reagent.  The  designation  as  a 
sulfhydryl  enzyme  has  often  been  meant  to  imply  that  the  catalytic  activity 
of  the  enzyme  is  dependent  on  SH  groups,  i.e.,  that  the  SH  groups  actually 
participate  in  the  enzyme  reaction.  As  Boyer  (1959)  has  pointed  out,  not 
a  single  enzyme  has  been  definitely  shown  to  involve  protein  SH  groups 
in  the  catalysis,  and  we  shall  see  that  the  inhibition  of  an  enzyme  by  an 
SH  reagent  does  not  prove  that  the  SH  groups  are  functional.  Hence,  a 
more  practical  definition  of  a  sulfhydryl  enzyme  at  the  present  time  is  an 
enzyme  that  shows  a  loss  of  activity  when  some  or  all  of  its  SH  groups 
are  modified. 


635 


636  4.    SULFHYDRYL    REAGENTS 

ROLE   OF   SH    GROUPS   IN    METABOLISM   AND    FUNCTION 

Cellular  components  containing  SH  groups  may  be  conveniently  grouped 
in  three  categories:  (1)  low  molecular  weight  thiols,  such  as  the  cofactors  li- 
poate,  coenzyme  A,  and  glutathione,  or  various  amino  acids  and  related 
compounds,  such  as  cysteine,  homocysteine,  2-thiolliistidine,  ergothioneine, 
and  thioglycolate,  (2)  nonenzyme  proteins,  probably  including  most  of  the 
cytoplasmic  proteins  (e.g.,  those  involved  in  movement,  such  as  actomyosin, 
ciliary  proteins,  and  proteins  of  the  mitotic  spindle),  plasma  membrane  pro- 
teins, and  structural  proteins,  and  (3)  enzymes  of  all  types  and  catalyzing 
every  variety  of  reaction.  Modification  of  or  reaction  with  any  of  these  SH 
groups  may  directly  or  indirectly  alter  cellular  metabolism  and  function. 
Even  reaction  with  nonenzyme  protein  SH  groups  may  disturb  metabolism, 
because  of  the  role  such  proteins  may  play  in  the  structural  organization  of 
the  metabolic  units  or  in  the  permeabilities  of  cells.  In  addition  to  the  free 
SH  groups,  many  proteins  and  enzymes  contain  disulfide  (S — S)  groups 
that,  in  the  case  of  enzymes,  are  probably  not  involved  directly  in  the  ca- 
talysis but  in  the  structural  stability.  These  disulfide  groups  can  under  cer- 
tain circumstances  be  reductively  cleaved  to  form  free  SH  groups,  with 
simultaneous  loosening  of  the  protein  structure,  or  can  perhaps  react  di- 
rectly with  certain  agents  to  form  mercaptides. 

In  the  early  days  of  interest  in  thiols,  it  was  believed  that  SH  reagents 
altered  metabolism,  and  were  sometimes  lethal,  as  a  result  of  reaction  with 
glutathione  or  other  low  molecular  weight  thiols,  but  it  was  soon  realized 
that  enzyme  SH  groups  are  a  much  more  important  site  of  attack.  Even 
today  the  role  that  such  small  thiols  play  in  metabolism  and  the  importance 
of  their  reaction  with  SH  reagents  are  not  well  understood,  except  in  the 
case  of  coenzyme  A  and  lipoate.  The  ubiquitous  glutathione  plays  at  pres- 
ent an  indeterminate  role  in  metabolism,  except  for  its  likely  participation 
in  the  reactions  of  phosphoglyceraldehyde  dehydrogenase,  glyoxalase,  mal- 
eate  isomerase,  maleylacetoacetate  isomerase,  formaldehyde  dehydrogenase, 
and  indolylpyruvate  tautomerase,  and  in  transpeptidation  and  folate  split- 
ting. Low  molecular  weight  thiols  have  also  been  supposed  to  regulate  me- 
tabolism by  redox  equilibria  with  enzyme  SH  groups,  maintaining  a  cer- 
tain fraction  of  these  in  the  reduced  or  active  state. 

The  SH  groups  of  enzymes  have  been  considered  to  bind  cofactors  or 
coenzymes  to  the  apoenzyme,  or  to  form  acyl  or  phosphoryl  complexes 
with  intermediates  derived  from  substrates,  or  to  function  directly  as  redox 
couples  in  electron  transfer,  but  there  is  little  evidence  for  any  of  these,  as 
likely  as  they  may  be.  The  SH  group  readily  donates  electron  pairs  and 
thus  is  one  of  the  most  reactive  enzyme  groups  with  regard  to  the  formation 
of  covalent  bonds,  so  it  would  not  be  surprising  if  covalent  intermediate 
complexes  occur.  Whatever  the  role  SH  groups  play  in  enzyme  catalysis, 


CHEMICAL    PROPERTIES    OF   SH    GROUPS  637 

their  modification  often  abolishes  activity  and,  since  metabolism  depends 
on  sulfhydryl  enzymes,  it  is  evident  that  most  important  metabolic  path- 
ways would  be  sensitive  to  SH  reagents.  In  addition,  coenzyme  A,  lipoate, 
and  glutathione  function  in  key  metabolic  positions.  Thus  glycolysis,  the 
tricarboxylate  cycle,  fatty  acid  oxidation,  photosynthesis,  phosphate  trans- 
fer, and  various  synthetic  pathways  are  inhibitable  by  SH  reagents.  Many 
effects  of  thiols  on  metabolism  have  been  observed  but  no  detailed  mechan- 
isms emerge.  Brain  respiration  and  glycolysis  in  vivo  proceed  at  only  a  frac- 
tion of  their  maximal  rates;  it  has  long  been  known  that  glutathione  stim- 
ulates aerobic  glycolysis  in  the  brain,  and  thus  it  has  been  implicated  in 
the  regulation  of  cerebral  metabolism.  Mcllwain  (1959)  reported  that  the 
aerobic  glycolysis  of  brain  slices  is  stimulated  by  glutathione,  cysteine, 
homocysteine,  2-mercaptoethanol,  and  other  thiols,  although  the  effects  on 
respiration  are  rather  slight.  However,  the  respiratory  stimulation  by  50  m.M 
KCl  is  depressed  by  these  thiols,  as  is  the  excess  respiration  in  the  presence 
of  dinitrophei|ol.  The  glycolytic  stimulation  is  accompanied  by  a  decrease 
in  creatine-P  and  a  rise  in  inorganic  phosphate,  these  changes  being  corre- 
lated with  the  metabolic  changes.  If  the  respiratory  augmentation  produced 
by  increased  functional  activity  were  mediated  through  glutathione  or 
similar  thiols,  there  would  have  to  be  a  fairly  large  change  in  their  concen- 
trations, or  in  the  ratios  of  the  oxidized  and  reduced  forms,  which  is  not 
observed.  The  metabolic  relations  are  clear  but  it  is  not  known  by  what 
mechanism  the  thiols  reduce  brain  creatine-P. 

Studies  of  the  effects  of  SH  reagents  on  cell  function  are  complicated  by 
the  fact  that  undoubtedly  some  of  the  proteins  of  the  functional  systems 
contain  SH  groups,  and  may  even  be  dependent  upon  them.  This  has  been 
investigated  principally  in  the  proteins  involved  in  motility;  for  example, 
the  polymerization  of  G-  to  F-actin,  the  interaction  of  actin  and  myosin, 
the  ATP-induced  contractions  of  glycerinated  flagella,  the  round-up  of  cul- 
tured fibroblasts,  the  formation  of  the  mitotic  apparatus,  and  many  other 
phenomena  appear  to  be  dependent  on  free  SH  groups.  Cell  excitability 
and  impulse  conduction,  based  on  ionic  fluxes  and  a  specific  membrane 
^ructure,  must  also  involve  SH  groups  in  the  membrane.  Thus  effects  of 
SH  reagents  on  cell  function  cannot  be  immediately  interpreted  in  terms 
of  an  enzymic  or  metabolic  site  of  action. 


CHEMICAL  PROPERTIES  OF  SH  GROUPS 

Only  a  few  characteristics  of  the  SH  group  that  are  particularly  important 
in  enzyme  inhibition  will  be  discussed.  A  brief  and  excellent  summary  of 
sulfur  chemistry  is  that  of  Calvin  (1954)  and  much  of  interest  may  be  found 
in  "Organic  Sulfur  Compounds"  edited  by  Kharasch  (1961),  as  well  as  in 
the  general  references  given  earlier  in  this  chapter. 


638  4.    SULFHYDRYL   REAGENTS 

The  reactions  of  most  SH  reagents  with  thiols  depend  on  the  pH  and  this 
undoubtedly  relates  to  the  ionizations  of  both  SH  reagent  and  the  SH  groups 
reacted.  In  most  cases,  as  with  the  mercaptide-forming  reagents  and  the  al- 
kylating agents,  the  rate  and  degree  of  reaction  increase  with  increasing 
pH,  and  it  is  likely  in  these  cases  that  the  ionized  mercapto  anion,  R — S", 
is  more  reactive  than  the  un-ionized  R — SH  form.  One  may  visualize  some 
of  these  reactions  as  a  competition  between  the  SH  reagent  and  a  proton 
for  the  R — S~  group.  The  reaction  with  a  heavy  metal  ion,  for  example, 
may  be  written  as: 

R— SH  +  Me+  ^  R— S— Me  +  H+ 

and  it  is  obvious  that  increase  of  the  pH  will  favor  the  formation  of  the 
mercaptide  complex,  or,  to  put  it  in  another  way,  that  the  reaction: 

R— S-  +  Me+  ^  R— S— Me 

will  proceed  more  readily.  On  the  other  hand,  reactions  with  double  bonds 
(as  with  maleate  or  quinones)  or  oxidations  to  the  disulfide  may  occur  more 
readily  when  the  SH  group  is  un-ionized.  In  any  event,  the  state  of  ioniza- 
tion of  the  SH  group  is  important  in  enzyme  inhibition  and  may  account 
partly  for  the  different  reactivities  of  protein  SH  groups.  The  ionization  of 
SH  groups  is  well  discussed  by  Edsall  and  Wyman  (1958),  mainly  on  the 
basis  of  work  by  Benesch  and  Benesch  (1955).  The  ionization  microcon- 
stants  for  cysteine  are  given  in  Table  1-14-4.  It  is  clear  that  the  pK^  of  an 
SH  group  is  markedly  dependent  on  the  electric  field  present,  that  is,  on 
the  vicinal  ionic  groups.  One  might  roughly  estimate  the  p^,/s  for  an  SH 
group  as  shown  in  the  following  tabulation: 


pK„ 


Near  a  +  charged  group  7.2-8.5 

No  electric  field  8.5-9.2 

Near  a  —  charged  group  9.2-10.2 


On  the  surface  of  a  protein  the  electric  field  will  be  the  resultant  of  all  the 
contributions  of  the  ionic  groups.  Enzyme  SH  groups  must  therefore  exhibit 
a  wide  range  of  ionizations  at  any  designated  pH,  and  in  most  cases  will 
exist  mainly  in  the  un-ionized  form  at  physiological  pH.  The  piii^  of  the  SH 
groups  on  aldolase  in  4  M  urea  is  around  8.66,  but  the  native  enzyme  has 
6  exposed  SH  groups  with  \)K„  values  near  10.5  and  buried  SH  groups  with 
an  apparent  p^„  of  11.5  (Donovan,  1964). 


CHEMICAL  PROPERTIES  OF  SH  GROUPS  639 

Many  of  the  atomic  and  bond  properties  will  be  found  in  the  tables  of 
Chapter  1-6.  The  bond  dipole  moments  are  fairly  high  (C — S  1.73  and  S — H 
0.68,  corresponding  to  fractional  atomic  charges  of  0.20  and  0.11,  respec- 
tively) and  the  bonds  with  sulfur  are  readily  polarized  (the  molar  refrac- 
tions are  C — S  4.43,  S — H  4.62,  and  S — S  7.41)  compared  with  most  other 
bonds  occurring  in  proteins.  Furthermore,  the  bond  energies  are  uniformly 
low  compared  with  the  corresponding  oxygen  bonds,  except  for  the  disulfide 
bond,  which  is  a  good  deal  stronger  than  the  peroxide  bond  (see  accompa- 
nying tabulation).  These  fundamental  properties  account  for  many  of  the 


Bond  energies  (kcal/mole) 

C— S    54 

C— 0     80 

S— H  87 

0— H  105 

S— S    66 

0—0     34 

characteristic  reactions  of  the  SH  group  and  the  relatively  unique  role  of 
these  groups  in  enzyme  activity  and  inhibitition.  The  inherent  dipole  mo- 
ments and  the  high  polarizability  of  sulfur  bonds  may  play  an  important 
role  in  the  interactions  of  enzymes  with  substrates  and  inhibitors,  whereas 
the  bond  energies  are  involved  in  determining  ionization  tendencies,  oxida- 
tion-reduction potentials,  and  the  equilibria  between  SH  groups  and  disul- 
fide structures. 

The  thiol-disulfide  equilibria  are  important  for  enzyme  structure  in  all 
probability  but,  in  addition,  may  well  be  determining  factors  in  the  states 
and  reactivities  of  the  SH  groups.  It  has  been  shown  recently  that  the  reac- 
tion of  a  thiol  with  a  disulfide  is  not  a  simple  oxidation-reduction  but  an 
exchange  reaction  involving  a  two  step  ionic  displacement  (Eldjarn  and 
Pihl,  1957  a,  b;  Parker  and  Kharasch,  1959;  Foss,  1961),  often  with  the 
formation  of  mixed  disulfides: 

X— SH  +  Y— S— S— Y  ±^  X— S— S— Y  +  Y— SH 
X— S— S— Y  -f  X— SH  ±?  X— S— S— X  +  Y— SH 

Low  molecular  weight  thiols,  such  as  glutathione  or  cysteine,  could  thus 
interact  with  enzyme  SH  and  disulfide  groups  to  form  mixed  disulfides.  Par- 
ticularly in  the  cell,  where  such  thiols  occur,  these  interactions  may  be  im- 
portant in  regulating  enzyme  activity,  and  could  easily  affect  the  reactivity 
of  enzymes  with  SH  reagents.  That  this  can  actually  occur  with  proteins 
was  shown  by  the  use  of  a  colored  disulfide,  with  which  seralbumin  and  /?- 
lactoglobulin  react  to  form  mixed  disulfides  (Klotz  et  at.,  1958).  If  an  SH 
enzyme  and  oxidized  glutathione  (GSSG)  are  allowed  to  react,  one  would 


640  4.    SULFHYDRYL    REAGENTS 

have  ESH,  ESSG,  ESSE,  GSSG,  and  GSH  (where  E  represents  the  enzyme) 
present,  perhaps  only  ESH  being  catalytically  active  and  the  forms  ESSG 
and  ESSE  protected  from  SH  reagents.  Although  it  is  usually  thought  that 
only  the  SH  form  can  react  with  most  SH  reagents,  it  is  possible  that  di- 
sulfides are  occasionally  reactive;  aryl  arsinites,  for  example,  can  exert  a 
nucleophilic  displacement  on  the  disulfide  bond: 

OH 
RSSR  +  R'AsO(OH)-  ±^  RSAsR'  +  RS~ 

O 

However,  such  reactions  are  probably  slower  than  with  free  SH  groups. 
The  new  reagent,  dithiothreitol  (HS— CH2— CHOH— CHOH— CHg— SH), 
which  has  a  low  redox  potential  (  —  0.33  v  at  pH  7),  is  highly  water-soluble 
and  reduces  protein  disulfide  groups  (Cleland,  1964).  It  was  suggested  that 
it  might  be  valuable  in  protecting  enzyme  SH  groups,  having  several  ad- 
vantages over  the  ones  commonly  used,  and  could  also  be  applied  for  the 
purpose  of  maintaining  enzymes  in  the  SH  state  for  inhibition  studies. 

Hydrogen  bonding  by  sulfur  should  be  mentioned  since  it  must  play  a 
role  in  both  intra-  and  intermolecular  interactions  of  the  SH  group,  but 
most  of  the  data  we  have  derives  from  studies  of  the  small  thiols  and  there 
is  very  little  information  on  hydrogen  bonding  of  protein  SH  groups.  Boyer 
(1959)  has  presented  the  evidence  for  the  occurrence  of  hydrogen  bonds  to 
sulfur  in  a  variety  of  compounds.  Sulfur  does  not  form  hydrogen  bonds  as 
readily  as  oxygen  or  nitrogen,  since  it  is  less  electronegative  (as  indicated 
by  the  smaller  dipole  moment  of  the  S — H  bond  compared  to  the  0 — H  and 
N — H  bonds)  (Table  1-6-1).  However,  there  is  evidence  for  intramolecular 
S — H  •  •  •  0  and  S — H  •  •  •  N  bonds  in  cysteine  and  its  peptides,  and  Benesch 
et  al.  (1954)  have  advanced  hydrogen  bonding  to  explain  some  of  the  dif- 
ferent reactivities  of  simple  thiols.  It  is  possible  that  hydrogen  bonding  of 
enzyme  SH  groups  can  modify  their  susceptibilities  to  various  SH  reagents. 

Detection    and    Determination   of  Enzyme   SH    Groups 

Valuable  reviews  of  the  general  methods  for  the  determination  of  SH 
groups  in  proteins  and  enzymes  have  been  provided  by  HeUerman  and 
Chinard  (1955)  and  R.  Benesch  and  R.  E.  Benesch  (1962).  Some  of  the 
most  reliable  methods  involve  the  use  of  mercurials  (to  be  discussed  later, 
pages  762  and  766).  Here  we  shall  mention  only  a  few  of  the  more  recently 
developed  reagents  which  may  be  applicable  in  inhibition  studies. 

Bis(2?-nitrophenyl)disulfide  reacts  with  thiols  at  pH  8  to  form  1  mole  of 
p-nitrophenol  per  mole  of  thiol,  and  this  anion,  being  highly  colored,  can 
be  used  to  determine  the  thiol  concentration  (Ellman,  1959).  However,  this 


CHEMICAL  PROPERTIES  OF  SH  GROUPS  641 

reagent  is  poorly  soluble  in  water,  so  carboxyl  groups  were  introduced  to 
solubilize  it;  this  compound  is  5,5'-ditliiobis(2-nitrobenzoate)  and  reacts 


0,N {'         V — S— S — V  ) — NO; 


5,  5'-Dithiobis  (2-nitrobenzoate) 

with  thiols  and  SH  groups  in  the  blood  and  tissues.  Another  water-soluble 
reagent  for  free  SH  groups  is  2,2'-dicarboxy-4,4'-diiodoaminoazobenzene, 
which  was  shown  to  react  only  with  the  SH  groups  on  denatured  meromyosin 
(Fasold  et  al.,  1964).  The  number  of  SH  groups  reacted  can  be  determined 

coo' 

I— CHi— CONH-Y^      \^N=N^/       ^V-NHCO— CH2— I 

"ooc 

2,  2'-Dicarboxy-4,  4'-diiodoaminoazobenzene 

spectrophotometrically  because  of  the  chromogenic  azo  link.  A  yellow  SH 
reagent,  A^-(4-dimethylamino-3,5-dinitrophenyl)maleimide,  was  introduced 
by  Witter  and  Tuppy  (1960)  and  found  to  react  with  the  free  SH  groups 
of  seralbumin.  The  treated  protein  could  be  hydrolyzed  with  pepsin  and 
the  A'-(4-dimethylamino-3,5-dinitrophenyl)succinimido-cysteine  peptides 
isolated  by  means  of  their  yeUow  color.  This  reagent  was  used  by  Gold  and 
Segal  (1964)  to  obtain  information  on  the  nature  of  the  active  site  of  3- 
phosphoglyceraldehyde  dehydrogenase.  Following  pepsin  treatment  the  sin- 
gle hexapeptide  -  Ala-Ser-(DDPS-Cys)-Thr-Thr-AspNH2  -  w^as  found  to 
contain  essentially  aU  the  color.  This  provides  evidence  that  the  three  active 
sites  on  the  enzyme  are  similar  in  structure,  at  least  in  part,  and  that  the 
reactive  SH  groups  are  not  those  of  glutathione,  which  occurs  on  the  enzyme. 

O.N 


0,N 


A/-(4-Dimethylamino-3,  5- 
dinitrophenyl)  maleimide 


Such  reagents  would  not  be  particularly  useful  for  the  inhibition  of  SH  en- 
zymes because  of  their  bulky  structure  and  the  presence  of  a  variety  of 


642  4.    SULFHYDRYL   REAGENTS 

groups,  but  they  could  be  applied  to  the  determination  of  changes  in  the 
SH  content  after  treatment  of  the  enzymes  with  the  usual  inhibiting  re- 
agents. 

TYPES  OF  SH   REACTION    IMPORTANT  IN    INHIBITION 

The  reactions  of  most  of  the  important  SH  reagents  have  usually  been 
classified  into  four  types.  The  SH  groups  have  been  written  as  un-ionized 
in  all  cases,  not  implying  that  this  is  necessarily  the  only  reactive  form. 
The  mechanisms  of  these  reactions  will  be  discussed  in  the  chapters  devot- 
ed to  the  individual  inhibitors. 

(I)  Oxidation  of  SH  growps 

2R-SH         ^         X     •*  >     R-S-S-^R        ^        XH, 

R  =  (SH),        +        X     •*  »  R^l  +        XH. 

S 

(X  may  be  an  acceptor  of  either  hydrogen  atoms  or  electrons.) 

Examples:  o-iodosobenzoate,  porphyrindin,  porphyrexide,  iodine,  alloxan 
(not  the  only  mechanism),  ferricyanide,  oxidized  glutathione,  tetrathionate, 
sulfite,  performic  acid,  and  oxygen  (catalyzed  by  metal  ions). 

(II)  Mercaptide  formation 

R-SH  *  X"    ^ jT  R-S-X  -  H 

2R— SH         +        X^  •  -^ ^  R-S— X-S— R         ■         2H 

. /S^ 

R=(SH),        +        X  '    "*  >:         R^        X  +        2H 

Examples:  HgClg,  organic  mercurials,  arsenite,  organic  arsenicals,  and 
various  heavy  metal  ions  (Cu++,  Pb++,  Cd++,  Ag+,  etc.). 

(III)  Alkylation  of  SH  groups  [alkyl  transfer) 

R— SH  +  X— R'  ±?  R— S— R'  +  X 

Examples:  iodoacetate,  iodoacetamide,  S-  and  N-mustards,  chloraceto- 
phenone,  chloropicrin,  bromobenzylcyanide,  and  fluoropjTuvate. 


FACTORS  •DETERMINING    THE    REACTIVITIES    OF    SH    GROUPS  643 

(IV)  Addition  of  SH  groups  to  double  bonds 

CH— R'       R— S— CH— R' 

R— SH  +11  ±?  I 

CH— R"  CH2— R" 

OH 

R— SH  +  0  =  C— R'  ±^  R— S— C— R' 

(This  may  also  be  considered  as  a  type  of  alkylation  reaction.) 

Examples:  maleate,  iV-ethylmaleimide,  quinones,  acrolein,  acetoacetate, 
and  methylglyoxal. 

FACTORS  DETERMINING  THE  REACTIVITIES  OF  SH  GROUPS 

The  SH  groups  of  various  simple  thiols,  peptides,  and  proteins  differ  mark- 
edly in  reactivity  with  SH  reagents.  Although  this  has  been  known  for  many 
years,  the  molecular  basis  for  this  differential  reactivity  is  poorly  understood. 
In  general  the  reactivity  is  maximal  in  simple  thiols  and  minimal  in  proteins, 
but  in  proteins  there  are  usually  reactive  and  unreactive  SH  groups.  Barron 
(1951)  classified  SH  groups  as  (a)  freely  reacting,  (b)  sluggish,  and  (c)  mask- 
ed, depending  on  whether  they  react  readily,  slowly,  or  not  at  all.  Although 
such  a  division  is  often  useful  in  discussing  SH  groups,  there  is  actually  a 
continuous  sequence  of  groups  from  highly  reactive  to  unreactive.  If  a  pro- 
tein is  allowed  to  react  with  an  SH  reagent  under  approximately  physiolo- 
gical conditions,  one  generally  finds  that  the  SH  groups  disappear  at  differ- 
ent rates,  perhaps  several  reacting  completely  before  others  are  affected.  A 
graded  response  is  clearly  seen  in  the  reaction  of  aldolase  with  p-chloromer- 
curibenzoate  (Swenson  and  Boyer,  1957).  Ten  SH  groups  react  relatively 
rapidly,  a  few  more  slowly,  and  the  rest  not  at  all  unless  the  enzyme  is  un- 
folded by  high  concentrations  of  urea.  Furthermore,  the  reaction  of  the  first 
10  SH  groups  does  not  alter  the  enzyme  activity,  indicating  that  these 
groups  are  not  part  of,  or  even  too  near,  the  active  center,  whereas  disap- 
pearance of  the  more  slowly  reacting  groups  abolishes  the  activity.  A  similar 
situation  has  been  observed  with  urease,  which  has  5  cysteine  residues  per 
molecule,  one  freely  reacting  and  4  sluggish,  the  catalytic  activity  being 
affected  only  by  modification  of  the  latter  (Hellerman,  1939;  Hellerman  et 
al.,  1943).  Thus  porphyrindin,  iodoacetamide,  and  iodosobenzoate  react  with 
one  SH  group  but  do  not  inhibit  (except  at  very  high  concentrations), 
whereas  p-chloromercuribenzoate  can  combine  with  another  SH  group  abol- 
ishing the  activity.  These  examples  —  and  we  shall  have  occasion  to  discuss 
many  others  —  illustrate  four  most  important  principles:  (1)  the  differential 
reactivity  of  enzyme  SH  groups,  (2)  the  increase  in  reactivity  of  many  of 
the  SH  groups  following  denaturation,  (3)  the  different  reactivities  of  var- 


644  4.    SULFHYDRYL   REAGENTS  • 

ious  SH  reagents,  and  (4)  the  lack  of  correlation  between  the  reactivity  of 
SH  groups  and  their  relationship  functionally  or  spatially  to  the  active 
center.  These  principles  are  central  to  the  problem  of  inhibition  by  SH  re- 
agents and  require  some  general  discussion  relative  to  the  possible  mecha- 
nisms involved. 

The  various,  and  mostly  obvious,  hypotheses  to  explain  the  differential 
reactivity  of  SH  groups  have  frequently  been  presented  with  a  prolixity 
inversely  proportional  to  the  amount  of  evidence  available.  Indeed,  at  the 
present  time  there  is  little,  if  any,  positive  evidence  for  any  explanation, 
but  there  are  a  number  of  factors  that  must  be  of  some  importance,  and 
these  can  be  enumerated.  It  should  be  emphasized  that  differential  reactivity 
should  be  based  on  accurate  spectrophotometric  or  argentimetric  titrations 
of  the  enzyme  SH  groups  under  various  conditions.  The  fundamental  prob- 
lem is  to  determine  the  cause  for  the  slow  reactions  of  all  or  a  fraction  of 
an  enzyme's  SH  groups,  the  total  number  of  such  groups  being  determined 
by  quantitative  titrations  of  the  enzyme  after  complete  unfolding.  The 
theories  assume  either  that  (I)  free  SH  groups  are  present  in  the  native  cat- 
alytically  active  enzyme,  but  are  for  some  reason  unable  to  react  readily 
with  SH  reagents,  or  that  (II)  the  unreactive  SH  groups  are  so  modified 
that  they  are  no  longer  free. 

(I)  Free  SH  groups  present  in  native  enzyme 

A.  Steric  factors  impede  reaction:  the  reagent  is  simply  unable  to  approach 
the  SH  group  because  it  is  located  in  a  pit  or  crevice  of  the  enzyme,  or  ac- 
tually within  the  protein  structure. 

B.  Electrostatic  factors  impede  reaction:  the  SH  group  is  in  the  electric 
field  of  surrounding  groups,  this  discouraging  reactions  with  reagents  of 
the  same  charge  sign  as  these  groups. 

C.  Ionization  state  impedes  reaction:  if  either  the  SH  or  the  S~  form  reacts 
preferentially  but  is  not  significantly  present  at  the  experimental  pH,  the 
reaction  will  be  appreciably  slowed. 

(II)  SH  groups  are  not  free  in  the  native  enzyme 

A.  Present  as  disulfide  groups 

B.  Hydrogen  bonded  to  adjacent  groups 

C.  Present  in  thiazolidine  or  thiazoline  rings 

D.  Reacted  with  some  component  of  enzyme  reaction:  for  example,  acylated, 
phosphorylated,  or  complexed  with  some  metal  ion. 

The  fact  that  complete  opening  up  or  unfolding  of  the  protein  structure 
invariably  increases  the  susceptibility  of  certain  SH  groups  to  attack  does 
not  necessarily  imply  that  the  groups  are  in  some  way  within  the  enzyme, 


FACTORS    DETERMINING    THE    REACTIVITIES    OF    SH    GROUPS  645 

since  denaturation  can  also  dissolve  disulfide  bonds,  hydrogen  bonds,  ring 
structures,  and  other  possible  chemical  interactions  of  the  SH  groups.  As 
long  as  one  studies  only  a  single  SH  reagent,  it  is  easy  to  postulate  a  reason- 
able mechanism  for  the  unreactivity  of  particular  SH  groups.  For  example, 
if  one  finds  that  iodoacetate  does  not  alkylate  an  enzyme  SH  group,  the 
group  may  be  thought  of  as  sequestered  within  the  protein  structure,  but 
if  subsequent  work  shows  that  p-chloromercuribenzoate  reacts  readily  with 
this  group,  this  hypothesis  must  be  abandoned  inasmuch  as  j^-chloromer- 
curibenzoate  is  a  larger  molecule  than  iodoacetate.  Likewise,  postulating 
that  negative  charges  surround  the  SH  group,  preventing  the  approach  of 
iodoacetate,  will  not  be  valid  if  p-chloromercuribenzoate  is  effective,  since 
both  of  these  reagents  are  negatively  charged,  as  has  been  pointed  out  by 
Boyer  (1959)  in  perhaps  the  best  discussion  of  differential  SH  group  reac- 
tivity. It  is  certainly  likely  that  steric  and  electrostatic  factors  are  occa- 
sionally important,  but  one  must  demonstrate  some  correlation  of  the  un- 
reactivity of  the  SH  groups  with  the  properties  of  a  variety  of  SH  reagents. 
Haurowitz  and  Tekman  (1947)  believed  that  protein  SH  groups  are  often 
inaccessible  to  reagents  because  of  the  tightly  folded  nature  of  the  polypep- 
tide chains,  rather  than  chemically  combined,  because  unfolding  is  accom- 
panied by  the  appearance  of  reactivity  in  other  than  SH  groups,  e.g.  phe- 
nolic groups.  Although  this  is  suggestive,  it  is  not  proof  for  the  inaccessibi- 
lity theory.  Unreactivity  due  to  an  unfavorable  ionization  state  has  per- 
haps been  insufficiently  considered,  particularly  for  inhibitors  such  as  iodo- 
acetate, and  there  is  no  question  but  that  the  very  low  concentration  of  S~ 
near  neutrality  for  some  SH  groups  must  be  important. 

Turning  to  the  second  group  of  theories,  no  one  denies  that  disulfide 
groups  occur  in  some  enzymes,  but  that  this  generally  cannot  explain  the 
differential  reactivity  of  SH  groups  is  obvious.  Indeed,  one  finds  a  wide 
range  of  reactivities  in  simple  thiols  where  cryptic  exclusion  or  disulfide 
bonding  may  be  eliminated.  Benesch  et  al.  (1954)  not  only  demonstrated 
different  nitroprusside  reaction  rates  with  various  biologically  important 
thiols,  but  showed  that  urea  increases  the  reactivity  of  the  more  slowly 
reacting  SH  groups,  just  as  it  does  in  proteins.  This  was  interpreted  in 
terms  of  the  breaking  of  hydrogen  bonds  and  thus  the  initial  sluggishness 
of  reaction  as  due  to  hydrogen  bonding  of  the  SH  groups  to  adjacent  amino 
or  peptide  groups.  This  may  occur  in  a  cysteine  peptide  in  the  following 
way: 

H,C  "H  H,C  H 

^11  "I  i 

R,— CONH-HC— CO— N— R,  R— CONH— HC— CO— NH— R, 

the  hydrogen  donator  depending  on  the  pH.  We  have  seen  that  SH  groups 
form  only  weak  hydrogen  bonds  (page  640),  so  that  it  is  likely  that  this 


646  4.    SULFHYDRYL   REAGENTS 

alone  cannot  depress  the  reactivity  too  greatly.  However,  in  addition  the 
hydrogen  bond  may  bring  an  adjacent  side  chain  into  the  region  of  the  SH 
group  and  this  second  steric  factor  may  further  reduce  the  reactivity.  Ben- 
esch  and  Benesch  (1953)  compared  the  peptides,  phenacetyl-L-cysteinyl- 
glycine  (PCG)  and  phenacetyl-L-cysteinyl-D-valine  (PCV),  with  respect  to 
the  polarographic  reduction  of  their  mercaptides  with  mersalyl  and  Ag+, 
and  found  that  a  relative  suppression  of  the  ability  of  the  PCV  SH  group 
to  react  with  these  reagents  is  evident.  This  was  interpreted  as  due  to  the 
steric  interference  of  the  isopropyl  group  of  valine  in  PCV,  brought  into 
the  proximity  of  the  SH  group  by  hydrogen  bonding,  whereas  in  PCG  there 
is  no  such  side  chain.  In  proteins  the  interference  may  be  even  greater. 
This  concept  thus  involves  both  a  reaction  of  the  SH  group  and  steric 
factors. 

The  hydrogen-bonded  structures  could  further  lose  water  to  form  more 
stable  thiazolidine  or  thiazoline  rings.  Linderstrom-Lang  and  Jacobsen 
(1941)  found  that  2-methylthiazoline  can  hydrolyze  under  certain  con- 
ditions to  release  an  SH  group  and  a  peptide  linkage,  such  as  occurs 
during  protein  denaturation.  Thus  in  a  cysteine  peptide,  where  the  hy- 
drogen bonding  is  now  to  the  keto  oxygen,  the  following  structures  can 
be  written: 

S— H 

/        \ 
H,C  O  HX S 


Ri — HC— NH— C— R,  R— HC— NH— C  — Rs 

OH 
H-bonded  form  Thiazolidine  form 


H,C S 

"I  I 

R  — HC— N=C-R2 

Thiazoline  form 

Such  transformations  have  been  more  recently  discussed  by  Calvin  (1954) 
in  connection  with  the  structure  of  glutathione  and  the  reactivities  of  SH 
groups  in  enzymes.  Indirect  evidence  for  these  rings  was  obtained  by  com- 
paring the  absorption  spectra  of  thiols  with  2-methylthiazoline  (although 
at  high  acidity  so  that  the  situation  near  neutrality  is  still  not  clear).  If 
such  structures  occur  in  enzymes,  they  could  unquestionably  account  for 
unreactivity. 


INTERPRETATION    OF    INHIBITIONS    BY    SH    REAGENTS  647 

It  is  quite  possible  that  many  or  all  of  these  mechanisms  contribute  in 
various  situations  to  the  differential  reactivity  of  SH  groups,  and  that  we 
should  not  be  too  eager  to  argue  for  a  single  dominant  factor.  After  all, 
there  is  a  graded  scale  of  reactivity,  which  in  itself  implies  multiple  mecha- 
nisms. It  would  probably  aid  in  the  characterization  of  SH  groups  if  some 
standard  method  for  designating  the  reactivity  could  be  used,  rather  than 
designating  them  by  terms  such  as  "sluggish,"  etc.  The  time  for  50%  reac- 
tion, where  determinable,  might  be  the  simplest  and  most  useful,  although 
some  form  of  reaction  rate  constant  would  be  preferable. 


INTERPRETATION  OF  INHIBITIONS  BY  SH  REAGENTS 

The  SH  reagents  are  used  most  commonly  to  determine  whether  a  parti- 
cular enzyme  is  an  "SH  enzyme"  or  not.  What  this  means  depends  on  one's 
definition  of  "SH  enzyme."  If  we  take  the  definition  proposed  earlier  (page 
635)  that  an  SH  enzyme  is  one  that  is  inhibited  by  SH  reagents,  not  a  great 
deal  has  been  achieved  by  proving  that  an  enzyme  belongs  to  this  class. 
In  the  past,  many  workers  have  been  satisfied  to  stop  at  this  distinction, 
and  the  designation  of  an  enzyme  as  an  SH  enzyme  has  been  deemed  suffi- 
cient without  further  discussion  as  to  the  significance  of  the  observation. 
On  the  other  hand,  some  have  assumed  immediately  that  inhibition  by  SH 
reagents  indicates  a  catalytically  functional  role  for  an  SH  group  at  the 
active  center,  and  this,  as  we  have  seen,  is  entirely  unjustifiable.  It  is  thus 
important  to  determine  as  far  as  possible  what  such  inhibition  means  and 
what  valid  conclusions  may  be  drawn  from  the  use  of  SH  reagents. 

The  various  mechanisms  by  which  SH  reagents  can  inhibit  pure  enzymes 
may  be  classified  in  the  following  way. 

(A)  The  SH  group  reacted  is  at  the  active  center  and  is  functional.  The  SH 
group  may  be  involved  in  the  binding  of  substrate,  coenzyme,  or  activator 
to  the  apoenzyme,  or  it  may  participate  in  the  transfer  of  groups  of  elec- 
trons. 

(B)  The  SH  group  reacted  is  at  the  active  center  but  is  nonfunctional.  It  is 
possible  that  an  SH  group  occurs  at  the  active  center  but  is  unrelated  to 
the  catalytic  process. 

(C)  The  SH  group  reacted  is  vicinal  to  the  active  center.  The  SH  reagent 
introduces  a  new  structure  on  the  enzyme,  and  if  this  is  near  enough  to 
the  active  center  it  may  either  sterically  or  electrostatically  modify  the 
reactions  proceeding  at  the  active  center. 

(D)  Reaction  of  the  SH  groups  alters  the  enzyme  protein  structure.  This 
could  also  apply  to  reaction  with  disulfide  groups  or  other  complexes  form- 
ed by  SH  groups.  The  change  in  protein  structure  would  then  reduce  the 


648  4.    SULFHYDRYL    REAGENTS 

rate  of  the  enzyme  reaction,  particularly  if  it  included  the  region  of  the 
active  center. 

(E)  Tlie  SH  group  reacted  is  on  the  substrate.  This  is  a  possibility  especially 
in  the  case  of  proteolytic  enzymes,  the  modification  of  the  peptide  or  protein 
substrate  preventing  normal  reaction  with  the  enzyme. 

(F)  The  SH  reagent  interferes  in  a  manner  unrelated  to  SH  groups.  Many 
SH  reagents  are  not  entirely  specific  for  SH  groups;  e.g.,  iodoacetate  also 
reacts  with  amino  groups  and  with  heavy  metal  ions  such  as  Cu++,  and  can 
often  form  complexes  with  protein  groups  other  than  SH,  especially  amino 
and  carboxylate  groups.  Also  the  SH  reagent  may  inhibit  because  it  is 
structurally  similar  to  the  substrate  and  can  compete  with  it  for  the  active 
site;  e.g.,  p-chloromercuribenzoate  may  act  like  a  substituted  benzoate  on 
certain  enzymes  rather  than  as  a  mercurial. 

Other  mechanisms  can  be  visualized  in  special  cases  and  particularly  for 
those  enzymes  comprising  several  units  and  catalyzing  complex  reactions, 
since  the  SH  reagents  can  conceivably  dissociate  the  functionally  related 
units,  just  as  /j-chloromercuribenzoate  can  split  the  relatively  simple  muscle 
phosphorylase  a  into  four  equivalent  fractions  (Madsen  and  Cori,  1956). 

It  is  very  difficult  to  distinguish  between  the  first  four  possibilities.  In- 
deed, proof  of  the  functional  role  of  SH  groups  usually  must  come  from 
evidence  other  than  inhibition.  Protection  of  the  enzyme  against  SH  re- 
agents by  the  substrate  does  not  provide  adequate  evidence  that  the  react- 
ed SH  group  is  part  of  the  active  center,  since  the  substrate  could  also  slow 
down  or  prevent  reaction  with  vicinal  groups  as  well,  and  could  also  sta- 
bilize the  protein  structure  around  the  active  center.  The  secondary  alter- 
ation of  protein  structure  brought  about  by  reaction  of  SH  groups  cannot 
always  be  detected  by  reversal  experiments  because  the  changes  may,  like 
certain  types  of  denaturation,  be  reversible.  One  must  therefore  conclude 
that  the  demonstration  of  inhibition  by  SH  reagents  indicates  at  best  (as- 
suming the  mechanisms  (E)  and  (F)  have  been  eliminated)  only  that  one 
or  more  SH  groups  are  sufficiently  near  the  active  center  to  interfere  with 
the  catalysis,  either  directly  or  by  structural  changes,  when  they  are  react- 
ed. It  must  be  admitted  that  such  a  conclusion  is  not  very  informative, 
especially  when  it  is  considered  that  most  enzymes  contain  5  to  30  SH 
groups  per  molecule  and  that  statistically  one  would  expect  one  or  more 
of  these  to  be  near  the  active  center.  Indeed,  it  is  rather  surprising  that  in 
some  instances  a  fair  number  of  SH  groups  can  be  reacted  without  altering 
the  catalytic  activity. 

One  characteristic  of  inhibition  by  SH  reagents  which  has  been  often 
neglected  is  that  the  reaction  with  the  enzyme  SH  groups  in  most  cases 
introduces  a  new  side  chain  onto  the  protein.  The  inhibition  may  be  as 
much  related,  if  not  more,  to  the  properties  of  this  side  chain  as  to  the 


INTERPRETATION    OF    INHIBITIONS    BY    SH    REAGENTS  649 

disappearance  of  a  free  SH  group.  These  new  groupings  have  varying  sizes 
and  frequently  electrical  fields.  It  is  quite  possible  that,  all  else  being  equal, 
a  smaller  reagent  of  a  particular  type  may  exert  less  inhibition,  due  simply 
to  the  fact  that  it  exerts  less  steric  hindrance  to  the  catalytic  process.  Thus 
the  inhibition  by  methylmercuric  chloride  may  be  different  from  that  by 
p-chloromercuribenzoate  for  this  reason.  The  groups  introduced  by  iodo- 
acetate  and  p-chloromercuribenzoate  are  negatively  charged,  whereas  those 
from  iodoacetamide  and  phenylmercuric  acetate  are  uncharged,  and  this 
could  well  be  responsible  for  some  of  the  differences  observed  between  these 
inhibitors.  This  is  one  reason  why  many  studies  with  SH  reagents  would 
profit  from  a  quantitative  comparison  of  the  effects  of  a  large  number  of 
inhibitors  of  different  types. 

Another  factor  of  some  importance  may  be  the  influence  of  the  reaction 
of  certain  SH  groups  on  the  reactions  of  other  SH  groups.  Further  reaction 
apparently  may  be  either  depressed  or  accelerated.  In  phosphorylase  a  the 
reaction  of  the  first  SH  group  seems  to  facilitate  the  combination  of  the 
remaining  18  groups  with  p-chloromercuribenzoate  (Madsen  and  Gurd,  1956). 
On  the  other  hand,  reaction  of  one  SH  group  on  hemoglobin  prevents  the 
further  reaction  of  one  or  two  other  groups  with  Ag+,  implying  that  the  SH 
groups  here  occur  in  clusters  (Ingram,  1955).  The  number  of  molecules  of 
SH  reagent  bound  to  the  protein  may  thus  not  be  equivalent  to  the  number 
of  SH  groups. 

Boyer  (1959)  has  emphasized  that  insufficient  consideration  has  usually 
been  given  to  the  possible  secondary  structural  changes  induced  in  enzymes 
by  reaction  with  SH  reagents.  If  some  SH  groups  are  unreactive  because 
of  steric  blocking  or  chemical  combination,  these  hindrances  must  be  over- 
come in  order  to  react  these  groups,  and  this  could  imply  a  modification 
in  the  protein  structure  that  in  itself  might  be  inhibitory.  It  has  been  ob- 
served frequently  that  the  poorly  reacting  SH  groups  are  more  important 
in  the  enzyme  structure  than  the  free  readily  reactive  ones.  One  of  the  first 
statements  of  the  importance  of  structure  in  the  inhibition  by  SH  reagents 
resulted  from  work  on  urease  by  Desnuelle  and  Rovery  (1949).  Phenyliso- 
cyanate  reacts  with  certain  SH  groups  rapidly  but  this  does  not  inhibit; 
inhibition  begins  when  the  unreactive  SH  groups  are  attacked,  and  this 
was  attributed  to  a  reversible  change  in  the  enzyme  structure.  Similarly, 
various  properties  of  aldolase  change  as  the  SH  groups  are  progressively 
reacted  with  p-chloromercuribenzoate:  after  3-5  are  reacted,  the  enzyme 
begins  to  be  more  readily  attacked  by  trypsin;  after  10  are  reacted,  the  tur- 
bidity increases,  denoting  marked  structural  changes,  and  inhibition  is  ob- 
served (Szabolcsi  and  Biszku,  1961).  There  is  a  progressive  labilization  of 
the  tertiary  structure,  accompanied  by  appearance  of  previously  masked 
SH  groups,  with  further  reaction  and  eventual  denaturation.  In  many  cases, 
the  initial  reactions  must  reduce  the  protein  stability,  perhaps  only  locally. 


650  4.    SULFHYDKYL    REAGENTS 

and  this  spreads  and  progresses  rapidly  as  further  groups  are  attacked,  just 
as  in  other  types  of  denaturation.  The  blocking  of  the  SH  groups  of  phos- 
phoglyceraldehyde  dehydrogenase  changes  the  optical  rotation  and  the  in- 
trinsic viscosity,  the  latter  increasing  linearly  with  the  equivalents  of  p- 
chloromercuribenzoate  reacted  (Elodi,  1960).  Reversibility  with  cysteine 
varies  with  the  time  of  exposure  to  the  mercurial,  at  first  the  effects  being 
completely  reversible  and  eventually  irreversible,  again  indicating  a  pro- 
gressive breakdown  of  the  protein  structure.  Elodi  postulated  three  phases: 
(1)  a  reversible  reaction  with  certain  SH  groups,  (2)  an  unfolding  of  the 
polypeptide  helices  as  a  result  of  the  alteration  of  the  SH  groups,  and  (3) 
precipitation  due  to  intermolecular  bridges  formed  between  the  new  groups 
appearing  on  the  protein  surface.  Ribonuclease  is  perhaps  another  example 
of  structural  changes  resulting  from  the  scission  of  disulfide  bonds,  of  which 
there  are  4  in  the  native  enzyme:  breaking  1  does  not  alter  the  activity, 
breaking  2  inhibits  about  20%,  breaking  3  inhibits  about  40%,  and  then 
suddenly  the  activity  drops  to  zero  as  the  last  disulfide  is  split  (Resnick 
et  al.,  1959).  The  — S — S —  bonds  were  believed  to  be  of  importance  in  pro- 
viding stability  to  the  secondary  structure  of  the  enzyme,  their  breaking 
leading  to  progressive  unfolding. 


PROTECTION   AND   INHIBITION    REVERSAL  BY  THIOLS 

Some  of  the  problems  involved  in  protection  and  reversal  experiments 
with  SH  reagents  and  thiols  were  discussed  at  some  length  in  Volume  I 
(pages  622-626).  A  few  of  the  conclusions  reached  there  will  be  briefly  sum- 
marized. (1)  Protection  or  reversal  by  a  thiol  depends  on  the  relative  affin- 
ities of  the  SH  reagent  for  the  enzyme  SH  groups  and  the  thiol,  and  the 
relative  concentrations  of  the  components,  and  hence  every  degree  of  re- 
versibility of  SH-inhibited  enzymes  may  be  observed.  (2)  Protection  or 
reversal  by  a  thiol  does  not  provide  conclusive  information  on  the  mecha- 
nism of  the  inhibition  or  the  enzyme  groups  attacked.  (3)  Irreversibility  is 
brought  about  not  only  by  very  tightly  bound  reagents,  but  by  progressive 
structural  changes  in  the  enzyme,  as  discussed  above.  This  type  of  irrever- 
sibility thus  increases  with  the  concentration  of  the  SH  reagent  and  the 
time  of  exposure.  (4)  The  amount  of  useful  information  relative  to  the 
mechanism  of  inhibition  obtained  from  such  studies  is  much  less  then 
commonly  believed. 

The  stability  of  the  product  formed  by  reaction  of  an  enzyme  SH  group 
with  an  SH  reagent  varies  with  many  factors,  most  of  which  have  been 
mentioned  in  connection  with  the  differential  reactivity  of  these  groups. 
Sometimes  the  product  is  completely  stable  for  all  experimental  purposes 
and  the  reacted  enzyme  is  permanently  altered;  such  would  be  the  case 
with  most  of  the  alkylating  agents.  Then  the  mercaptide  complexes  vary 


GENERAL  CONSIDERATIONS  OF  THE  USES  OF  SH  REAGENTS    651 

greatly  in  stability,  so  that  in  some  instances  they  can  be  split  at  a  rate 
too  rapid  to  be  technically  measurable,  while  in  others  the  rate  is  too  slow 
to  measure.  It  is  not  necessary  that  this  stability  be  correlated  with  struc- 
tural changes  in  the  enzyme  or  irreversible  inactivation.  The  inability  to 
reactivate  an  enzyme  inhibited  by  an  SH  reagent  can  be  attributed  to  a 
variety  of  factors,  some  of  which  are  listed  below. 

(A)  The  binding  of  the  SH  reagent  to  the  enzyme  is  stronger  than  to  the 
reversor;  one  must  use  the  proper  reversor  and  concentration  (e.g.,  dimer- 
caprol  will  reverse  some  inhibitions  untouched  by  cysteine). 

(B)  The  enzyme  is  chemically  altered  by  the  SH  reagent  so  that  it  is  not 
a  question  of  a  tightness  of  binding;  reversal  can  occur  only  by  a  chemical 
transfer  of  the  attached  group  to  another  radical. 

(C)  The  enzyme  is  structurally  altered  irreversibly  (denatured)  by  the 
blocking  of  the  SH  groups. 

(D)  The  SH  reagent  has  caused  a  splitting  off  of  some  coenzyme  or  co- 
factor,  which  must  be  added  back  following  restoration  of  the  SH  group 
for  activity  to  be  evident. 

(E)  The  reversor  may  in  some  manner  inhibit  the  enzyme,  even  though 
restoring  the  free  SH  groups  initially,  as  when  the  SH  groups  are  oxidized 
by  disulfides  formed  from  the  oxidation  of  the  added  thiols. 

In  order  that  irreversibility  be  correctly  attributed  to  protein  denaturation, 
these  other  possibilities  must  be  ruled  out.  Complete  reversibility  is  more 
easily  interpretable,  and  one  can  quite  confidently  say  that  at  least  no  per- 
manent derangements  in  the  enzyme  structure  have  been  induced  by  the 
SH  reagent. 

Protection  experiments,  in  which  some  thiol  is  added  previous  to,  or  with, 
the  SH  reagent,  are,  as  has  been  emphasized  earlier,  of  little  value,  since  all 
one  is  doing  is  reducing  the  concentration  of  the  free  SH  reagent  (assuming 
that  it  reacts  with  the  thiol).  Actually,  it  is  a  little  difficult  to  speak  of  this 
as  protection,  inasmuch  as  one  usually  would  not  call  a  reduction  in  inhib- 
itor concentration  a  type  of  protection.  It  is  very  unlikely  that  worthwhile 
information  can  be  obtained  from  such  experiments,  aside  from  the  practical 
determination  of  the  ability  of  substances  to  reduce  the  toxic  effects  of  SH 
reagents. 


GENERAL  CONSIDERATIONS  OFTHE  USES  OF  SH  REAGENTS 

Numerous  types  of  reagent  are  available  for  satisfactorily  specific  reaction 
with  SH  groups  but  no  single  one  is  adequate  for  all  purposes.  The  most 
useful  information  on  the  nature  of  enzyme  SH  groups,  their  locations  and 


652  4.    SULFHYDRYL   REAGENTS 

relation  to  the  catalysis,  can  be  obtained  by  the  proper  use  of  several  types 
of  SH  reagent.  It  is  also  advisable  to  use  different  concentrations  of  the 
reagents  (it  is  not  very  informative  to  report  that  1  mM  of  some  SH  reagent 
inhibits  100%)  and  calculate  a  K,  that  characterizes  the  potency  of  the  in- 
hibition and  the  affinity  of  the  enzyme  for  the  reagent.  In  order  to  present 
the  kinetics  properly,  it  is  necessary  to  determine  if  the  inhibitions  under 
the  experimental  conditions  used  are  reversible,  and  for  this  purpose  it  is 
best  to  perform  the  reversal  study  in  nitrogen.  It  is  also  useful  to  determine 
the  degree  of  inhibition  and  the  number  and  type  of  SH  groups  reacted 
simultaneously  in  order  to  correlate  reactivity  and  relationship  to  the  active 
center.  Finally,  at  least  some  simple  rate  studies  should  be  done  to  deter- 
mine if  the  inhibitions  observed  are  for  equilibrium  conditions.  In  many 
reports  one  finds  only  that  the  enzyme  was  incubated  with  the  SH  reagent 
for  a  certain  period  (even  this  information  is  frequently  omitted)  and  it  is 
impossible  to  determine  if  the  inhibition  observed  is  maximal  or  not.  SH 
reagents  have  perhaps  been  used  during  the  past  several  years  more  com- 
monly than  any  other  type  of  enzyme  inhibitor  and  yet  they  have  been 
used  with  little  concern  for  the  many  complexities  involved  in  the  interpre- 
tation of  the  results,  with  a  few  notable  exceptions. 

Despite  the  generally  good  specificity  of  these  reagents  for  SH  groups, 
they  are  not  specific  inhibitors  from  the  metabolic  standpoint  in  most  cases. 
Since  SH  groups  are  present  not  only  in  many  enzymes  but  in  most  other 
proteins  of  the  cell,  one  must  expect  that  in  complex  systems  there  will  be 
many  components  reacted.  Whether  some  of  these  reactions  will  be  of  im- 
portance in  what  is  measured  will  depend  on  the  nature  of  the  work.  It  is 
probably  justifiable  to  suggest  that  the  use  of  most  SH  reagents  be  restrict- 
ed at  the  present  time  to  enzyme  studies  for  the  purpose  of  determining  the 
nature  of  the  enzyme  SH  groups.  As  the  complexity  of  the  system  increases, 
the  value  of  SH  reagents  diminishes,  at  least  if  one  is  trying  to  correlate 
some  enzymic  or  metabolic  process  with  over-all  cellular  metabolism  or 
function.  In  this  connection  it  is  interesting  to  note  that  although  the  most 
reactive  SH  reagents  are  generally  best  for  pure  enzyme  work,  this  is  not 
necessarily  true  for  more  complex  systems.  What  one  usually  requires  in 
metabolic  or  functional  investigations  is  specificity  with  respect  to  a  partic- 
ular enzyme  or  metabolic  pathway.  Thus  iodoacetate,  although  it  is  a  rather 
poor  reagent  for  the  detection  of  SH  groups  and  has  frequently  been  ma- 
ligned for  this  purpose,  is  actually  more  valuable  in  cellular  work  than  most 
of  the  others  since  it  has  the  ability,  if  used  properly,  of  inhibiting  the  phos- 
phoglyceraldehyde  dehydrogenase  and  glycolysis  without  affecting  other 
systems  significantly,  whereas  a  more  reactive  inhibitor,  such  as  p-chloro- 
mercuribenzoate,  is  valueless  for  producing  specific  metabolic  blockade.  The 
choice  of  SH  reagent  to  be  used  should  always  be  made  on  the  basis  of  the 
type  of  work  to  be  done.  Another  factor  to  be  considered  in  work  with  eel- 


GENERAL  CONSIDERATIONS  OF  THE  USES  OF  SH  REAGENTS    653 

hilar  preparations  is  the  penetrabihty  of  the  SH  reagent,  and  those  reagents 
should  be  chosen  that  have  the  most  likelihood  of  reaching  the  systems  to 
be  attacked.  Thus  iodoacetamide  is  often  a  better  choice  than  iodoacetate 
for  intracellular  inhibition  because  it  is  uncharged  and  probably  enters  cells 
more  readily. 

The  treatment  of  the  individual  SH  reagents  in  the  following  chapters 
must  be  eclectic  in  view  of  the  immense  amount  of  reported  work,  partic- 
ularly during  the  past  few  years.  The  attempt  will  be  made  to  select  the 
results  of  those  investigations  done  most  carefully  and  thoroughly,  and  to 
include  work  on  the  most  important  or  interesting  facets  of  inhibition  by 
SH  reagents.  A  third  aim  is  to  present  all  the  available  accurate  data  that 
may  aid  in  the  assessment  of  the  specificity  of  these  inhibitors  in  order  to 
use  them  more  profitably  in  complex  systems. 


CHAPTER  5 

OXIDANTS 


Many  reagents  have  been  used  to  oxidize  protein  and  enzyme  SH  groups 
for  the  purpose  of  either  estimating  these  groups  or  determining  the  relation- 
ship of  the  groups  to  the  enzyme  activity.  Most  of  these  oxidants  at  present 
are  of  little  importance  in  the  study  of  enzymes  or  metabolism,  mainly  be- 
cause of  their  lack  of  specificity  for  SH  groups.  General  over-all  oxidation 
of  an  enzyme,  involving  several  types  of  group  and  ending  in  partial  or  com- 
plete denaturation,  provides  no  useful  information.  If  oxidants  are  to  be 
used  for  the  specific  modification  of  SH  groups  it  is  necessary  that  the  choice 
of  oxidant  and  the  experimental  conditions  be  made  very  carefully.  The 
oxidizing  activity  of  the  reagent  must  be  neither  too  high  nor  too  low  (i.e., 
its  oxidation-reduction  potential  must  be  in  the  proper  range  relative  to  the 
SH  groups  under  the  selected  conditions)  and  the  ability  of  the  substance 
to  react  in  other  ways  with  the  enzyme  must  be  minimal.  Of  the  factors 
determining  the  rates  of  oxidation  of  SH  groups  and  the  specificity  of  an 
oxidant,  the  pH  and  the  temperature  are  the  most  important.  Some  of  the 
oxidants  that  have  been  abandoned  in  enzyme  work  might  well  be  applicable 
in  certain  studies  if  the  optimal  conditions  for  their  use  were  known. 

The  formation  of  enzyme  disulfide  groups  during  oxidation  requires  SH 
groups  that  are  close  enough  to  link  together  in  S — S  bonds,  or  are  so  located 
as  to  be  able  to  approach  each  other  readily.  The  SH  groups  may  be  on  the 
same  enzyme  molecule  or  on  different  molecules: 


SH  S 

Intramolecular  oxidation:  R,         ^  Xqx  >'-         R 


SH 


\ 


^red 


Intermolecular  oxidation: 


2  R— SH    -r    Xox     •*  *     R— S-S  — R    ^    Xred 


The  hydrogen  atoms  may  be  transferred  directly  to  the  oxidant,  or  may 
form  H+  ions,  the  oxidant  accepting  only  electrons.  A  lone  SH  group  on  an 
enzyme,  although  fully  exposed,  may  not  be  oxidized  if  sterically  it  cannot 
associate  with  another  SH  group.  The  formation  of  enzyme  aggregates,  or 

655 


656  5.    OXIDANTS 

actual  precipitation,  upon  oxidation  has  occasionally  been  taken  as  evidence 
for  intermolecular  disulfide  bonding,  but  this  is  perhaps  not  always  valid, 
since  the  oxidation  may  bring  about  a  dissolution  of  the  protein  structure 
leading  to  such  intermolecular  reactions  as  occur  during  any  type  of  de- 
naturation.  If  reversal  of  aggregation  can  be  induced  by  reducing  agents,  it 
is  more  likely  that  disulfide  bonds  are  responsible.  One  factor  of  primary 
importance  in  determining  whether  a  disulfide  bond  can  be  formed  is  the 
steric  relationship  between  the  interacting  groups.  The  C — S — S — C  group- 
ing is  not  linear,  or  even  planar;  the  S — S — C  bond  angle  is  around  107°  and 
the  dihedral  angle  between  the  two  C — S  bonds  is  close  to  90°  (due  to  the 
electrostatic  repulsion  between  unbonded  electron  pairs).  Thus  such  bonds 
will  be  formed  readily  only  when  the  residues  to  which  the  sulfur  atoms  are 
attached  can  assume  the  proper  orientations. 
The  thiol-disulfide  equilibrium: 

R— SH  ^  1/2  (R— S— S— R)  +  H+  -f  e- 

has  not  been  easy  to  determine,  due  to  abnormal  electrode  reactions  and 
the  usual  sluggishness  of  such  systems,  and  hence  values  for  the  oxidation- 
reduction  potential  vary  with  the  method  used.  It  appears  that  Eq  at  pH  7 
for  various  low  molecular  weight  thiols  generally  lies  between  —0.35  and 
0.0  (Calvin,  1954;  Clark,  1960,  p.  486).  The  values  of  E^'  for  protein  SH 
groups  are  not  known,  but  it  is  likely  that  they  would  lie  mainly  in  this 
range  also.  It  is  possible,  however,  that  some  SH  groups,  due  to  their  par- 
ticular molecular  environment,  may  have  positive  potentials,  i.e.,  would  be 
less  easily  oxidized  than  most  SH  groups  of  the  smaller  compounds.  It  is 
certainly  true  that  certain  SH  groups  on  proteins,  although  readily  acces- 
sible to  alkylation  or  mercaptide  formation,  are  not  oxidized  readily,  but 
whether  this  is  due  to  an  especially  high  oxidation-reduction  potential  or 
steric  factors,  as  discussed  above,  is  not  known.  The  values  of  Eq  depend 
strongly  on  the  pH,  which  must  be  taken  into  account  when  experiments 
are  run  at  pH's  varying  from  neutrality.  In  any  event,  the  oxidant  should 
have  a  rather  high  potential  (probably  0.2  or  higher)  in  order  to  oxidize 
the  susceptible  SH  groups  to  virtual  completeness.  On  the  other  hand,  it  is 
usually  desirable  to  oxidize  the  SH  groups  only  to  the  disulfide  stage.  Strong 
oxidants  can  occasionally  not  only  oxidize  SH  groups  to  sulfonate  but  at- 
tack enzyme  groups  other  than  SH  so  that  specificity  is  lost.  Thus  iodate 
oxidizes  gluten  and  thiolated  gelatin  mainly  to  the  disulfide  stage: 

6  R— SH  +  lOg^  -►  3  R— S— S— R  +  1+3  H^O 

but  further  oxidation  also  occurs  simultaneously: 

R— SH  +  IO3-  ->  R— SO3-  +  H+  +  I- 


OXIDANTS  657 

(Hird  and  Yates,  1961).  o-Iodosobenzoate  usually  oxidizes  only  to  the  di- 
sulfide stage  at  pH  7,  but  if  too  high  a  concentration  is  used,  or  the  pH  is 
much  below  7,  sulfinate  or  sulfonate  groups  are  produced  (Hellerman  et  at., 
1941).  In  addition,  methionine  residues  may  be  oxidized  to  the  sulfoxide 
stage.  Sizer  (1942  a,b,  1945)  studied  the  effects  on  enzymes  of  many  oxida- 
tion-reduction systems  over  a  wide  range  of  Eq  and  found  that  as  the  Eq 
is  increased  from  around  —  0.5  there  is  little  effect  on  activity  until  a  crit- 
ical value  is  reached,  which  is  +0.6  for  /?-fructofuranosidase,  +0.35  for 
intestinal  phosphatase,  and  +0.58  for  chymotrypsin,  inactivation  increas- 
ing rapidly  above  these  values.  Of  course,  it  is  not  entirely  a  matter  of  the 
Eq  ,  the  nature  of  the  oxidant  being  also  very  important.  The  enzymes 
used  are  not  those  containing  SH  groups  most  easily  oxidized,  but  it  does 
indicate  that  rather  strong  oxidants  must  be  used  with  many  enzymes. 

Oxidants  can  inhibit  enzymes  by  mechanisms  other  than  oxidation  of  SH 
groups.  They  may  (1)  oxidize  other  enzyme  groups,  (2)  be  chemically  in- 
corporated into  the  enzyme  (e.g.,  the  iodination  of  tyrosine  residues  by 
iodine),  or  (3)  inhibit  reversibly  by  any  of  the  mechanisms  observed  with 
nonoxidizing  inhibitors.  Other  enzyme  groups  susceptible  to  oxidation  are 
the  hydroxy!  groups  of  tyrosine  and  serine,  the  hydrocarbon  chain  of  leu- 
cine, the  indole  ring  of  histidine,  and  perhaps  the  amino,  guanidine,  and 
peptide  groups.  A  few  examples  will  be  mentioned  and  others  will  be  dis- 
cussed in  the  sections  on  the  individual  oxidants.  Lieben  and  Bauminger 
(1933  a)  showed  that  several  amino  acids  are  attacked  by  ^permanganate. 
During  the  oxidation  of  casein,  the  arginine  content  falls,  urea  and  dixan- 
thylurea  appearing.  Haas  et  al.  (1951)  emphasized  the  importance  of  tyro- 
sine and  tryptophan  in  the  actions  of  permanganate  on  proteins.  Although 
phenylalanine  is  quite  refractory,  tyrosine  and  tryptophan  are  oxidized,  as 
shown  by  changes  in  the  ultraviolet  spectra.  The  spectra  of  insulin  and  pep- 
sin treated  with  permanganate  (0.1  mM  at  pH  2)  change  in  a  manner  sim- 
ilar to  that  of  the  free  amino  acids,  so  it  is  likely  that  oxidation  of  these 
amino  acids  occurs  when  they  are  part  of  the  protein  structure.  Oxidation 
of  several  proteins  by  periodate  releases  formaldehyde,  which  probably  arises 
from  hydroxylysine  (Desnuelle  and  Antonin,  1946).  One  mole  of  ovalbumin 
reduces  30  moles  of  periodate  to  iodate,  the  protein  losing  all  of  its  cysteine 
and  cystine,  one  third  of  its  tryptophan,  and  a  small  fraction  of  its  tyrosine 
(Desnuelle  et  al.,  1947).  Oxidation  of  seralbumin  by  periodate  results  in 
destruction  of  certain  amino  acids,  producing  changes  in  spectral  and  elec- 
trophoretic  properties  (Goebel  and  Perlmann,  1949).  Periodate  releases  acet- 
aldehyde  from  chymotrypsin,  arising  from  terminal  threonine,  although  this 
is  not  responsible  for  the  inhibition  of  the  enzyme  inasmuch  as  it  occurs 
before  inactivation  starts,  and  in  this  case  no  ultraviolet  spectral  changes 
are  observed  (Jansen  et  al.,  1950,  1951).  Nitrous  acid  not  only  oxidizes  cer- 
tain enzyme  groups,  such  as  SH,  but  attacks  free  tyrosine  and  amino  groups 


658  5.    OXIDANTS 

(Philpot  and  Small,  1938;  Weill  and  Caldwell,  1945  a).  Hypochlorite  oxidizes 
a  number  of  amino  acids,  only  glycine  being  resistant,  and  spectral  changes 
occur  with  proteins  indicating  oxidation  of  tyrosine  and  tryptophan  residues 
(Lieben  and  Bauminger,  1933  b);  in  addition  it  chlorinates  amino  groups 
(Wright,  1926).  Sizer  (1942  b)  noted  in  his  work  with  many  oxidants  that 
SH  groups  are  by  no  means  the  only  susceptible  groups  on  enzymes,  the 
tyrosine  residues  being  particularly  oxidizable.  These  results  point  to  the 
importance  of  exercising  great  caution  in  the  choice  of  oxidants  and  condi- 
tions for  treatment  of  enzymes  if  specific  oxidation  of  SH  groups  is  desired. 
The  need  for  characterizing  well  the  enzyme  changes  —  e.g.,  disappear- 
ance of  SH  groups  as  determined  by  the  standard  methods,  or  alterations 
in  the  ultraviolet  spectrum  —  upon  oxidation  is  also  indicated. 

Oxidation  of  Enzymes   by  Molecular  Oxygen 

Thiols  and  enzyme  SH  groups  are  not  oxidized  by  Og  unless  certain  metal 
ions  are  present.  Thus  papain  is  oxidized  by  Og  in  the  presence  of  Cu++  or 
Fe+++  and  the  consequent  inactivation  of  the  enzyme  is  readily  reversed  by 
glutathione  (Hellerman  and  Perkins,  1934).  Enzymes  such  as  papain  and 
urease  were  the  earliest  studied  with  respect  to  the  effect  of  oxidation  on 
their  catalytic  activities,  and  this  work  led  to  the  concept  wherein  the  redox 
state  of  SH  groups  is  an  important  regulating  mechanism  in  cell  metabolism 
(Hellerman,  1939).  The  initial  over-all  reaction  may  be  written  as: 

2  R— SH  +  O2  ±5  R— S— S— R  +  H2O2 

but  the  hydrogen  peroxide  can  produce  further  oxidation: 

2  R— SH  +  H2O2  ^  R— S— S— R  +  2  H^O 

or  it  can  oxidize  other  components  present.  The  kinetics  of  the  Cu++-  and 
Fe+++-catalyzed  oxidations  are  complex  and  the  mechanism  is  not  comple- 
tely understood.  One  theory  involves  the  formation  of  a  Fe++-thiol  radical 
from  a  Fe+++-thiol  complex;  two  such  radicals  would  combine  to  form  the 
disulfide  and  free  Fe++,  which  is  reoxidized  by  O2  (Williams,  1956).  A  sec- 
ond theory  postulates  a  Fe++ (thiol )2  chelate  complex,  which  is  oxidized 
by  O2  to  the  ferric  complex,  within  which  electron  transfer  occurs  to  form 
the  disulfide  and  Fe++  (Martell  and  Calvin,  1952).  Since  some  type  of  com- 
plex between  metal  ion  and  thiol  must  occur,  it  is  evident  that  the  suscepti- 
bility of  various  SH  groups  to  this  type  of  oxidation  must  vary  greatly. 
It  should  be  noted  that  the  rates  of  such  oxidations  depend  on  the  nature 
of  the  buffer  used  and  the  pH. 

The  toxic  effects  of  high  tensions  of  Og  on  cell  metabolism  may  depend 
on  the  oxidation  of  enzyme  SH  groups.  Brain  respiration  is  slowly  inhibited 
by  O2  and  there  is  increasing  inability  of  the  tissue  to  oxidize  glucose,  pyru- 


OXIDANTS  659 

vate,  lactate,  fructose,  and  succinate  (Dickens,  1946  a).  It  was  suggested 
that  the  most  sensitive  system  is  perhaps  the  pyruvate  oxidase  due  to  the 
involvement  of  SH  groups.  This  inactivation  is  not  mediated  through  the 
H2O2  formed,  inasmuch  as  the  concentrations  are  never  great  enough  due 
to  the  catalase  present.  The  following  enzymes  are  inactivated  by  high  Og 
tensions:  succinate  dehydrogenase,  phosphoglyceraldehyde  dehydrogenase, 
choline  oxidase,  phosphoglucomutase,  and  other  SH  enzymes  (Dickens, 
1946  b).  Lactate  dehydrogenase,  malate  dehydrogenase,  D-amino  acid  oxi- 
dase, and  yeast  hexokinase  are  resistant.  It  is  interesting  that  malonate 
(1  mM)  and  Mn++  (0.25  mM)  protect  succinate  dehydrogenase  against  oxi- 
dation by  O2,  and  that  NAD  protects  phosphoglyceraldehyde  dehydrogen- 
ase, indicating  that  either  O2  or  some  intermediate  must  react  directly 
with  the  enzyme  SH  groups.  Haugaard  (1946)  also  found  a  good  correlation 
between  the  SH  nature  of  enzymes  and  their  susceptibility  to  O2,  the  follow- 
ing sensitive  enzymes  being  added  to  the  list  above:  a-ketoglutarate  oxidase, 
pyruvate  oxidase,  glutamate  dehydrogenase,  and  xanthine  oxidase.  Dickens 
stated  that  succinate  dehydrogenase  is  irreversibly  inactivated  by  0^,  but 
Haugaard  found  reactivation  by  cysteine  or  glutathione,  indicating  a  simple 
disulfide  formation.  Inactivation  of  certain  enzymes  during  extraction  and 
purification  is  probably  due  to  oxidation  by  Og  since  metal-chelating  agents, 
such  as  ethylenediaminetetraacetate  (EDTA),  are  able  to  protect  against 
such  inactivation. 

The  cytochrome  system  may  be  involved  in  the  inactivation  of  enzymes 
by  O2.  Since  cysteine  is  oxidized  by  O2  through  the  cytochrome  system  to 
form  cystine,  and  since  cystine  will  in  turn  oxidize  certain  enzyme  SH  groups 
(see  page  661 ),  enzyme  extracts  in  which  cysteine  is  present  may  be  unstable. 
Thus  cysteine  inhibits  succinate  dehydrogenase  (Potter  and  DuBois,  1943), 
but  in  mouse  kidney  homogenates  the  inhibition  shows  a  lag  period  which  is 
interpreted  as  due  to  the  necessary  oxidation  of  cysteine  to  cystine  by  the 
cytochrome  system  (Ames  and  Elvehjem,  1944  a,  b). 

Various  Minor  Oxidants 

A  number  of  strong  oxidants,  such  as  permanganate,  perchlorate,  dichro- 
mate,  and  related  compounds,  have  been  used  in  the  past  to  oxidize  enzyme 
groups.  Most  of  these  have  dropped  out  of  use  because  they  were  felt  to 
lack  specificity  toward  SH  groups,  or  other  groups.  Actually  none  of  these 
oxidants  has  been  studied  thoroughly  with  respect  to  what  enzyme  groups 
are  oxidized,  or  to  the  optimal  conditions  for  achieving  specificity.  It  is 
quite  possible  that,  at  certain  pH's  and  concentrations  and  temperatures, 
these  reagents  may  be  specific  oxidants.  Certainly  some  enzymes  are  quite 
susceptible  and  others  very  resistant,  and  it  would  be  interesting  to  know 
the  reasons.  For  example,  permanganate  at  10  mM  inhibits  a-amylase 
>  75%  (Di  Carlo  and  Redfern,  1947),  at  0.05  mM  inhibits  /^-amylase  85% 


660  5.    OXIDANTS 

(Ghosh,  1958),  at  1  mM  inhibits  green  gram  flavokinase  48%  (Giri  et  at., 
1958),  at  0.1  mM  inhibits  /?-fructofuranosidase  94%  (Sizer,  1942  a),  at 
10  mM  inhibits  /5-glycerophosphatase  100%  (Rao  et  al,  1960),  at  2  mM 
inhibits  Aerobacillus  hydrogenlyase  23%  (Crewther,  1953),  at  0.01  mM  in- 
hibits intestinal  phosphatase  95%  (Sizer,  1942  b),  at  1  mM  inhibits  Pseudo- 
monas  proteinase  100%  (Morihara,  1963),  and  at  0.1  and  1  mM  inhibits 
beef  liver  urocanase  13%  and  87%,  respectively  (Feinberg  and  Greenberg, 
1959).  Since  these  experiments  were  done  at  different  pH's,  temperatures, 
and  incubation  times,  it  is  difficult  to  compare  the  results  accurately.  In- 
deed, no  thorough  investigation  of  the  effects  of  pH  or  temperature  on  such 
oxidations  has  been  made.  The  moderate  inhibition  (12%)  of  liver  arginase 
by  5  mM  permanganate  was  believed  due  to  an  effect  on  the  Mn++  cofactor 
rather  than  on  the  enzyme  (Greenberg  et  al.,  1956).  Although  yeast  /?-fruc- 
tofuranosidase  is  so  sensitive  to  permanganate,  it  is  inhibited  only  16%  by 
10  mM  dichromate  (Sizer,  1942  a)  and  only  43%  after  90  min  incubation 
with  123  mM  periodate  (Myrback,  1957  b).  Dichromate  is  also  less  effective 
than  permanganate  on  /5-glycerophosphatase  (Rao  et  al.,  1960)  and  /?-amyl- 
ase  (Ghosh,  1958).  In  this  connection,  it  must  be  remembered  that  the  prod- 
ucts of  the  reduction  of  the  oxidant  may  also  be  inhibitory,  e.g.,  the  MnOg 
or  Mn++  from  permanganate.  The  results  of  Taylor  and  Gale  (1945)  on  E. 
coli  amino  acid  decarboxylases  are  interesting  in  that  the  effects  of  perman- 
ganate were  found  to  depend  on  the  substrate  used.  For  example,  0.1  mM 
permanganate  inhibits  the  decarboxylation  of  histidine  15%,  arginine  17%, 
glutamate  41%,  ornithine  98%,  lysine  100%,  and  tyrosine  100%.  Whether 
there  are  different  enzymes  or  different  effects  with  the  various  substrates 
is  not  known.  Permanganate  and  other  strong  oxidants  can  occasionally 
act  on  substrates  or  other  components  of  the  reaction.  This  is  illustrated  in 
the  effects  of  permanganate  and  p-benzoquinone  on  the  growth  of  Fusarium 
conidia  (Braune,  1963).  Both  are  inhibitory  alone  but  when  present  together 
nullify  each  other  and  may  actually  stimulate  growth.  This  was  shown 
not  to  be  due  to  some  oxidation  product  of  jo-benzoquinone  or  reduction 
product  of  permanganate.  The  formation  was  postulated  of  a  substance  X 
which  protects  against  p-benzoquinone  and  heavy  metal  ions.  Indeed,  treat- 
ment of  maleate  or  tartrate  with  permanganate  gives  rise  to  substance  X. 
Although  such  effects  in  cellular  systems  are  complex,  related  actions  must 
be  expected  in  certain  enzyme  systems. 

Results  with  nitrous  acid  are  difficult  to  interpret,  but  in  all  cases  the 
inhibition  progresses  very  slowly  (Myrback,  1926).  Whereas  the  of-amylase 
from  B.  suhtilis  cannot  be  reactivated  by  HgS  after  inhibition  by  nitrite 
(Di  Carlo  and  Redfern,  1947),  the  /9-amylase  of  barley  is  completely  reac- 
tivated (Weill  and  Caldwell,  1945  a).  In  the  former  case  it  was  concluded 
that  SH  groups  are  not  involved  in  the  inhibition,  and  in  the  latter  case 
that  they  are.  Similarly,  inhibitions  by  redox  dyes  may  or  may  not  be  at- 


DISULFIDES  661 

tributed  to  SH  oxidation,  but  because  of  the  molecular  complexity  of  most 
dyes  it  would  be  very  unlikely  that  specific  effects  on  SH  groups  could  be 
obtained.  Inhibitions  by  dyes  will  be  discussed  in  a  separate  chapter. 


DISULFIDES 

Cystine  oxidizes  certain  protein  SH  groups  and  was  first  used  for  the  de- 
termination of  these  groups  by  Mirsky  and  Anson  (1935).  They  found  that 
those  protein  SH  groups  reacting  with  nitroprusside  are  completely  oxidized 
by  cystine.  It  was  stated  by  Mirsky  and  Anson,  and  has  been  restated  by 
others,  that  cystine  is  one  of  the  most  specific  oxidants  of  protein  SH  groups. 
One  disadvantage  of  cystine  is  its  low  solubility;  dithioglycolate  was  sug- 
gested as  superior  in  this  regard  but  has  been  used  very  little.  Another  dis- 
advantage is  perhaps  that  the  oxidation-reduction  potential  of  the  cystine- 
cysteine  couple  is  not  high  enough  to  oxidize  all  the  SH  groups,  which,  if  it 
is  assumed  that  the  mean  potential  of  protein  SH  groups  is  similar  to  free 
cysteine,  must  be  true.  A  further  complication  is  the  formation  of  mixed 
disulfides  (see  page  639): 

E— SH  +  R— S— S— R  ±5  E— S— S— R  +  R— SH 

which  is  not  the  simple  oxidation  of  enzyme  SH  groups  previously  suppos- 
ed, a  portion  of  the  disulfide  being  bound  to  the  enzyme. 

A  few  results  on  enzyme  inhibition  are  summarized  in  Table  5-1.  The  in- 
hibitions are  not  to  be  taken  quantitatively  because  the  conditions  and  the 
incubation  times  vary  greatly.  Particularly  important  is  the  duration  of 
contact  between  the  enzyme  and  the  disulfide,  since  in  almost  all  instances 
the  reaction  has  been  found  to  proceed  very  slowly.  Rapkine  (1938)  found 
that  phosphoglyceraldehyde  dehydrogenase  requires  up  to  5  hr  for  maximal 
inhibition  by  GS,SG,  and  Hopkins  et  al'.  (1938)  showed  that  the  inhibition 
of  succinate  dehydrogenase  develops  steadily  over  2  hr,  and  probably  con- 
tinues after  that  time.  Whether  this  is  due  to  the  slowness  of  the  oxidation 
of  enzyme  SH  groups  or  to  secondary  factors  is  not  known.  One  reason  for 
the  lack  of  inhibition  of  certain  enzymes  by  disulfides  may  well  be  that  suf- 
ficient time  was  not  allowed  for  reaction.  Such  slow  reactions  limit  the  use 
of  the  disulfides  for  either  SH  group  determination  or  enzyme  studies.  In- 
sufficient examination  of  the  reversibility  of  disulfide  inhibitions  by  reduc- 
tion also  makes  it  difficult  to  evaluate  the  mechanism.  Rapkine  (1938)  found 
that  both  GSH  and  cysteine  reactivate  GSSG-inhibited  phosphoglyceral- 
dehyde dehydrogenase,  and  Hopkins  and  Morgan  (1938)  obtained  similar 
results  with  succinate  dehydrogenase,  but  reversibility  has  not  been  at- 
tempted in  most  work.  Many  enzymes  require  SH  groups  for  activity  but 
others  are  active  only  when  disulfide  bonds  are  formed.  Cytochrome  oxidase 


662 


5.    OXIDANTS 


P5 


M*   ^ 

2  ^ 


P5 


CO  CO 


•5  ^  ^ 


^  ^ 


^   * 


T3   i^ 


2        ^ 


I   O   ?3     §   -^ 


5^ 


5      0      0;,= 

in  s  ;^  H 


-^  to 

e 

e3 


^    00 


'S  -r! 


"  c 

C  ^  _2  ft 

03  "o  3  ^ 

>  Q  S  « 


^    e  05 


^^    «3      e8      O    -^ 
<M      IS  


oo 


0) 

.S    c 
>  Q 


^  M  s  s  -^  ;:h 

^  i 


H 


03    J4 
S     O 


c  ^^ 
9  ^ 


•S  s 

G  '-I 

5  ^ 

02  W 


oooot-oo«<i 


lOMOOOOOO^OeOiMO 
(MC0C5  ^»0  -<tlTj<coc5Tt< 


cS    — - 


o  o  o  o  ^ 


»0    M    »C 
(M    O 


o  o  o  o 

I— 1    (M    O    -^ 


03       ^ 


o 

o 

cc 

M 

CK 

CO 

O 

o 

.S  .S  o  -S  O  o  o 

-g    -g   M    "S   M   M   M 

O  O  O  O  O  O  O 


cs 

•S  o  o 

o 

•S  O 

•S  O  O 

M 

■g  W  M 

OT 

"S    «2 

"S    W    M 

OJ 

■^m  m 

02 

>>  cc 

>^.m  m 

O 

o  o  o 

o 

o  o 

o  o  o 

CO     >> 


P5 


3         4)    s    >.  -ti  :r3 


ft     (»      tH 
2      C      C      > 


B      o3 


S    s    3  Is 


^  pq  W  ffi  tf        ^ 


-:«        -^    e 


g  s  s 

"S  -►^  -^ 

^  ;s  3 

Ah  Ph  tf 


-^  PP  Ph 


o     cS 


a     e8 

V       (11 


(1h    ^-5 


<S, 


s 

SI 


^  t^  :2 

•■  nH  ft  «  « 

-5  :2  .s  s  a 

^  o  a  <  < 


Q  Q 


-g    ft 


2^2 


3 

?>  aj  S  4) 

ft  03  gr"  -t^ 

CS  «2  cS 

«  .S  a>  2 

•S  S  .S  1^ 

£  t«  £  -jS 


>s  >s    -  2 

w    ^    w    i^  ^  o    >> 

^2       ^      Jl^       03  ro       rH       r- 

^  |2i  JS| 

O         O  Ah  Ph  cc 


Ah 


1^ 


Ah  -g 

$  c 

m  .3 

O  o 

^  o 

S  CO 


be    o 


DISULFIDES  663 

normally  contains  disuL&de  groups  and  can  be  inactivated  by  cysteine,  GSH 
and  other  thiols;  reactivation  occurs  with  cystine  or  GSSG  (Cooperstein, 
1963).  This  may  also  be  the  case  with  lens  aminopeptidase,  which  is  inhib- 
ited rather  strongly  by  cysteine  and  GSH  but  not  at  all  by  GSSG  (Spec- 
tor,  1963).  Another  reason  for  the  lack  of  response  to  a  disulfide  is  that 
the  environment  of  the  enzyme  SH  group  may  be  unfavorable  for  the  ap- 
proach of  the  disulfide  or  may  affect  the  redox  potential  in  such  a  manner 
as  to  deter  the  interaction.  If  seralbumins  are  incubated  with  excess  cystine, 
the  SH  content  as  determined  by  p-mercuribenzoate  titration  drops  to  zero; 
half  the  SH  groups  are  lost  from  bovine  hemoglobin  (Isles  and  Jocelyn, 
1963).  GSSG  has  no  effect  on  either  type  of  protein.  Incidentally,  the  reac- 
tion of  cystine  with  bovine  seralbumin  leads  to  the  disappearance  of  1  mol- 
ecule of  cystine  for  each  pair  of  SH  groups  lost,  so  it  is  likely  that  the 
cysteine  formed  is  reoxidized  and  mixed  disulfides  are  formed  according  to: 

2  Cy— S— S— Cy   -  2  Prot— SH   ±-.  2  Cy— SH  -  2  Trot— S— S— Cy 
2  Cy— SH   ^      Cy— S— S— Cy    i-  2  H+  +  2  e- 

Cy— S— S— Cy   f  2  Prot— SH   ^-.  2  Prot— S— S— Cy  +  2  H^  -f  2  e- 

Some  evidence  for  the  formation  of  mixed  disulfides  during  the  inhibition 
of  enzymes  has  been  reported.  Beef  liver  catalase  contains  8  SH  groups  ti- 
tratable  with  p-chloromercuribenzoate.  Reaction  with  cystine-S^^  results  in 
the  oxidation  of  4  of  these  groups  to  disulfides  and  the  formation  of  4  mixed 
disulfides  {YM  et  al,  1961): 

SH^     SH  SH       SH  X— S— S         S— S         S  — S— X 

\     I       I     /  _^ \     I       I     / 

catalase^  -^     C— S  — S— C                   >.:                      catalase 

/     I       I     \  /     I      I     \ 

SH       SH  SH       SH  X— S— S        S— S         S— S— X 

This  inhibition  is  spontaneously  reversible;  cystine  at  0.037  mM,  pH  7,  and 
370  inhibits  maximally  17%  around  5  min  and  then  the  inhibition  decreases. 
Cystamine  monosulfoxide  likewise  appears  to  form  mixed  disulfides  with 
phosphoglyceraldehyde  dehydrogenase,  and  this  is  reversible  by  thiols  (Pihl 
and  Lange,  1962).  Cystamine  itself  inhibits  glucose  utilization  in  erythro- 
cytes and  this  block  is  believed  to  be  at  hexokinase  (Eldjarn  and  Bremer, 
1962).  The  inhibition  is  reversible  by  thiols,  and  mixed  disulfide  formation 
was  postulated;  however,  there  was  no  direct  evidence  for  this  as  opposed 
to  simple  oxidation.  Dithioglycolate  and  cystine  were  reported  to  form 
mixed  disulfides  with  myosin  ATPase,  but  inhibition  of  the  enzyme  activity 
does  not  occur  until  the  last  2  or  3  SH  groups  are  altered  (Barany,  1959). 
The  optimal  temperatures  and  pH's  for  reaction  of  enzyme  SH  groups 
with  disulfides  are  difficult  to  determine  in  our  present  state  of  knowledge. 
Apparently  the  temperature  coefficient  is  quite  high;  succinate  dehydrogen- 


664  5.    OXIDANTS 

ase  is  inhibited  20%  at  20°,  78%  at  30°,  and  93%  at  40^  by  100  mM  GSSG 
after  4  hr  incubation  (Hopkins  et  al.,  1938).  Thus  one  might  expect  incon- 
veniently slow  reactions  at  low  temperatures.  Yet  Barany  (1959)  ran  his 
experiments  with  ATPase  at  0^,  although  up  to  15  hr  was  sometimes  re- 
quired for  satisfactory  reaction.  Increase  in  the  pH  favors  the  oxidation  of 
enzyme  SH  groups  to  disulfides,  since  the  — S~  form  presumably  reacts 
more  readily.  Hence,  incubation  of  the  enzyme  with  the  disulfide  at  pH's 
around  9  may  be  useful  where  possible. 


PORPHYREXIDE  AND   PORPHYRINDIN 

Porphyrindin  was  originally  synthesized  by  Piloty  and  Schwerin  (1901  a, 
b,  c;  Piloty  and  Vogel,  1903)  but  only  much  later  attracted  attention, 
when  it  was  studied  by  Kuhn  et  al.  (1934)  as  the  first  clearly  demonstrated 
double  free  radical,  and  shown  by  Kuhn  and  Franke  (1935)  to  have  one  of 
the  highest  oxidation-reduction  potentials  among  organic  substances.  It  was 
introduced  as  a  reagent  for  the  determination  of  protein  SH  groups  by  Kuhn 
and  Desnuelle  (1938)  because  of  its  high  potential  and  applied  particularly 
by  Greenstein  (1938;  Greenstein  and  Edsall,  1940;  Greenstein  and  Jenrette, 
1942)  for  this  purpose.  Meanwhile  its  synthesis  was  improved  by  Porter  and 
Hellerman  (1939).  For  the  past  20  years  it  has  been  used  sporadically  to 
inactivate  enzymes  by  oxidation  of  the  SH  groups.  Possibly  it  has  been 
neglected  in  enzyme  studies  since,  although  it  is  probably  not  as  specific 
for  SH  groups  as  is  o-iodosobenzoate,  it  is  certainly  more  selective  than 
most  oxidants  and,  furthermore,  reacts  more  rapidly  and  more  completely. 

Chemistry 

The  structures  of  porphyrindin  and  porphyrexide  may  be  written  in  sev- 
eral different  ways  because  of  resonance.  Both  in  the  crystalline  state  are 
paramagnetic,  the  values  indicating  one  unpaired  electron  in  porphyrexide 
and  two  in  porphyrindin  (Kuhn  et  al.,  1934).  The  paramagnetism,  however, 
increases  with  temperature  (Miiller  and  Miiller-Rodloff,  1935),  suggesting 
equilibria  between  diamagnetic  and  paramagnetic  forms.  Thus  the  resonance 
structures  for  porphyrindin  may  be  written  as: 

HoC     O"  ~0     CH,  HjC     O'  b     CH3 

1       L  .11  I       1+  +11 

H3C-C— N  N— C— CH3  H3C-C— N                         N-C— CH, 

I       \\  //I  \       ■\  /-I 

C— N=N— C                      -« >-  C=N— N  =  C 

I        /  \        I  I        /  \        I 

HN— C— N  N— C  =  NH  HN=C— N                         N— C  =  NH 

H  H  H                         H 

(diamagnetic)  (paramagnetic) 

Porphyrindin 


PORPHYREXIDE    AND    PORPHYRINDIN  665 

while  for  porphyrexide  only  the  free  radical  form  is  possible: 

H3C    6 

I      U 
H,C— C— N 

I      *\ 

C  =  NH 

HN=C— N 
H 

Porphyrexide 

The  diamagnetic  form  is  more  stable  than  the  paramagnetic  by  about  0.56 
kcal/mole.  Magnetic  studies  have  indicated  the  free  electrons  to  be  fairly 
well  localized  and  not  diffusely  distributed  over  the  molecule.  Possibly  these 
free  electrons  contribute  to  the  color  of  these  substances:  porphyrexide  is 
red  and  porphyrindin  a  deep  blue.  Upon  reduction  the  color  disappears 
(leucoporphyrindin  may  be  slightly  yellow),  this  being  the  basis  for  the  col- 
orimetric  titration  of  SH  groups.  Porph>Texide  has  an  absorption  maximum 
at  460  m//  and  porphyrindin  at  653  m//  (Kuhn  and  Franke,  1935). 
Reduction  involves  the  change  from 

O-  O-  OH 

I  I  I 

— N+=    or   — N+—  to  — N  — 

and  a  disappearance  of  free  radicals.  The  oxidation-reduction  potentials  at 
pH  7  and  18°  are: 

Porphyrexide:      £"„'  =  +  0.725  v 
Porphyrindin:       E^'  =  +  0.565  v 

SO  that  porphyrexide  approaches  the  oxygen  potential  and  both  are  well 
above  most  systems  commonly  seen  in  biological  work.  The  oxidation  of 
simple  thiols  is  very  rapid,  cysteine  and  glutathione  reacting  almost  instan- 
taneously. 

Porphyrindin  is  not  very  stable  and  the  solid  crystalline  dye  should  be 
stored  at  low  temperatures  and  desiccated.  Greenstein  (1938)  pointed  out 
that  porphyrindin  solutions  are  stable  enough  for  an  hour  but  the  activity 
then  decreases.  Brand  and  Kassell  (1940)  reported  that  solutions  at  0^  show 
3%  deterioration  in  1  hr,  5%  in  2  hr,  and  9%  in  4  hr.  At  25^  deterioration 
occurs  at  a  rate  of  about  0.5%  per  minute.  Reactions  of  enzyme  SH  groups 
can  usually  be  carried  out  at  0°,  but  in  work  with  tissues  at  physiological 
temperatures  this  spontaneous  decomposition  must  be  borne  in  mind.  Por- 
phyrindin should  be  quantitatively  determined  in  all  accurate  work  since 
it  is  seldom  pure;  this  may  be  done  by  titrating  with  asorbic  acid  (Chinard 
and  Hellerman,  1954). 


666  5.    OXIDANTS 

Spiro  analogs  of  both  porphyrexide  and  porphyrindin  were  synthesized 
by  Porter  and  Hellerman  (1944)  and  found  to  have  high  oxidation-reduc- 

O"  O  0 

N— C— NH  / \     N=C— N=N— C=N, 


C— NH  ' '     C— NH  HN— C 

II  II  II 

NH  NH  NH 

Spiroporphyrexide  Spiroporphyrindin 

tion  potentials.  The  Ef^  at  pH  7  for  spiroporphyrexide  is  +  0.69  v,  the  po- 
tential increasing  at  lower  pH's.  The  spiroporphyrindin  is  quite  insoluble 
and  may  be  more  polymerized  than  indicated.  An  interesting  aspect  of  spiro- 
porphyrexide is  that  it  does  not  inhibit  urease,  despite  its  high  potential, 
indicating  possible  steric  effects  of  the  cyclohexyl  ring.  This  is  a  good  exam- 
ple that  not  only  is  oxidation-reduction  potential  important  in  the  oxida- 
tion of  SH  groups  on  enzymes  but  that  structural  configuration  is  a  factor, 
as  in  any  inhibition. 

Oxidation   of  Thiols  and   Amino  Acids 

Porphyrexide  and  porphyrindin  react  very  rapidly  with  thiols  at  neutral- 
ity and  the  end-point  is  generally  quite  sharp;  cysteine,  glutathione,  and 
cysteinylcysteine  are  titrated  quite  comparably  (Greenstein,  1938).  The  oxi- 
dation of  cysteine  at  pH  7.2  is  complete  within  3-60  sec  and  glutathione  is 
oxidized  only  slightly  more  slowly  (Brand  and  Kassell,  1940).  No  reaction 
under  ordinary  conditions  is  seen  with  cystine,  cysteate,  tryptophan,  hydro- 
xyproline,  histidine,  methionine,  serine,  phenylalanine,  or  threonine.  Tyro- 
sine, however,  is  oxidized  slowly  with  the  formation  of  a  pink-orange  color, 
the  reaction  taking  around  30  min  for  completion  (2  equivalents  of  porphy- 
rindin per  mole  of  tyrosine)  at  pH  7.2  and  0°.  Tyrosine  and  other  phenols 
are  oxidized  more  rapidly  in  alkaline  solutions  and  at  higher  temperatures, 
but  in  most  cases  the  reaction  is  much  slower  than  the  oxidation  of  SH 
groups  (Greenstein  and  Edsall,  1940).  At  pH  7.33  and  25^  the  oxidation  of 
tyrosine  may  be  fairly  rapid,  and  even  tryptophan  may  be  slowly  reacted 
(half-reaction  time  around  2  hr)  (Barron  et  al.,  1941).  Porphyrindin  can 
also  oxidize  a  variety  of  other  substances,  such  as  ascorbate  or  thiamine 
(Kuhn  and  Desnuelle,  1938),  and  in  complex  systems  or  cellular  prepara- 
tions the  effects  may  not  be  due  entirely  to  SH  group  oxidation. 

Oxidation  of  Proteins 

Kuhn  and  Desnuelle  (1938)  showed  that  native  ovalbumin  does  not  react 
with  porphyrindin  (in  common  with  other  SH  reagents)  but  that  following 


PORPHYREXIDE    AND    PORPHYRINDIN  667 

heat  denaturation,  titration  gives  results  comparable  to  other  methods  for 
total  cysteine.  This  was  confirmed  by  Greenstein  (1938),  who  used  urea  and 
guanidine  for  denaturation.  The  rapidity  of  SH  oxidation  was  noted.  Brand 
and  Kassell  (1940),  on  the  other  hand,  did  not  find  good  end-points  with 
denatured  ovalbumin  and,  because  of  the  pink  color  developed,  felt  that 
tyrosine  groups  are  also  oxidized.  This  was  criticized  by  Greenstein  et  al. 
(1940)  on  the  basis  that  far  too  much  porphyrindin  was  used,  and  they  em- 
phasized that  such  high  concentrations  are  probably  not  specific  and  should 
be  avoided.  The  method  was  somewhat  improved  (Greenstein  and  Edsall, 
1940;  Greenstein  and  Jenrette,  1942)  by  reducing  the  reaction  time  and 
lowering  the  pH  to  6.4-6.8  and,  using  myosin,  seralbumin,  ovalbumin,  and 
tobacco  mosaic  virus,  it  was  believed  that  accurate  titration  of  the  SH 
groups  could  be  achieved  with  little  interference  from  tyrosine  oxidation. 
Barron  et  al.  (1941)  treated  scarlet  fever  toxin  with  1  mM  porphyrindin  for 
1  hr  at  pH  7  and  found  that  the  activity  of  the  toxin,  as  determined  by  skin 
tests,  is  abolished.  Since  other  SH  reagents  do  not  inactivate  the  toxin,  it 
was  felt  that  oxidation  of  groups  other  than  SH  is  involved.  However,  the 
SH  reagents  used  (iodoacetate,  iodoacetamide,  hydrogen  peroxide,  alloxan, 
and  Cu++)  are  not  the  most  satisfactory  for  the  demonstration  of  SH  groups, 
so  this  evidence  is  not  conclusive.  In  order  to  achieve  specificity  toward  SH 
groups  it  is  advisable  to  (1)  avoid  alkaline  conditions,  (2)  reduce  the  reac- 
tion time  with  porphyrindin  as  much  as  possible,  (3)  use  as  low  concentra- 
tions of  porphyrindin  as  possible,  and  (4)  determine  the  disappearance  of 
SH  groups  by  some  secondary  titration. 

Inhibition  of  Enzymes 

Results  of  treating  enzymes  with  these  oxidants  are  shown  in  Table  5-2; 
one  cannot  help  but  be  surprised  that  so  little  use  has  been  made  of  these 
substances,  especially  during  the  past  few  years.  It  is  evident  that  quite 
low  concentrations  are  needed  for  those  enzymes  which  have  susceptible 
SH  groups  and  that  porphyrindin  and  porphyrexide  are  among  the  most 
potent  oxidant  inhibitors. 

Balls  and  Lineweaver  (1939  b)  attempted  to  titrate  papain  with  por- 
phyrindin but  found  that  no  clear  end-point  could  be  obtained  at  room 
temperature,  and  at  2-3°  no  bleaching  of  the  dye  occurred  during  several 
minutes  when  dilute  concentrations  were  used.  Higher  concentrations  pro- 
duce a  pink  coloration,  even  at  pH  4.6.  The  native  papain  SH  groups  are 
thus  not  reactive  with  porphyrindin;  neither  are  they  reactive  with  nitro- 
prusside.  On  the  other  hand,  iodoacetate  and  iodoacetamide  react  and  in- 
hibit; papain  treated  with  these  alkylating  agents  still  gives  rise  to  the  pink 
coloration,  indicating  that  tyrosine  is  oxidized,  although  not  necessary  for 
enzyme  activity.  E.  L.  »Smith  (1958)  concluded  that  the  SH  group  which  is 


668 


5.    OXIDANTS 


M 


O  05  C5 

fo  CO  fo 

- — •  Oi  a  C5 

■*       -H  r-  — . 


cS      he 
(D       CO 


.22       0)       0^ 

MOO 


CO      GO 
— (      02 


M 


=«    P    ^ 


Jri    J^    ^    -d 
O      O      O      C 

bc     bt}     bC     d 


O    O    O    W    K 


eo    CO    t- 

C5     C3     c> 


U  Xh        --I        F-H  Xh 


^ 

G5 

^3 

CO 

SO 

< 

>o 

05 

Oi 

Ti 

1— H 



<V 

^      ' 

U 

o3 

m 

<u 

0 

-*^ 

> 

T! 

;3 

a 

0 

C 

K   pq 


^   J    c   4-         ^   ^   J= 


S      S       3  rt      rt 

^   ^   W        pq   ffi 


^^ 


00000 


-H       O      O 


o 

odd 


G      «      XI 
I        I        I 

Ph      Pm      PL| 


X     X     t^     c     c 
I      I      I      I      I 

p^     PLi     CM     (1^     C^H 


Pi    Ph    Pm 


Ph     PLI 


-5    ^    -S 

cS      aj      cS 

C5    >H    P5 


cS      £P     cS      D 

P5    Ph    Ph    >H 


PlH 


o 
a;    ^ 

M  — 


.&    >> 


H    ^    -5    n1    -r^ 


•r^    bfi    ^  dj 


(11     o    rr   -T? 


jj    ^    o 


CO 

a) 

(» 

^ 

p 

cS 

4) 

0 

2 
Is 

0) 

in 

"S 

2 

^ 

>. 

.2 

0 

C 

2 

m 

fin 

< 

'3 
ft 

Ph 

■ft 
0 

0) 
0 

o 


Oh 


>> 

03 
T3 

0) 

x: 

-0 

0 

PORPHYREXIDE    AND    PORPHYRINDIN  669 

at  or  near  the  active  center  of  papain  is  in  a  high  energy  state,  perhaps  as  a 
thiol  ester,  and  this  may  explain  why  it  is  resistant  to  porphyrindin. 

The  reactive  SH  groups  of  urease  are  rapidly  oxidized  by  porphyrindin 
and  the  nitroprusside  test  becomes  negative  (Hellerman,  1939).  However, 
no  inhibition  occurs.  Porphyrindin-treated  enzyme  is  inhibited  by  jo-chloro- 
mercuribenzoate,  so  that  certain  SH  groups  required  for  activity  are  resist- 
ant to  porphyrindin.  If  an  excess  of  porphyrindin  is  used,  inactivation  oc- 
curs slowly  and  is  irreversible.  Oxidation  past  the  disulfide  stage  or  oxida- 
tion of  other  groups  is  possible.  Since,  p-chloromercuribenzoate  protects  the 
enzyme  from  high  concentrations  of  porphyrindin,  it  appears  that  SH  groups 
are  indeed  involved.  It  was  established  later  that  there  are  two  types  of  SH 
group  in  urease:  reactive  a  groups  not  necessary  for  enzyme  activity,  and 
less  reactive  b  groups  at  the  active  center.  Porphyrindin  reacts  with  the 
former  but  only  slowly  with  the  latter  at  higher  concentrations  (Hellerman 
et  al.,  1943).  If  urease  is  denatured  with  guanidine,  many  more  SH  groups 
appear  and  react  with  porphyrindin. 

Xanthine  oxidase  is  inhibited  readily  by  porphyrindin  but  this  is  not  re- 
versible with  cysteine  (Harris  and  Hellerman,  1956).  The  inhibition  by  o- 
iodosobenzoate  is  also  irreversible.  This  problem  comes  up  repeatedly  with 
inhibitions  by  oxidants  and  seems  on  the  surface  to  indicate  that  a  simple 
oxidation  to  the  disulfide  stage  does  not  occur.  However,  it  is  also  possible 
that  (1)  the  oxidation-reduction  potential  of  the  groups  involved  is  such 
that  cysteine  will  not  reduce  them,  (2)  the  enzyme  structure  is  altered  by 
the  formation  of  disulfide  bonds,  (3)  oxidation  past  the  disulfide  stage  has 
occurred,  or  (4)  the  formation  of  intermolecular  disulfide  linkages  prevents 
the  access  of  cysteine  to  the  group.  It  is  impossible  to  distinguish  at  this 
time  between  these  different  possibilities. 

Effects   on    Tissue    Function 

Porphyrindin  at  0.37  0.75  mM  produces  an  increase  in  the  contractile 
amplitude  of  the  frog  heart  and  this  effect  can  last  for  as  long  as  90  min 
(Mendez,  1946;  Mendez  and  Peralta,  1947).  Higher  concentrations  of  3.7- 
7.5  mM  bring  about  a  progressive  contracture,  the  heart  stopping  in  systole 
in  around  30  min.  The  rate  is  simultaneously  slowed.  The  atria  continue  to 
beat  after  the  ventricles  have  stopped.  These  effects  can  be  prevented  by 
glutathione  but  not  reversed,  as  expected.  In  these  respects  the  heart  re- 
sponds to  porphyrindin  much  as  it  does  to  other  SH  reagents.  The  site  or 
sites  of  action  are  not  known,  and  it  is  useless  to  speculate  since  the  relative 
sensitivities  of  the  possible  enzymes  involved  are  undetermined. 

A  few  miscellaneous  and  unrelated  observations  will  be  mentioned.  The 
short-circuit  current  and  potential  of  frog  skin  are  altered  by  oxidants  and 
reductants,  such  as  quinones,  dyes,  and  iodine,  but  there  is  little  effect  of 
porphyrindin  at  1  mM;  the  potential  may  drop  temporarily  but  the  cur- 


670  5.    OXIDANTS 

rent  is  unaffected  (Eubank  et  al.,  1962).  Porphyrindin  injected  at  a  dose  of 
200  mg  in  a  pregnant  mouse  produced  neuroblastic  necrosis  in  the  fetus, 
but  not  as  much  as  p-chloromercuribenzoate,  oxophenarsine,  or  o-iodosoben- 
zoate,  these  damaging  neuroblasts  in  a  pattern  similar  to  radiation  (Hicks, 
1953).  SH  reagents  usually  cause  blebbing  of  Sarcoma  37  ascites  cells.  Por- 
phyrindin has  no  effect  at  2  nxM  but  produces  symmetrical  blebs  at  8  roM 
(Belkin  and  Hardy,  1961).  Such  blebbing  involves  a  raising  of  the  entire  cell 
membrane  and  presumably  is  due  to  some  disturbance  in  water  transport. 

FERRICYANIDE 

Ferricyanide  has  been  used  widely  as  a  fairly  specific  oxidant  for  the  de- 
termination of  protein  SH  groups.  Furthermore,  it  is  commonly  used  as  an 
electron  acceptor  in  various  dehydrogenase  systems,  since  it  is  reduced  by 
some  of  the  components  in  electron  transport  before  cytochrome  c,  and  fer- 
rocyanide  has  often  served  as  an  electron  donor  in  studying  the  cytochrome 
system.  When  ferricyanide  or  ferrocyanide  is  used  for  such  purposes,  it  is 
important  to  consider  the  possibility  of  inhibition  of  the  enzymes  involved, 
particularly  the  dehydrogenases,  and  to  use  as  low  concentrations  as  pos- 
sible. The  first  use  of  ferricyanide  for  the  determination  of  SH  groups  was 
by  Flatow  (1928)  and  this  was  simplified  by  Mason  (1930)  so  that  the  ferro- 
cyanide formed  could  be  colorimetrically  estimated  after  transformation  to 
Prussian  blue,  this  being  suggested  by  Folin's  ferricyanide  method  for  blood 
glucose.  This  reaction  has  been  used  for  the  histochemical  localization  of  SH 
groups  but  is  not  very  satisfactory.  Anson  and  Mirsky  (1931)  noted  that 
hemoglobin  treated  with  ferricyanide  yields  a  globin  that  no  longer  reacts 
with  nitroprusside,  but  it  remained  for  Schiiler  (1932)  to  show  that  ferri- 
cyanide oxidizes  more  than  the  heme  group  and  that  globin  itself  reacts 
after  separation  from  heme.  Mirsky  and  Anson  during  the  next  10  years 
elucidated  the  nature  of  the  reactions  between  ferricyanide  and  proteins, 
and  applied  their  results  to  determination  of  protein  SH  groups. 

Chemistry 
The  oxidation  of  thiols  may  be  written  as: 

2  Fe{CN),^-  +  2  R— SH  ±^  2  Fe(CN)6''-  +  R— S— S— R  +  2  H+ 

When  the  SH  groups  are  on  different  molecules,  the  kinetics  are  complex 
and  the  detailed  mechanism  of  the  reaction  is  not  understood.  The  ferro- 
cyanide formed  is  usually  determined  by  addition  of  Fe+++,  forming  Prus- 
sian blue,  but  in  work  with  proteins  it  is  advisable  to  determine  also  the 
disappearance  of  SH  groups  with  nitroprusside  or  ??-chloromercuribenzoate, 
since  groups  other  than  SH  may  be  oxidized.  The  oxidation-reduction  po- 


FERRICYANIDE  671 

tential  of  the  ferricyanide-ferrocyanide  couple  is  +  0.36  v  and  does  not 
change  from  pH  4  to  10.  The  potential  is,  however,  rather  strongly  depen- 
dent on  ionic  strength.  The  potential  is  thus  sufficiently  high  for  SH  groups 
to  be  oxidized  completely  if  they  are  available  to  the  ferricyanide,  and  the 
reaction  is  generally  quite  rapid. 

Most  commercial  preparations  of  ferricyanide  contain  ferrocyanide,  which 
may  be  detected  by  the  Prussian  blue  method,  and  the  latter  should  be  re- 
moved by  addition  of  a  little  bromine  water  if  SH  determinations  are  done 
by  the  colorimetric  technique.  Solutions  of  ferricyanide  should  be  stored  in 
the  cold  and  dark  to  avoid  changes. 

Oxidation   of  Thiols  and   Proteins 

The  reactions  of  ferricyanide  with  proteins  have  direct  bearing  on  the 
effects  of  ferricyanide  on  enzymes  so  that  it  is  necessary  to  discuss  the  re- 
sults in  some  detail.  Although  titration  of  cysteine  and  glutathione  with 
ferricyanide  is  rapid  and  provides  good  end-points,  reactions  with  proteins 
may  not  be  so  clear-cut.  The  conditions  for  the  reaction  are  very  important. 
There  is  a  marked  effect  of  pH,  as  shown  originally  by  Mirsky  and  Anson 
(1936  a)  for  hemoglobin  (see  tabulation).  Indeed,  at  pH  6.8  one  may  spe- 


pH         Total  SH  groups  oxidized  (%) 


6.8 

0 

7.3 

28 

9.0 

44 

9.5 

65 

cifically  oxidize  the  heme  iron  to  form  methemoglobin  without  affecting  SH 
groups.  The  conditions  for  reaction  were:  83  mill  ferricyanide  incubated  with 
the  protein  for  30  min  at  room  temperature  —  all  reactive  SH  goups  are 
oxidized,  as  shown  by  titration  of  SH  groups  in  denatured  globin.  Kolthoff 
and  Anastasi  (1958)  have  also  observed  that  oxidation  of  the  SH  groups  of 
denatured  seralbumin  is  faster  at  pH  9  than  7.  They  noted  that  the  reac- 
tion is  accelerated  by  Cu++,  and  Katyal  and  Gorin  (1959)  found  in  a  study 
of  ovalbumin  that  iodide  also  catalyzes  the  oxidation  by  ferricyanide. 

Ferricyanide  is  not  specific  for  SH  groups,  however,  unless  the  conditions 
are  rigorously  controlled,  as  shown  early  by  Mirsky  and  Anson  (1936  b)  in 
proteins  not  containing  cysteine  (zein  and  serum  globulin)  but  nevertheless 
reducing  ferricyanide,  or  in  proteins  previously  treated  with  cystine  to  oxi- 
dize all  the  available  SH  groups.  These  groups  are  oxidized  more  slowly  than 
the  SH  groups  and  are  more  difficult  to  oxidize  (e.g.,  milder  oxidants  than 


672  5.    OXIDANTS 

ferricyanide  will  not  oxidize  them),  but  their  total  reducing  capacity  (the 
amount  of  ferricyanide  they  can  reduce)  is  often  greater  than  for  the  SH 
groups.  Furthermore,  the  rate  and  degree  of  oxidation  of  these  non-SH 
groups  depend  on  the  same  factors  as  reaction  with  SH  groups;  thus,  the 
rate  is  accelerated  by  rise  in  pH,  rise  in  temperature,  and  denaturation, 
Native  ovalbumin  is  not  oxidized  at  all  bj^  ferricyanide,  but  denatured  oval- 
bumin treated  with  cystine  to  remove  SH  groups  reduces  ferricyanide.  In 
other  words,  these  other  groups  become  available  during  unfolding  of  the 
protein.  /5-Lactoglobulin,  which  contains  2  SH  grovips  per  molecule  (molec- 
ular weight  of  37,000),  reacts  very  slowly  with  ferricyanide  in  the  native 
state  but  rapidly  in  the  presence  of  urea  or  guanidine  (Leslie  et  al.,  1962  a). 
The  stoichiometry  indicates  that  the  SH  groups  are  oxidized  beyond  the 
disulfide  stage.  Since  the  reaction  was  carried  out  under  fairly  mild  condi- 
tions (0.1-0.8  mM  ferricyanide,  pH  7,  37^,  and  30-45-min  incubation),  it 
is  evident  that  one  cannot  generally  assume  the  simple  formation  of  disul- 
fides from  the  actions  of  ferricyanide  on  enzymes.  Ferricyanide  can  oxidize 
tyrosine  and  tryptophan,  but  not  histidine,  and  the  characteristics  of  the 
oxidation  parallel  oxidations  of  proteins.  Mirsky  and  Anson  suggested  that 
tyrosine  and  tryptophan  are  the  residues  responsible  for  ferricyanide  re- 
duction. Gelatin,  which  contains  no  (or  very  little)  tyrosine  and  no  tryp- 
tophan, scarcely  reacts  with  ferricyanide,  supporting  this  view.  Anson 
(1939  b)  observed  that  at  pH  9.6,  where  much  previous  work  had  been 
done,  the  oxidation  is  nonspecific,  but  that  at  pH  6.8  combined  with  the 
treatment  of  the  protein  with  Duponol  PC  the  SH  groups  react  rapidly  and 
specifically  if  not  too  much  ferricyanide  is  used  (2-5  mM  is  best);  under 
these  conditions  there  is  no  reaction  with  cystine,  tyrosine,  tryptophan,  or 
proteins  that  do  not  contain  cysteine.  The  specificity  of  SH  group  oxida- 
tion could  also  be  shown  by  pretreatment  of  denatured  ovalbumin  with 
iodoacetamide,  following  which  ferricyanide  is  no  longer  reduced  by  the 
protein.  Katyal  and  Gorin  (1959)  also  demonstrated  such  specificity  by 
blocking  SH  groups  with  p-chloromercuribenzoate.  Mirsky  (1941)  discussed 
the  method  in  detail  and  showed  that  when  properly  run  the  oxidation  oc- 
curs within  1  min.  Barron  (1951)  has  also  reported  in  detail  his  modifica- 
tion of  the  method.  Various  oxidations  by  ferricyanide  have  been  reviewed 
by  Thyagarajan  (1958). 

Inhibition   of  Enzymes 

One  must  conclude  from  the  results  with  proteins  that  application  of  ferri- 
cyanide to  enzymes  cannot  be  done  haphazardly  if  specific  effects  on  SH 
groups  are  to  be  anticipated.  Unfortunately  most  studies  have  used  ferri- 
cyanide along  with  numerous  other  inhibitors  under  the  same  conditions 
of  pH,  temperature,  and  incubation  time,  without  considering  that  rather 
stringent  conditions  have  been  proposed  for  the  use  of  ferricyanide.  Some 


FERRICYANIDE  673 

inhibitions  are  summarized  in  Table  5-3.  Certain  enzymes  which  possess  SH 
groups  reactive  with  other  reagents,  e.g.  urease,  are  not  inhibited  by  even 
high  concentrations  of  ferricyanide.  One  might  imagine  ferricyanide  to  be 
unable  to  gain  access  to  the  SH  groups.  Ferricyaiade  is  not  only  a  fairly 
large  ion  but  has  a  strong  negative  charge.  If  the  enzyme  SH  group  occu- 
pied a  region  of  high  negative  charge,  this  might  repel  the  ferricyanide  and 
reduce  the  reaction.  Indeed,  one  must  always  consider  the  possibility  that 
ferricyanide  inhibits  certain  enzymes  by  mechanisms  other  than  oxidation, 
and  related  more  to  its  charge  and  structure.  For  example,  it  would  not  be 
so  surprising  if  ferricyanide  inhibits  succinate  dehydrogenase  to  some  ex- 
tent because  it  can  interact  with  the  cationic  groups  normally  binding  suc- 
cinate. One  notes  also  that  ferrocyanide  generally  inhibits  aconitase  more 
strongly  than  does  ferricyanide,  and  here  redox  reactions  may  be  of  no  sig- 
nificance (Rahatekar  and  Rao,  1963).  On  the  other  hand,  some  enzymes 
are  inhibited  just  as  rapidly  and  completely  by  ferricyanide  as  by  the  more 
commonly  used  SH  reagents;  myosin  ATPase  is  one  example  (Singer  and 
Barron,  1944).  The  inhibitions  of  papain  and  aldolase  are  quite  reversible 
with  cysteine,  indicating  that  reversible  oxidation  is  the  mechanism  of  the 
inhibitions.  Oxidation  of  coenzymes  or  cof"ctors  can  also  occur.  Ferricyanide 
can  directly  oxidize  NADH  but  the  rate  is  slow  (Schellenberg  and  Heller- 
man,  1958).  In  the  case  of  homogentisicase  it  may  well  be  the  Fe++  that  is 
oxidized  but,  on  the  other  hand,  there  appears  to  l)e  a  tyrosine  phenolic 
group  at  the  active  site  (Tokuyama,  1959). 

It  is  interesting  that  Weill  and  Caldwell  (1945  b)  report  /5-amylase  to  be 
not  readily  inhibited  by  either  ferricyanide  or  Cu+^  alone  but  strongly  in- 
hibited when  both  are  present,  even  when  the  ferricyanide  is  at  a  concen- 
tration noninhibitory  by  itself  (see  accompanying  tabulation).  Could  this 


Ferricyanide  Cu++ 

(mM)  (mM) 


%   Inhibition 


0.2 

— 

12 

— 

0.32 

4 

0.02 

0.32 

93 

relate  to  the  observation  of  Katyal  and  Gorin  (1959)  that  Cu++  accelerates 
the  action  of  ferricyanide  ?  Or  does  the  Cu"^+  in  some  manner  alter  the  en- 
zyme structure  so  that  ferricyanide  can  attack  the  SH  groups  more  easily? 

Effects  on   Cellular   Metabolism   and    Function 

Mendel  (1937)  reported  that  Balogh  mouse  tumor  glycolysis  is  markedly 
depressed  by  10  mM  ferricyanide  and  that  the  inhibition  is  maintained 


674 


5.    OXIDANTS 


fa 


P5 


fa 


O  O  O  O  O 

c^  c^  c^  c3  cS 

P5  tf  «  D5  tf 

""O  'C  '^  '^  "^ 


f— *      c5     cd     cfi     c3     rf 
—    ^    ^    ^    ^    ^ 

OJ      4)      tU      0)      <B 


cS      c3      C3      cS      cS 
2      -G-^-Sj3_C^^_C^-fi 


Pi   W 


o 

o 

o 

O 

o 

fe- 

^^ 

^ 

t2 

trt 

rt 

c. 

rt 

d 

es 

X 

o 

-« 

T3 

^ 

T-5 

tf 
TJ 

« 
TJ 

05 

^ 

TJ 

fi 

C 

C 

c 

'ci 

- 

k4 

ID 

C 
c3 

03 

U 

(1 

t^ 

o 

c 

o 

Ti 

0) 

^ 

« 

M 

« 

03 

C5 

o 

_o 

c 

5 

"S 

« 

c3 

6 

'S 

fs^PiCiiPiPifiiPiPiJbMPj^O 


5-ICOGOlOOOOO 


C<l    O      O      O      00 


O     O    O    ■^     (N 


lO  »o  lo  irj  ic  Lo  o  in  »c  LQ    CO 

oooooooooo    — < 


s 


V1 

o 

Si 

S 

? 

o 

o 

w. 

03 

CS 

(S 

s 

o 

'^ 

;■-» 

5^ 

c 

-u 

5=5 

X> 

IC      IC      M 

odd 


S     > 


c     p,    2 


(1^   ft^-^Q^^a5a,CQ>HfiH^  P5  oq  « 


c 

<M 

r^ 

u 

3 

OJ 

e 

cS 

O 

pq 

oq 

m 

a 

N 
fa 


c    .ii 


<t3    <3 


C     T3 


c8       P<     ■— 


O     .3     .5 


a 


FERRICYANIDE 


675 


eo 

CO 

CO 

»o 

»c 

05 

05 

C5 

o 

k4 

b. 

C5 

■'S" 

^ 

00 

o 

1 

s 

(M 

1— 1 

05 

05 

c3 

o 

'5 

C5 

1 

IS 

c3 
C5 

a: 
C 
ki 

0) 

T3 

CD 

c 

CO 

o 

o 

o 

;4 
O 

O 

Q 

IS 

CO 

03 

> 

O 
C 

2 

Is 

>• 

c 

Hargreaves  ( 
Mounter  and 
Mounter  and 

C 

;5 

C 

0) 

XI 

1 

o 

O 
02 

^    (M  C- 

'c  2  -^ 

S    CB    § 


t-     C    O    O    O     O 


1— I    fO     O     (M     O 


o  o  d 


00  ^   ,-, 


O     O    O    "H    <M 


?5, 


o 

c$ 

+=> 

o 

ft 


P    X. 


J  .S    g 

.§  ^  1 


C3      P 


w  w 

ft 

ft 

— 

— 

-u 

-kJ 

4i 

Ol 

02 

O) 

cS 

rt 

cS 

OJ 

<D 

aj 

>H 

;^ 

K^ 

a,  M 


bH  ,-= 


o 
g 


3 

4) 

M 

cc 

c3 

O 

c: 

OJ 

T3 

"3 

a; 

05 

CO 

<0 

05 
-2 

a5 

ft 

'>. 

cS 

cS 

03 

'3 

03 
O 

■ft 
P 

iS    S    c 


p*.  ft  g  t:  s 


^^07; 


j   <1   •<! 


0000 


CO 

0 

ID 

CO 

c3 

43 
00 

-0 

2 
0 

'a 

t 

10^ 

0 

cS 

03 

0 

c 

1 

<0 

e 
0 

X 

0 

i 

C 
0 

4> 

0 

T3 

"E 

0 

'S 

-2 

0 

ft 

X 

0 

P 

ft 

S 

0 

Q 

0 

0^ 

fe 

<3i. 

T!    .2 


j«  a 


676 


5.    OXIDANTS 


« 


fe 


S 
c 


05     Oi 


o 


O    ?3 


t5    S   ^    ^    M 


C  Oi  -<*  M 

eg  i-H  O  :C3 

'^  a      tl  r^  S 

=5   -S   s    c   =:  -2  oi  -^. 


S  05    s 


Q   fe   W   OQ  M   M 


^  CO     ^      »0     t-     -H      ■* 


-^    rN   (M   'Ki 


.0     c 


> 


W  3 


!M     CO     tJ< 

O     -H     O     iM     --H 


?^    S 


S       <D     -S 


T3     rt      cc      u 


h-1   PM 


0 

0 

_>; 

c3 

8 
g 

0 

?^ 

M 

& 

(U 

42 

-« 

%- 

oj 

« 

0) 

c3 

0 

^ 

S 

P^ 

< 

pq 

P^ 

;h 

. • 

03 

^-^ 

, , 

»C 

cS 

"o 

t> 

00 

0 

t/2 

T3 

+3 

0 
10 

_ 

05 

10 

C5 

05 
CO 

C 

C5 

0 

05 

fl 
£ 

"e 

ce 

S 

s 

eS 

0 

<4) 

c3 

e3 

^ 

0 

c3 

Jj 

(1 

s 

"0 

e« 

0 

PlH 

c3 

0 
1— 1 

s 

-y       ID 


s 

1^ 

<u 

^        S 

eS 

"o 

ID      D 

c 

(» 

e 

>     ~' 

0 
^ 

^ 

0 

CO 

0 

IS 
0 

M 
0 
ft 

0 

T3 

a 

0) 

1 

'i 

e3 

0 

SJ 

03 

1 

2 

4) 
03 

(» 

0 

0 

"So 

CO 

03 

T3 

c 

Ol 

12 

'3 

CO 

4) 

0 

0 

T3 

CI 

4) 
03 

'S 

iS 

cS 

0 

M 

<u 

c* 

0 

c< 

_e 

-tJ 

0 

03 

0) 

0 

<>!? 

JU 

c 

2 

^ 

S 

a 

^J 

■*^ 

-t-3 

c 

•^ 

W 

K 

fl 

'0 
1 

"a 

T3 

c 
'S 
0 

> 

03 
> 

e3 

0 

0 

03 

<1 

■3 
& 

c8 

0 

^ 

"5 

T3 

3 
u 

0 

ft 

H 

133 

2 

^ 

1^; 

PL^ 

P4 

PLH 

CO 

P^ 

P-l 

Plh 

OJ 

H 

<5i. 

^ 

FERRICYANIDE  677 

when  the  ferricyanide  is  washed  out.  Ferricyanide  slowly  injected  intraven- 
ously into  mice  (1.5  g/kg  of  sodium  salt)  produces  no  disturbance  of  tumor 
metabolism,  but  when  the  various  tissues  are  treated  with  re++"'",  only  the 
tumor  turns  blue.  The  anaerobic  glycolysis  of  no  other  tissue  is  depressed 
by  ferricyanide,  which  in  this  respect  differs  from  other  glycolytic  inhibitors 
(Mendel  and  Strelitz,  1937).  Renal  medulla  was  studied  particularly  because 
it  has  a  significant  rate  of  aerobic  glycolysis  and  ferricyanide  was  found  to 
have  no  effect  (and  in  some  cases  even  stimulated  somewhat).  This  action 
was  not  investigated  further  until  Birkenhager  (1959,  1960)  attempted  to 
locate  the  site  of  inhibition.  He  confirmed  that  10  mM  ferricyanide  does 
indeed  inhibit  both  aerobic  and  anaerobic  glycolysis  in  Walker  and  Crocker 
tumors,  but  not  in  Ehrlich  ascites  cells,  and  further  showed  that  it  inhibits 
the  extra  glycolysis  brought  about  by  dinitrophenol.  The  respiration  in  the 
presence  of  glucose  rises  30-60%  in  the  presence  of  ferricyanide  in  the  two 
former  tumors,  but  not  in  the  ascites  cells.  The  problem  of  what  happens 
to  the  glucose  taken  up,  since  this  is  not  depressed  as  much  as  lactate  for- 
mation, remains  unsolved.  A  small  accumulation  of  pyruvate  was  found 
under  aerobic  conditions  but  not  anaerobically,  and  no  other  glycolytic  in- 
termediates could  be  detected.  Use  of  glucose- 1-C^*  and  glucose-6-C^*  point- 
ed to  the  conclusion  that  ferricyanide  either  directly  or  indirectly  inhibits 
glycolysis  at  the  level  of  phosphohexose  isomerase  or  phosphohexokinase; 
this  would  make  more  hexose  phosphate  available  for  the  pentose  phosphate 
shunt.  However,  aldolase  was  found  to  be  very  sensitive  to  ferricyanide 
(88%  inhibition  at  0.5  mM)  and  addition  of  aldolase  to  a  tumor  extract  in 
which  glycolysis  has  been  abolished  by  ferricyanide  leads  to  recovery.  Phos- 
phoglyceraldehyde  dehydrogenase  is  not  sensitive  to  ferricyanide  nor  does 
its  addition  reverse  the  glycolytic  inhibition.  Birkenhager  ascribed  the  dif- 
ference between  cells  and  extracts  in  susceptibility  to  ferricyanide  as  due 
to  permeability  factors.  Certainly  one  might  expect  an  ion  such  as  ferri- 
cyanide to  enter  cells  with  difficulty.  However,  the  initial  observation  of 
Mendel  that  tumor  tissue  is  specifically  sensitive  to  ferricyanide  remains  to 
be  explained.  If  such  a  difference  exists,  it  must  be  due  to  ferricyanide 
penetrating  into  tumor  cells  more  readily,  since  none  of  the  enzymes  con- 
sidered to  be  the  point  of  attack  differs  markedly  in  tumor  cells  compared 
with  normal  tissues. 

Inasmuch  as  ferrocyanide  is  presumably  formed  in  tissues  during  the 
reduction  of  ferricyanide,  the  effects  of  ferrocyanide  on  the  tricarboxylate 
cycle  may  play  a  role  in  any  over-all  action.  Martin  (1955)  noted  that 
growth  of  Aspergillus  niger  is  inhibited  by  ferrocyanide  at  concentrations 
below  0.002  mM.  However,  acid  production  may  not  be  simultaneously  in- 
hibited, and  is  depressed  50%  only  at  0.4  mM.  An  accumulation  of  citrate 
is  actually  observed  at  1  mM  ferrocyanide  and,  at  this  concentration,  iso- 
citrate  dehydrogenase  is  inhibited  100%  (Ramakrishnan  et  al.,  1955).  In 


678  5.    OXIDANTS 

A.  tereus  0.1  mM  ferrocyanide  has  no  effect  on  glucose  utilization  but  in- 
creases the  yield  of  itaconate,  due  presumably  to  the  piling  up  of  citrate 
(Bentley  and  Thiessen,  1957).  The  uptake  and  metabolism  of  itaconate  are 
inhibited  by  ferrocyanide,  which  is  reasonable  on  the  basis  of  the  inhibition 
of  isocitrate  dehydrogenase,  and  the  assumption  that  itaconate  feeds  into 
the  cycle  (Shimi  and  Nour  El  Dein,  1962).  There  has  been  very  little  work 
on  the  effects  of  either  ferro-  or  ferricyanide  on  cycle  enzymes,  but  it  ap- 
pears likely  that  ferrocyanide  blocks  tricarboxylate  steps  selectively,  where- 
as ferricyanide  less  potently  inhibits  pyruvate  oxidation.  More  work  should 
be  done  on  these  effects  since  no  other  specific  inhibitor  of  isocitrate  dehy- 
drogenase is  known. 

A  few  miscellaneous  observations  on  ferricyanide  may  be  interpreted 
when  more  is  known  of  the  basic  actions.  Thus,  10  mM  ferricyanide  inhibits 
P^2  incorporation  into  phospholipids  19%  while  stimulating  respiration  17% 
in  M.  tuberculosis,  in  this  way  acting  more  like  the  uncoupling  agents  (azide, 
dinitrophenol,  and  arsenate)  than  the  common  SH  reagents  (Tanaka,  1960). 
Whereas  other  oxidants  and  SH  reagents  frequently  cause  mitochondrial 
swelling,  ferricyanide  is  without  effect,  which  could  scarcely  be  due  to  per- 
meability factors  (Rail  et  al.,  1962).  Porphyrin  synthesis  from  porphobili- 
nogen is  strongly  inhibited  by  1  mM  ferricyanide;  this  may  be  partly  the 
result  of  direct  oxidation  of  the  porphyrins  (Rimington  and  Tooth,  1961). 
The  eggs  of  Urechis  and  Hemicentrotus  are  very  sensitive  to  ferricyanide, 
0.02  mM  elevating  the  fertilization  membrane  in  the  former  and  causing 
cytolysis  in  the  latter  (Isaka  and  Aikawa,  1963).  The  dorsal-ventral  modifi- 
cation i^roduced  hy  ferricyanide  in  Dendraster  eggs,  whereby  either  dorsal 
induction  or  ventral  inhibition  is  manifest,  is  similar  to  that  produced  by 
iodoacetate  or  iodine,  but  is  unexplainable  since  the  factors  involved  in 
bilaterality  are  completely  unknown  (Pease,  1941).  The  effects  of  ferricy- 
anide on  the  naturally  occurring  quinones  must  occasionally  be  important. 
For  example,  ferricyanide  potentiates  very  markedly  the  growth-inhibiting 
activity  of  menadione  on  yeast,  due  to  the  fact  it  reoxidizes  the  reduced 
menadione  and  hence  maintains  the  naphthoquinone  in  the  active  form 
(Kiesow.  1960  b). 

IODINE 

Iodine  has  been  used  more  frequently  than  the  oxidants  previously  dis- 
cussed for  the  oxidation  of  protein  8H  groups  and  in  enzyme  studies,  and 
yet  it  seems  under  most  conditions  to  be  less  specific  than  the  others.  Al- 
though it  is  quite  a  potent  inhibitor  of  many  enzymes,  unless  one  can  de- 
termine if  a  particular  protein  group  is  oxidized,  or  otherwise  attacked,  the 
information  obtained  is  negligible.  Another  complication  in  the  use  of  iodine 
is  the  multiplicity  of  forms  in  solution  and  the  difficulty  in  characterizing 
the  nature  of  the  oxidation  reaction. 


IODINE  679 

Chemistry 

Iodine  is  soluble  to  the  extent  of  1.33  mM  in  water  at  20°,  which  is  much 
less  than  the  other  halogens.  There  is  interaction  with  the  water,  which 
initially  was  written  as  a  hydration: 

I,  +  H,0  ^  I,  ■  H^O 

but  evidence  pointing  to  the  highly  polarized  state  of  iodine  in  the  complex 
has  suggested  the  following  reaction: 

I,  +  H3O  t:,  1+  .  H,0  +  I- 

The  equilibrium  constant  for  this  reaction  has  been  estimated  as  roughly 
lO^^*'.  Iodine  may  also  undergo  hydrolysis: 

I2  +  H2O  ^  I-  +  HOI  +  H+ 

the  equilibrium  constant  being  3  X  10~^^.  The  hypoiodous  acid  formed  has 
a  p^^  of  12.3  so  is  little  ionized  at  physiological  pH's.  The  hypoiodous  acid 
can  also  go  to  iodate,  especially  in  alkaline  solution: 

3  HOI  ->  IO3-  +  2  I-  +  3  H+ 

A  third  reaction  of  iodine  is  with  the  iodide  ion: 

I,  +  I-  ±^  I3- 

to  form  the  triiodide,  which  is  the  principal  reason  for  the  greater  solubility 
of  iodine  in  KI  solutions.  The  equilibrium  is  given  by: 

(I2)  (I-) 

1.38  X  10-3 


(I3-) 


Thus  iodine  would  be  soluble  in  50  mM  KI  solution  to  the  extent  of  46  mM, 
an  appreciable  increase  over  the  1.33  mM  in  water.  The  production  of  nas- 
cent oxygen  by  the  reaction: 

I2  +  H2O  ->  2  H+  +  2  I-  +  O 

which  has  been  believed  to  be  involved  in  protein  oxidations,  does  not  occur 
with  iodine,  although  the  other  halogens  react  to  some  extent  in  this  way. 
In  most  biological  work,  iodine  is  dissolved  in  fairly  strong  KI  or  Nal  so- 
lution. This  not  only  serves  to  increase  the  solubility,  but  limits  the  fraction 
of  the  iodine  in  other  forms  (HOI,  IO3",  and  I+'HaO);  the  principal  form 
here  is  presumably  the  Ig"  anion.  However,  there  will  always  be  significant 
concentrations  of  iodine  present.  The  relative  importance  of  the  Ig  and  Ig" 


680  5.    OXIDANTS 

forms  in  the  oxidation  of  SH  groups  is  not  known.  The  oxidation-reduction 
potential  for  iodine  varies  with  the  type  of  reaction  in  which  it  participates 
and  the  pH,  but  is  usually  sufficiently  high  to  oxidize  any  accessible  SH 
groups.  It  is  important  in  certain  enzyme  studies  to  realize  that  iodine  may 
disappear  fairly  rapidly  from  solution,  independently  of  reaction  with  or- 
ganic materials;  such  is  favored  by  lack  of  iodide  and  high  pH. 

Reaction  of  Iodine  with  Thiols 

Iodine  is  able  to  oxidize  SH  groups  to  four  different  states:  the  disulfide 
(S — S),  the  sulfenate  (SO"),  the  sulfinate  (SOg"),  and  the  sulfonate  (SO3-). 
Apparently  it  is  quite  easy  to  oxidize  beyond  the  disulfide  state  with  iodine. 
The  stoichiometry  of  a  particular  reaction  will  depend  not  only  on  the  state 
of  oxidation  of  the  SH  groups,  but  also  on  the  degree  of  reduction  of  the 
iodine,  this  varying  with  the  pH.  For  example,  the  following  reactions  can 
be  written  for  oxidation  to  the  disulfide  state: 

I2  +  2  R— SH  ->  R— S— S— R  -I  2  I-  +  2  H+ 

2  I2  +  2  R— SH  +  H2O  ->  R— S— S— R  +  HOI  +  3  I"  +  3  H+ 
4  I2  +  2  R— SH  +  3  H2O  ->  R— S— S— R  +  IO3-  +  7  I-  +  8  H+ 

However,  it  is  likely  that  near  neutrality  the  first  reaction  is  dominant.  In 
the  oxidation  of  cysteine,  3  equivalents  of  iodine 'are  taken  up  to  form 
cysteate: 

3  I2  +  R— SH  +  3  H2O  -►  R— SO3-  +  6  I-  +  7  H+ 

It  is  interesting  that,  at  pH  3.2,  iodine  oxidizes  cysteine  well,  but  does  not 
react  with  cystine,  tyrosine,  or  histidine.  This  indicates  that  the  first  prod- 
uct in  the  oxidation  of  cysteine  is  not  cystine  but  free  radicals,  which  can 
either  combine  to  form  disulfides  or  be  further  oxidized  to  sulfonate  groups 
(Anson,  1940).  At  pH  6.8,  cystine  is  the  major  product.  In  most  instances, 
especially  with  proteins,  several  reactions  will  occur  and  mixed  products 
will  be  found.  In  addition  to  these  straightforward  oxidations,  we  shall  see 
that  there  is  now  evidence  for  the  formation  of  sulfenyl  iodide  groups  (SI), 
so  that  a  certain  fraction  of  the  iodine  can  be  incorporated  into  the  thiol 
molecule. 

Reactions  of  Iodine  with   Proteins 

The  SH  groups  of  denatured  ovalbumin  are  oxidized  by  iodine  within  5 
min  at  pH  3.2  and  37°  (Anson,  1940).  The  rate  of  the  reaction  decreases  as 
the  pH  is  raised  to  6.8.  In  acid  media  iodine  does  not  react  with  tyrosine  or 
proteins  containing  tyrosine  (e.g.  pepsin),  whereas  at  neutrality  it  readily 
iodinates  tyrosine.  By  proper  choice  of  pH  and  iodine  concentration  it  is 
possible  to  oxidize  the  SH  groups  of  denatured  ovalbumin  without  appre- 


IODINE  681 

ciably  altering  tyrosine  residues  (Anson,  1941).  Iodine  reacts  only  with  the 
tyrosine  residues  of  seralbumin,  and  denaturation  accelerates  the  formation 
of  diiodotyrosine  (Li,  1945).  The  rate  of  the  reaction  is,  however,  quite  slow 
in  native  seralbumin  (half-reaction  time  around  100  min).  Human  seral- 
bumin iodinated  at  low  temperature  takes  up  36  atoms  of  iodine  per  mole 
of  albumin,  but  only  12  diiodotyrosyl  groups  are  found  (Hughes  and  Straes- 
sle,  1950).  The  remainder  was  believed  to  be  incorporated  into  histidyl  re- 
sidues. Some  oxidation  of  cysteinyl  residues  also  occurs  and  this  presumably 
is  beyond  the  disulfide  stage,  since  2.2  moles  of  iodine  are  taken  up  per  SH 
group.  In  any  particular  case,  the  amount  of  disulfide  formed  will  depend 
to  a  large  extent  on  steric  factors,  i.e.,  how  readily  the  sulfhydryl  radicals 
can  combine;  the  seralbumin  molecule  is  fairly  large  and,  not  surprisingly, 
disulfide  groups  are  not  found  after  oxidation.  Although  no  degradative 
changes  in  seralbumin  are  observed,  protein  structure  is  certainly  modified 
by  treatment  with  iodine,  since  the  water  binding  capacity  is  increased 
(Jensen  et  at.,  1950)  and  the  rates  of  pepsin  and  trypsin  digestion  are  de- 
creased (Raghupathy  et  al.,  1958). 

We  have  noted  that  2  to  3  atoms  of  iodine  are  occasionally  utilized  for 
each  protein  SH  group.  This  might  indicate  (1)  oxidation  of  SH  beyond 
the  disulfide  state,  (2)  reduction  of  the  iodine  beyond  the  iodide  state,  or 
(3)  some  substitution  of  iodine  in  the  cysteinyl  residue.  This  problem  was 
studied  by  Fraenkel-Conrat  (1955)  with  tobacco  mosaic  virus  protein.  It 
is  possible  that  sulfenate  or  sulfenyl  iodide  groups  might  be  produced,  but 
it  has  always  been  thought  that  such  groups  are  quite  unstable  and  cannot 
exist  for  appreciable  time.  However,  the  virus  protein  SH  groups  react  with 
2  atoms  of  iodine  fairly  rapidly,  and  this  was  shown  to  be  accompanied  by 
the  formation  of  sulfenyl  iodide  groups: 

V— SH  +  I2  ->  V— SI  +  I-  +  H+ 

This  group  appears  to  be  stable  in  this  particular  protein.  Fraenkel-Conrat 
pointed  out  that  further  reaction  with  thiols  can  form  mixed  disulfides: 

V— SH  +  R— SH  ±^  V— S— S— R  +  H+  -f-  I" 

Such  reactions  have  been  studied  further  in  /5-lactoglobulin  by  Cunningham 
and  Nuenke  (1959,  1960,  1961),  using  a  spectrophotometric  method.  This 
protein  reacts  with  4  equivalents  of  iodine  to  form  2  sulfenyl  iodide  groups 
per  mole  of  protein: 

P(— SH),  +  2  I,  ->  P(— SI),  +  2  I-  +  2  H+ 

Ovalbumin  reacts  similarly  but  6  equivalents  of  iodine  are  taken  up.  The 
sulfenyl  iodide  groups  are  quite  stable  in  these  proteins,  but  can  react  with 
simple  thiols  (e.g.,  glutathione,  cysteine,  and  others)  to  form  mixed  disul- 


682  5.    OXIDANTS 

fides.  Intermolecular  disulfide  formation  was  ruled  out  for  these  proteins. 
Ovalbumin  has  4  SH  groups  and  2  S — S  groups;  the  protein  treated  with 
iodine  has  2  SH  groups  and  2  S — S  groups,  and  has  incorporated  1  iodine 
atom  (Winzor  and  Creeth,  1962).  Since  5  atoms  of  iodine  are  taken  up,  it 
is  not  a  simple  oxidation  to  disulfide.  It  was  suggested  that  the  following 
reactions  occur: 

2  P(— SH)2 

HS-P— S— S-P-SH 

IS-P-S-S-P— SI 

+  2  I2  +  3  HjO 
IS-P— SO      +     bzS-P-SI 

where  P  represents  that  portion  of  the  protein  not  reacting  with  iodine. 
Further  oxidation  of  the  sulfenate  group  to  sulfinate  may  occur  to  give  a 
homogeneous  product.  Therefore  the  formation  of  sulfenyl  iodide  groups 
and  oxidation  of  SH  groups  to  sulfenate  and  sulfinate  must  be  considered 
as  likely  possibilities  in  enzymes  treated  with  iodine. 

Inhibition  of  Enzymes 

Many  enzymes  have  been  found  to  be  readily  inhibited,  often  by  low  con- 
centrations of  iodine  (Table  5-4).  It  is  impossible  to  know  in  most  cases 
whether  the  inhibition  is  due  to  reaction  with  SH  groups  or  to  iodination 
of  tyrosine.  The  fact  that  most  studies  have  been  done  at  pH's  around 
neutrality  implies  that  both  SH  and  tyrosyl  groups  could  be  reacted,  so 
that  the  relative  importance  would  depend  on  the  accessibility  of  the  groups 
and  their  location  with  respect  to  the  active  center.  Fixation  of  iodine  into 
an  enzyme  does  not  imply  inhibition;  an  example  is  Aspergillus  protease 
(Dhar  and  Bose,  1962).  The  inhibition  of  certain  enzymes  by  iodine  is 
probably  related  to  oxidation  of  SH  groups:  papain,  creatine  kinase,  urease, 
aldolase,  lactate  dehydrogenase,  succinate  dehydrogenase,  pyruvate  decar- 
boxylase, and  adenosinetriphosphatase.  Other  enzymes,  such  as  pepsin  or 
peroxidase,  are  inhibited  through  tyrosine  iodination,  and  in  some  instances 
a  mixed  mechanism  is  probable. 

One  way  of  determining  if  SH  group  oxidation  is  responsible  for  enzyme 
inhibition  is  to  attempt  reversal  with  thiols.  Complete  reversal  certainly 
implies  such  a  mechanism,  but  negative  results  can  be  interpreted  in  various 
ways.  Even  oxidation  to  disulfide  groups  is  not  necessarily  reversed  by 
thiols  if  steric  factors  prevent  reaction,  and  oxidation  past  the  disulfide 


IODINE  683 

stage  would  not  be  expected  to  be  reversed.  No  reactivation  of  /3-galactosi- 
dase  (Knopfmacher  and  Salle,  1941 )  or  a-amylase  (Di  Carlo  and  Redfern, 
1947)  is  observed;  however,  in  both  cases  there  is  some  reason  for  believing 
that  SH  groups  are  involved.  Partial  reactivation  of  /5-amylase  (Weill  and 
Caldwell,  1945  b)  and  phosphoglyceraldehyde  dehydrogenase  (Rapkine, 
1938)  was  taken  to  mean  that  at  least  SH  group  oxidation  is  responsible 
for  the  inhibition.  Essentially  complete  reactivation  with  thiols  has  been 
found  for  urease  (Hellerman,  1939),  papain  (Hellerman  and  Perkins,  1934), 
cholinesterase  (Nachmansohn  and  Lederer,  1939),  and  lactate  dehydrogen- 
ase (Nygaard,  1955)  so  that  an  SH  mechanism  seems  assured  for  these. 
The  mechanism  of  the  inhibition  of  /5-fructofuranosidase  is  still  unknown, 
although  it  is  the  first  enzyme  studied  with  iodine.  The  enzyme  is  inhibit- 
ed fairly  rapidly  to  the  extent  of  45  50%,  but  further  inactivation  proceeds 
very  slowly  (Myrback,  1926).  A  "Jodsaccharase"  was  assumed,  but  the 
iodine  must  not  be  fixed  at  the  active  center  since  there  is  no  decrease  in 
the  affinity  for  the  substrate.  There  is  no  reactivation  by  reduction  (Myr- 
back, 1957  a)  so  there  is  little  evidence  for  SH  group  oxidation.  Sulfenyl 
iodide  groups  may  be  involved. 

There  has  been  little  study  of  the  disappearance  of  SH  groups  during  in- 
hibition by  iodine.  Cardiac  lactate  dehydrogenase  SH  groups  are  rapidly 
oxidized  by  iodine,  as  determined  with  Ag+  and  spectrophotometrically  with 
p-mercuribenzoate,  and  the  inhibition  develops  in  parallel  fashion  (Nygaard, 
1956).  The  lactate  dehydrogenase  from  rabbit  muscle,  on  the  other  hand, 
incorporates  iodine  at  0^  and  pH  8  over  many  hours;  when  1  atom  of  iodine 
is  incorporated  per  molecule  of  enzyme,  the  inhibition  is  30%,  and  the  in- 
hibition increases  until  21  atoms  of  iodine  are  incorporated.  Both  NAD  and 
oxalate  protect  the  enzyme  against  iodination.  Although  the  results  with 
iodoacetamide  indicate  an  SH  group  at  the  active  site,  one  cannot  be  cer- 
tain if  this  is  the  initial  point  of  attack  for  the  inhibition  (Dube  et  al.,  1963). 
It  is  likely  in  situations  like  this  that  both  SH  group  oxidation  and  iodina- 
tion of  tyrosine  occur. 

Pepsin  is  not  an  SH  enzyme  but  is  inhibited  by  iodine,  and  here  it  is 
highly  probable  that  tyrosine  iodination  occurs.  The  activity  of  pepsin  de- 
creases with  the  amount  of  iodine  incorporated;  it  is  inactive  when  35-40 
atoms  of  iodine  are  bound  (Herriott,  1937).  3-Iodotyrosine  has  been  isolat- 
ed from  inhibited  pepsin  (Herriott,  1947),  confirming  the  importance  of 
tyrosine  for  the  enzyme  activity.  Of  the  6  tyrosyl  groups  in  ribonuclease, 
3  are  unreactive,  and  the  problem  of  where  these  are  in  the  polypeptide 
chain  was  studied  with  iodine  (Cha  and  Scheraga,  1961  a,b).  At  pH  9.4  and 
10°  —  conditions  favorable  for  tyrosine  iodination  with  minimal  effects  on 
other  groups  —  3  tyrosyl  residues  are  iodinated;  the  others  can  be  iodinated 
only  very  slowly.  The  iodinated  tyrosyl  residues  were  located  in  the  amino 
acid  sequence.  Such  techniques  will  undoubtedly  become  more  common 


684 


5.    OXIDANTS 


tf 


1^ 

o    a 


o 

05 

a 

3 

>> 

Pm 

0) 

P4 

eS 

^-^ 

^3 

-* 

C 

O 

-* 

c^ 

-»^ 

05 

O 

t>i 

g 

^-^ 

Ji 

IS 

to 

"S 

03 

N 

m 

^     ci 


:^  3 


■* 

J4 

05 

_o 

^H 

^.-^ 

r— 1 

§ 

1 

05 

^ 

2 

"e 

-0 

a 

d 

cS 

-s 

^ 

CI 

c 

S 

o 

X! 

1 

to 
bC 

O 

Q 

1 

05     Ph        125 


Ph 


oq 


>, 

to. 

•| 

Sh 

C 

-T^ 

ert 

fi.^ 

M 

'^ 

-^ 

"w" 

eg 

<?-. 

"H. 

-C 

ro 

D 

EC 

&< 

cd 

a 

e 

O 

oa 
O 

"ft 

W 

lO 

.4^ 

i 

« 

« 

OJ 

ce 

05 

"m 

+^ 

'm 

;h 

o 

o 

o 

4J 

c 

S 

0) 

T3 

S 

03 

'C 

(H 

T3 

Q^ 

<: 

<1 

<t; 

<; 

^ 


73  O 


IODINE  685 


;^  CO       rL      X;   ^ 


M 

~t^ 

t^ 

2 

Eh 

s 

^ 

» 

<D 

o 

^3 

^ 

s 

cS 

'^ 

cS 

c 

'5' 

13 

g 

H-5 

-tJ 

> 

■* 

W 

t3 

"^ 

^ 

P2 


o 


«     ^  OS 


^ 

0 

r~^ 

OS 

c8 

c 

-H 

g 

^.-^ 

c 

<ij 

t^ 

a 

« 

0 

0 

N 

c 

c3 

CO 

ui 

C5 

LO 

0 

10 

0 

0 

0 

0  0 

r> 

00 

IC 

0 

« 

0 

LO 

0 

O       -H  ^ 


bo 

■1        ^ 

00 

0         Tt 

»o 

CO 

0 

0 

s    r^  t-    -H 

10 

0  10 

i       d 

d   d 

d 

d 

d 

§   d  d   d 
d 

d  d 

<»    d, 


ft      j2 


ft  ft 


CUM  ci2&i|il(iHMQ>^>'>^!>Hf*5  a,f^ 


Is 

_       _  _  O 

Mc3ooo  00      6'^  «a 


1^ 
0 

ft 

'J2 

r- 

? 

C 

Oj 

0 

c 

0 

"o 

0 

>, 

a 

0 

-C 

-C 

>1 

0 

0 

0 

0 

rt 

^ 

ft 

0 

0) 

,  <^ 

en 

ft 

as 

0 

C 

« 

be 

0 

n 

t^, 

t4 

ts 

0 

>. 

■as. 

X 

686 


5.    OXIDANTS 


P5 


Oh 

xn 


^ 


C5      CI  '-^  31      C5 


o  ^  ^-  o  o 

5  =*  i£  S  ?; 

o  :3  3  o  o 

"o  ^  >  "o  "o 

M  ^  l>  M  02 


pq 


P4  O      O 


-2 

oil 

2 

T3 

C 

> 

a 

>. 
"S 

>, 

-o 

a 

s 

» 

S- 

o 

-k^ 

» 

-+:» 

cS 

tc 

"o 

p 

-i-i 

03 

73 
1— 1 

i? 

3 

>> 
^ 


c    _C 


c 

S 

c3 

u 

•^_-' 

r^ 

<o 

S 

Ph 

"e 

o 

-73 

"S 

hJ 

c3 

r. 

'^ 

C 

O 

s 

c3 

s 

U 

c 

l-l 

<o 

-2 

J=> 

£ 

o 

« 

"O) 

A 

m 

ffi 

(M 

cc 

CO      -H 

ii  II 

o 

_, 

d 

<M    lO 

W  W 

d 

d  — 1  o 

O        Ph    Pm 


pq 


o 


Ph 


PM 


a 


eu 


IODINE  687 


rS         —  r^  -A  C3  CD  "J 

^        ci       -r^  ^  CO  B  rt-H 

Oi      2      ti        ^  i-o        .S        .52        -^       ^ 

?5  2  2  2  <^r^S4^^cQ;5i3 

2  ^         —         ^  w^'^m.S'^^^!^ 

«  .2  .2  .2  c      cs     ^         g  'c         fe         =«         S         c 

P5  Ph  PQ  M  Qt^Wfes^MOPHOO 


05  ^^rt  C5000 

OfCO         O-^         O^         C^  0      0-*         ooo 


CO        Co      ^  O  4)  O 

"^      "^      O  >H  >^  W 


>> 

ci 

M 

<U 

O 

&£ 

-C 

O 

o. 

Ch 

03 

73 

o 

>■. 

^ 

j2 

Ph 

73 

,5 
o 


e8 

fl 

(C 

tiO 

0 

S 

0) 

1 

r2 

© 

'S 

n3 

0 

688  5.    OXIDANTS 

when  more  enzymes  are  susceptible  to  sequential  analysis.  Another  interest- 
ing approach  to  elucidating  enzyme  binding  groups  with  iodine  is  illustrat- 
ed by  the  study  of  the  old  yellow  enzyme  by  Theorell  (1956).  Flavin  is 
bound  to  the  apoenzyme  through  its  imino  group  and  fluorescence  is  quen- 
ched; Weber  had  suggested  that  a  tyrosine  hydroxyl  group  might  bind  this 
imino  group.  This  was  examined  by  reaction  of  the  apoenzyme  with  iodine; 
since  no  SH  groups  are  present,  this  is  relatively  easy.  It  was  found  that 
very  low  concentrations  of  iodine  decrease  the  coupling  rate  of  FMN  to  the 
apoenzyme,  and  90%  of  the  iodine  which  disappears  is  recovered  as  diiodo- 
tyrosine. 

Some  results  on  the  variation  of  inhibition  with  pH  appear  to  point  to 
the  importance  of  tyrosine  iodination.  The  iodination  of  pepsin  is  very  slow 
below  pH  4.5  and  rises  suddenly  as  the  pH  is  increased  to  become  maximal 
around  pH  5.5  (Herriott,  1937).  This  is  essentially  the  same  pH  dependence 
as  found  for  glycyltyrosine.  The  inhibition  of  /3-fructofuranosidase  by  iodine 
is  minimal  at  pH  5.14  and  much  faster  at  pH's  above  6  (Myrback,  1926), 
which  might  support  the  importance  of  tyrosine  iodination  in  the  inhibition. 
Although  one  might  expect  the  effect  of  iodine  on  cathepsins  to  be  mainly 
through  reaction  with  SH  groups,  Maver  and  Thompson  (1946)  found 
greater  inhibition  by  0.25  mM  iodine  at  pH  7  (71%)  than  at  pH  3.5  (20%). 
Results  with  various  SH  reagents  do  not  favor  the  importance  of  SH 
groups.  This  may  well  be  a  case  where  there  is  a  mixed  mechanism  for  the 
inhibition. 

An  interesting  situation  occurs  in  the  reaction  of  the  exopenicillinase  of 
B.  cereus  with  iodine  (Citri  and  Garber,  1960,  1961).  This  enzyme  can  exist 
in  two  antigenically  different  states  —  a  and  y  —  and  these  differ  in  re- 
sponse to  iodine,  although  the  enzyme  activity  is  the  same  for  both.  a-Pen- 
icillinase  is  quite  resistant  to  iodine  whereas  ^-penicillinase  is  inhibited 
by  0.5-1  mM  iodine.  The  enzyme  must  be  flexible  since  in  the  presence  of 
the  competitive  inhibitor,  6-(2,6-dimethoxybenzamido)penicillanic  acid,  it 
becomes  sensitive  to  iodine.  Pretreatment  with  this  inhibitor,  followed  by 
its  removal,  does  not  alter  hydrolysis  of  benzylpenicillin,  so  that  any  struc- 
tural change  that  occurs  is  not  permanent.  It  is  very  difficult  to  understand 
how  a  rather  large  competitive  inhibitor,  which  must  cover  the  active  cen- 
ter, could  allow  reaction  of  any  group  at  the  active  center  with  iodine,  un- 
less the  reacted  group  is  just  vicinal  to  the  active  center  and  either  SH  oxi- 
dation or  iodination  alters  the  structure. 

Effects  on  Cellular  Metabolism  and  Function 

The  uncoupling  of  oxidative  phosphorylation  by  iodine  has  been  claimed 
to  relate  to  the  effects  of  thyroxine.  Klemperer  (1955)  reported  that  al- 
though iodide  exhibits  no  uncoupling  activity,  iodine  is  quite  effective  in 
rat  liver  mitochondria  with  /5-hydroxybutyrate  as  substrate  (see  accom- 


IODINE  689 

panying  tabulation).  Iodine  appears  to  fulfill  the  requirements  of  an  un- 
coupler,  in  that  it  can  reduce  the  P:0  ratio  significantly  without  depressing 
the  respiration,  although,  to  completely  uncouple,  the  O2  uptake  must  be 


KI 

(mM) 

I2 

{mM) 

0, 

;  uptake 

P:0 

0 

0 

2.3 

1.7 

20 

0 

2.1 

1.8 

20 

0.02 

2.3 

1.5 

8 

0 

2.0 

1.9 

8 

0.05 

2.1 

1.0 

8 

0.2 

0.8 

0 

inhibited.  Middlebrook  and  Szent-Gyorgyi  (1955)  found  an  uncoupling  in 
mitochondria  when  Cl~  is  partially  replaced  with  I^;  at  25  mM  I~,  phos- 
phorylation is  almost  abolished  without  depression  of  respiration.  It  is  not 
known  if  this  is  due  to  I"  itself,  to  reduction  in  CI",  or  to  iodine  formed 
from  I~.  Iodine  causes  mitochondrial  swelling  at  a  concentration  as  low  as 
0.005  raM,  and  in  this  it  resembles  thyroxine  (Rail  et  al.,  1962).  Iodine  is 
able  to  oxidize  NADH  but  addition  of  NADH  does  not  reverse  the  swelling. 
Other  oxidizing  agents  do  not  duplicate  this  effect.  Furthermore,  agents  that 
inhibit  thyroxine-induced  swelling  also  inhibit  that  caused  by  iodine.  De- 
spite the  superficial  similarities  in  the  actions  of  iodine  and  thyroxine,  it  is 
difiicult  to  understand  the  nature  of  any  relationship.  It  is  very  unlikely 
that  thyroxine  releases  its  iodine,  and  since  thyroxine  is  always  more  potent 
than  iodine,  not  enough  iodine  could  be  released  in  any  event.  It  is  possible 
that  iodine  does  not  act  directly,  but  iodinates  tyrosine  or  some  protein, 
and  that  this  product  is  the  active  uncoupler. 

One  might  expect  iodine  to  be  an  effective  inhibitor  of  glycolysis,  inas- 
much as  this  pathway  involves  a  number  of  SH-dependent  steps.  Yeast 
fermentation  is  indeed  quite  sensitive  to  iodine,  0.017  mM  inhibiting  22% 
and  0.085  roM  inhibiting  100%  (Schroeder  et  al,  1933  b).  There  is  simul- 
taneously a  loss  of  GSH  and  probably  other  SH  groups.  This  does  not 
prove  that  the  glycolytic  inhibition  is  related  to  SH  groups,  but  is  sugges- 
tive. The  locus  of  action  is  not  known;  one  thinks  of  aldolase,  because  of 
its  great  sensitivity  to  iodine,  but  most  of  the  glycolytic  enzymes  have  not 
been  tested.  That  iodine  can  oxidize  SH  groups  in  cells  was  shown  by  Ca- 
fruny  et  al.  (1955  a).  Kidney  sections  incubated  with  iodine  exhibit  85% 
loss  in  SH  groups  in  the  proximal  and  distal  tubules.  However,  so  little 
work  has  been  done  on  cell  metabolism  with  iodine  that  it  is  impossible  to 
predict  if  any  pathways  are  inhibited  selectively;  it  would  appear  to  be 
unlikely,  unless  glycolysis  proves  to  be  more  susceptible  than  other  systems. 


690  5.    OXIDANTS 

Iodine  can  interfere  with  the  transport  of  substances  across  cell  mem- 
branes. Hemolysis  by  glycerol  and  other  nonelectrolytes  is  quite  potently 
inhibited  by  iodine  at  0.08  mM  (LeFevre,  1947,  1948)  as  it  is  by  various 
SH  reagents.  The  effects  are  readily  reversed  by  thiols.  This  was  taken  to 
indicate  that  SH  groups  are  in  some  manner  involved  in  the  transport  of 
these  substances  into  the  erythrocytes.  However,  it  is  not  necessarily  evi- 
dence for  an  active  transport,  since  membrane  permeability  could  be  af- 
fected directly  or  indirectly.  Iodine  also  inhibits  the  transport  of  phosphate 
into  staphylococci  and  it  was  claimed  that  this  process  involves  SH  groups 
(Mitchell,  1954).  Finally,  iodine  at  0.1  mM  reduces  the  short-circuit  current 
and  electrical  potential  of  frog  skin  (Eubank  et  al.,  1962),  but  there  is  no 
evidence  as  to  the  site  or  mechanism  of  this  action. 

Iodine  has  been  studied  a  great  deal  in  connection  with  its  germicidal 
activity  (Gershenfeld  and  Witlin,  1950)  but  not  a  great  deal  has  been  done 
from  the  metabolic  standpoint.  The  effects  of  pH  on  the  ability  of  iodine 
to  kill  bacteria,  fungi,  or  spores  are,  however,  of  interest,  since  they  would 
presumably  apply  to  work  with  any  cells.  It  has  generally  been  considered 
that  at  lower  pH's  there  is  more  free  iodine,  and  hence  greater  penetrability 
into  cells  and  greater  activity.  It  is  true  that  more  iodate  would  be  formed 
in  alkaline  solutions  and,  in  the  absence  of  much  iodide,  more  hypoiodous 
acid.  Wyss  and  Strandskov  (1945)  found  the  bactericidal  activity  to  de- 
crease at  higher  pH's  and  attributed  this  to  a  greater  formation  of  HOI 
and  lOg".  When  iodide  is  present,  the  formation  of  HOI  is  suppressed,  and 
the  pH  does  not  affect  the  activity.  It  was  also  observed,  as  would  be  ex- 
pected, that  the  action  of  iodine  is  strongly  dependent  on  temperature,  re- 
quiring about  4  times  as  long  to  kill  Bacillus  metiens  spores  for  each  10° 
drop  in  temperature. 

PEROXIDES 

Hydrogen  peroxide  and  other  peroxides  occasionally  depress  enzymes  and 
metabolism  potently  but  little  is  known  about  the  specificity  with  respect 
to  SH  groups.  In  comparison  with  other  oxidants,  no  thorough  studies  of 
the  effects  of  hydrogen  peroxide  on  proteins  have  been  made.  Mirsky  and 
Anson  (1935)  mention  that  hydrogen  peroxide  is  convenient  to  use  in  the 
oxidation  of  SH  groups,  but  it  has  never  been  widely  applied  for  this  pur- 
pose. The  interesting  effects  of  hydrogen  peroxide  on  glycolysis  and  a  few 
enzymes  justify  a  brief  discussion. 

Chemistry 

Hydrogen  peroxide  is  a  nonlinear  molecule  that  is  quite  miscible  with 
water: 

H2O2  +  H2O  -?  H3O+  +  OOH-  K  =  2Ax  10-1'' 


PEROXIDES  691 

The  ion  product  (H+)  (00H-)  is  1.55  X  10-^2  at  20°.  Thus  it  is  a  very  weak 
acid  and  the  ion  OOH"  is  probably  unimportant  in  its  reactions.  Hydrogen 
peroxide  can  function  as  both  oxidant  and  reductant.  It  is  a  strong  oxidiz- 
ing agent  in  both  acid  and  alkaline  media,  but  a  relatively  poor  reductant. 
Although  the  oxidation-reduction  potential  would  be  more  favorable  for 
oxidation  in  acid  medium,  the  rate  of  oxidation  is  often  greater  in  alkaline 
conditions.  Hydrogen  peroxide,  of  course,  is  an  unstable  substance,  especi- 
ally in  the  presence  of  organic  material,  and  this  must  be  considered  in  its 
use.  Despite  its  instability  ( — AF  =  23.4  kcal/mole),  it  is  rather  stable  in 
pure  solution,  but  its  decomposition  is  catalyzed  by  heavy  metal  ions,  and 
is  more  rapid  in  alkaline  than  acid  media. 

Inhibition  of  Enzymes 

The  few  results  summarized  in  Table  5-5  are  not  comparable  with  each 
other  because  the  conditions  were  quite  different  in  the  various  studies. 
However,  there  is  no  doubt  that  some  enzymes  are  very  sensitive  to  hydro- 
gen peroxide.  The  inhibition  develops  very  slowly  in  some  cases;  with  yeast 
/5-fructofuranosidase  the  inhibition  by  2.9  M  hydrogen  peroxide  is  0%  at  1 
min,  10%  at  30  min,  47%  at  3  hr,  and  100%  at  21  hr  (Myrback,  1957  b). 
Of  course,  at  this  very  high  concentration  one  has  no  idea  of  the  mechanism 
of  the  inhibition,  and  can  only  marvel  at  the  resistance  of  this  enzyme.  The 
inhibition  of  ATPase  depends  on  the  pH  at  which  the  reaction  is  run:  Thus 
the  enzyme  was  incubated  with  hydrogen  peroxide  at  pH  7  for  15  min,  and 
the  inhibition  was  found  to  be  51%  when  the  ATPase  reaction  was  tested 
at  pH  6.3  and  95%  when  tested  at  9.2  (Mehl,  1944).  The  reason  for  this 
strange  behavior  is  unknown.  One  factor  that  has  not  been  generally  consid- 
ered is  the  possible  presence  of  heavy  metal  ions  in  the  hydrogen  peroxide. 
Holmberg  (1939)  believed  that  the  inhibition  of  uricase  he  observed  was 
due  to  traces  of  Cu++,  inasmuch  as  diethyldithiocarbamate  prevents  the 
inhibition.  It  is  also  possible  that  some  metal  ion  may  be  necessary  to 
catalyze  the  oxidation  of  the  enzyme  and  that  the  inhibition  is  not  due 
to  the  Cu++  itself. 

The  inhibition  of  /5-galactosidase  by  hydrogen  peroxide  is  completely  re- 
versible by  HgS  or  cyanide,  while  that  by  iodine  is  not,  indicating  that  here 
one  may  oxidize  the  SH  groups  more  specifically  with  the  peroxide  (Knopf- 
macher  and  SaUe,  1941).  Reactivation  of  ATPase  inhibited  by  hydrogen 
peroxide  was  observed  by  both  Mehl  (1944)  and  Ziff  (1944),  using  cysteine 
or  glutathione,  so  that  specific  oxidation  of  SH  groups  may  occur  with  this 
enzjTne.  Simultaneously  there  is  a  suppression  of  the  interaction  of  actin 
and  myosin,  which  is  believed  to  depend  on  SH  groups  (Bailey  and  Perry, 
1947).  Papain  inhibited  up  to  90%  by  hydrogen  peroxide  can  also  be  reac- 
tivated by  cysteine,  but  beyond  this  there  is  apparently  oxidation  beyond 
the  disulfide  stage  (Sanner  and  Pihl,  1963).  Blocking  the  SH  groups  with 


692 


5.    OXIDANTS 


P5 


m 


5    S   -i3 

PQ  <5  S 


PM 

Ni  pq 


i-:i  W 


c4 

Q 

T3 

„.— V 

C 

oo 

cS 

lO 

Oi 

c8 

a 

o 

CO 

eS 

05 

^ 

>> 

l-H 

0) 

S 

o 

-ii 

~t^ 

^ 

O 

-fc3 

o 

H 

O 

^~.  "S    <M    -C 


ft 

o 

c 


g      ft     P<     <»      03 


'^      ^      ^      ^ 


a  a 


.-    -»^    -»^ 


03    ^    +a     -P    -J2    X! 
QD    •«*      c3      c6      c5      c5 

>H  u  p^  (25  tf  pq 


m     oo 

a  a 

.1^  .-s  -►^ 


-  i  I 

I  "§  -S 

5=      o       cS 


jj^     '^     lis  .3     ™ 


•a 


c3 

J3 

ft 

a: 

O 

^ 

ft 

IB 

'ti 

GQ 

-1-3 

CS 

0) 

.g 

,g 

'3 

o 

o 

c 

c 

(0 

« 

T3 

-0 

<! 

<i3 

^g   CO 


CD 

'« 

OJ) 

-3 

T) 

O 

to 

C 

-2 

c« 
C 

S      S 

g3 
C 

c3 

,3    'to 

03 
03 

-5     Is 

>>  x> 

'5b 

1 

o  -is 

'o 

C     O 

rt 

x> 

iH      cs 

73 

73 

1      <K 

ft 
to 

o3 

^  ^5 

<i 

QQ.  <3 

^Jl 

o 

<*i.  CQ. 

PEROXIDES 


693 


1—t 

o 

CD 

CO 

CO 

05 

05 

05 

1—1 

-U 

-U 

'-- 

4) 

c 

c 

05 

M 

<B 

4) 

C 

1 

02 

CO 

Ci 

C 

l> 

t3 

2 

-C 

b 

-tJ 

-li 

->J 

o 

_fi 

C 

a 

^ 

T3 

2 

t 

£ 

>> 

£ 

'o 

C 

O 

Q 

CO 

o 

Q 

^H        (D        TO 


2    ^ 


c    c  ZS 


T3     O     (» 


O 

C 


~       05 

o 


o    p    a>  -C 

P  fe  m  p 


PQ 


»o  !2  05 
ic  ®  ^ 

05     — I     — ' 


.£j    ft        ^ 


;ir     f^ 


5  1^  s 


c 

C 

fl 

C 

© 

O 

2 

2 

"o 

"o 

K 

W 

ti^ 


^_^ 

^ 

c 

o 

s 

o 

->* 

o 

05 
CO 

5 

05 

TS 

"-g 

'O 

c 

T\         e        ^  c 


OOOCO^Ct^t^t-- 
-H     -^     Cq     O     05     -< 


OOOOOiC-hQOOJSOOOOCOOO 


'-3   "^   '-3 

CO    CO    CO 


O      -H      O      -H 


CO    CO    CD 

O    -H    CD 
O    O    O    CO 


CD 


o  o  o 


O    <M    >0    (M 


O    '^    s^    ■<* 


o  o  -< 

o  o  o 

O  O  O    -H 

■rit  <6  <6  'Z'  <6  -^ 


s   s 

Si    5!) 


o 

S 

o 

•<?a 

o 

^ 

o 

%^ 

E-1    5^    "^    tf    >H    CQ 


QQ    5-1 


e 

«j 

?!, 

00 

-S 

s 

eg 

•-s. 

s 

>i 

S' 

•^ 

03 

!U 

^ 

a 

?5, 

^::? 

c6 

«o 

> 

e^ 

^ 

^ 

fl      o3 

fin  PL, 


W 


p^  pq  !S 


>. 

^ 


^  Ph   |, 

CO    CO    "g 


o    S    >< 

O    T3 


o  o  w  ^ 


PL,     PlH 


(M      Lh      Co 
ID    D    >^ 


694  5.    OXIDANTS 

p-mercuribenzoate  prevents  their  reaction  with  hydrogen  peroxide.  The 
single  SH  group  of  papain  is  about  5  times  as  susceptible  to  oxidation  by 
hydrogen  peroxide  as  the  SH  group  of  phosphoglyceraldehyde  dehydrogen- 
ase. a-Chymotrypsin  has  two  methionine  residues,  one  being  3  residues 
away  from  the  active  site  serine  and  the  other  15  residues  removed.  The 
Met-3  is  oxidized  by  hydrogen  peroxide  specifically,  whereas  both  SH  groups 
are  oxidized  in  the  urea-denatured  enzyme  (Schachter  et  at.,  1963).  It  is 
obvious  that  disulfide  bonds  cannot  be  formed  intramolecularly  in  the  na- 
tive enzyme  and  it  was  shown  that  methionine  sulfoxide  is  the  product. 
Glutathione  is  also  oxidized  to  the  sulfoxide  by  hydrogen  peroxide  (Utzin- 
ger  et  al.,  1963). 

Hydrogen  peroxide  generated  during  oxidation  in  enzyme  preparations 
or  cells  is  sometimes  inhibitory  and  for  many  years  it  has  been  assumed 
that  at  least  one  function  of  catalase  is  to  protect  cells  against  it.  Dixon 
(1925)  was  the  first  to  demonstrate  this  with  a  purified  enzyme  system. 
When  purines  are  oxidized  by  oxygen  in  the  presence  of  xanthine  oxidase 
there  is  a  progressive  inactivation  of  the  enzyme,  which  is  due  to  hydrogen 
peroxide  formed,  since  it  can  be  prevented  by  catalase.  Hydrogen  peroxide 
was  claimed  to  stimulate  xanthine  oxidase  at  very  low  concentrations  (be- 
tween 0.00001  mM  and  0.01  mM)  (Table  5-5),  and  to  inhibit  above  0.1  mM 
(Bernheim  and  Dixon,  1928).  The  rate  of  activation  is  slow,  maximal  effects 
of  0.001  mM  hydrogen  peroxide  occurring  in  around  100  min.  The  mecha- 
nism for  this  is  unknown;  metal  impurities  seem  unlikely  because  they  would 
be  at  extremely  low  concentrations.  The  inhibition  of  Aspergillus  aconitase 
depends  on  the  strain  of  the  organism,  the  sensitivity  varying  over  at  least 
a  4-fold  range  of  concentration  (Bruchmann,  1961  a).  Certain  strains  tend 
to  accumulate  citrate  under  specified  conditions  and  this  is  believed  to  be 
due  to  the  hydrogen  peroxide  formed  and  the  particular  sensitivity  of  aco- 
nitase in  these  strains  (Bruchmann,  1961  b).  Adding  exogenous  hydrogen 
peroxide  augments  the  accumulation  of  citrate  (Bruchmann,  1961  c). 

Succinyl  peroxide  is  a  radiomimetic  substance  and  has  been  found  to 
inhibit  several  SH  enzymes  while  having  much  less  action  on  non-SH  en- 
zymes (Wills,  1959).  The  enzymes  inhibited  are  amylase,  /5-fructofuranosi- 
dase,  urease,  succinate  dehydrogenase,  phosphoglyceraldehyde  dehydrogen- 
ase, papain,  tyrosinase,  and  cholinesterase.  Urease,  for  example,  is  inhib- 
ited completely  in  8  min  by  0.033  mM  succinyl  peroxide.  Reversal  of  inhi- 
bitions by  cysteine  is  obtained  only  if  the  period  of  exposure  to  the  peroxide 
is  brief.  It  is  questionable  if  the  action  is  exerted  by  succinyl  peroxide  itself, 
since  it  is  immediately  hydrolyzed  to  persuccinate  in  aqueous  solution,  and 
peracids  have  long  been  known  to  oxidize  SH  groups  (Freudenberg  and 
Eyer,  1932;  Swan,  1959).  The  inhibition  of  catalase  by  monoethyl  peroxide 
is  probably  not  due  to  SH  group  oxidation  but  to  an  analog  type  of  inhi- 
bition (Blaschko,  1935). 


PEROXIDES  695 

Effects  on  Metabolism 

Hydrogen  peroxide  inhibits  brain  respiration  and  especially  the  oxidation 
of  succinate  (Dickens,  1946  a).  If  one  compares  the  effects  on  various  tis- 
sues, the  sensitivity  to  hydrogen  peroxide  depends  on  the  relative  concentra- 
tions of  catalase;  the  more  catalase,  the  less  inhibition.  In  brain,  75  roM  hy- 
drogen peroxide  inhibits  respiration  36%  and  succinate  oxidation  95%  over 
a  60  min  period.  It  was  thought  initially  that  the  toxic  effects  of  high 
oxygen  tension  on  brain  might  be  due  to  hydrogen  peroxide  released,  but 
this  was  shown  not  to  be  true. 

If  Lactobacillus  is  grown  anaerobically,  the  cells  lose  their  iron  enzymes 
and  catalase;  if  they  are  then  exposed  to  oxygen,  hydrogen  peroxide  is  form- 
ed and  the  cells  are  killed  (Warburg  et  al.,  1957).  Since  cancer  cells  possess 
an  anaerobic  type  of  metabolism  and  contain  much  less  catalase  than  nor- 
mal cells,  it  was  postulated  that  this  may  be  the  cause  of  the  greater  sensi- 
tivity of  cancer  cells  to  hydrogen  peroxide.  It  was  found  that  1  mM  hy- 
drogen peroxide  has  no  effect  on  the  aerobic  or  anaerobic  glycolysis  of  em- 
bryo tissue,  but  inhibits  both  almost  completely  in  ascites  cells.  Since  the 
catalactic  activity  of  embryo  tissue  is  around  10-fold  that  of  the  ascites 
cells,  this  could  explain  the  differential  susceptibility.  Inasmuch  as  radia- 
tion of  cells  can  induce  hydrogen  peroxide  formation,  this  may  be  one  reason 
for  the  more  selective  effects  of  radiation  on  cancer  cells.  Holzer  and  Frank 
(1958)  extended  these  observations  in  ascites  cells  to  show  that  hydrogen 
peroxide  at  0.056  mM  not  only  inhibits  glycolysis  86%,  but  simultaneously 
reduces  the  NAD  concentration  very  markedly  (0.31  to  0.05  //moles/ml). 
Triose-P  and  fructose-diP  rise,  indicating  a  block  of  the  phosphoglyceral- 
dehyde  dehydrogenase.  However,  they  found  the  extracted  enzyme  to  be 
inhibited  only  37%  by  0.079  mM  hydrogen  peroxide,  so  that  concentra- 
tions effectively  blocking  glycolysis  would  have  little  effect  on  this  enzyme 
(assuming  the  same  sensitivities  of  the  intact  and  extracted  enzymes).  They 
thus  postulated  that  the  inhibition  is  due  to  a  reduction  of  NAD  and  that 
this  suppresses  the  oxidation  of  triose-P.  Nicotinamide  can  protect  both 
NAD  and  glycolysis  from  hydrogen  peroxide,  and  Pantlitschko  and  Seelich 
(1960)  showed  that  it  could  overcome  the  inhibition  when  added  1  hr  after 
the  hydrogen  peroxide.  Baker  and  Wilson  (1963)  confirmed  the  inhibition 
of  anaerobic  glycolysis  in  Ehrlich  ascites  carcinoma  ceUs,  although  the  ef- 
fects were  not  as  marked  as  observed  previously  —  some  inhibition  at  0.3 
mM,  around  50%  at  0.9  mM,  and  80%  at  2  mM  —  and  further  showed 
that  during  the  oxidation  of  unsaturated  fatty  acids  some  hydrogen  per- 
oxide is  formed  and  may  depress  glycolysis.  Piitter  (1961)  studied  the 
possible  relationship  between  glycolysis  and  transplantability  of  ascites 
cells,  but  encountered  the  difficulty  that  the  hydrogen  peroxide  used  to 
inhibit  glycolysis  is  fairly  rapidly  decomposed  so  that  the  inhibition  dis- 
appears. Thus  it  requires  above  1  mM  to  interfere  with  transplantability. 


696  5.    OXIDANTS 

Hydrogen  peroxide  is  by  no  means  the  ideal  glycolytic  inhibitor  for  this 
type  of  work. 

Effects   on    Tissue    Function    and    in    Whole   Animals 

The  spontaneous  motility  of  the  rat  intestine  is  extremely  sensitive  to 
hydrogen  peroxide  inasmuch  as  10%  stimulation  of  the  amplitude  occurs 
with  0.00057  mikf  (Goodman  and  Hiatt,  1964).  At  0.057  mM  hydrogen  per- 
oxide the  stimulation  of  contraction  by  acetylcholine  is  blocked  and  67% 
of  the  total  SH  groups  of  the  tissue  are  reacted.  Although  hydrogen  per- 
oxide at  0.001  vaM  has  no  definite  effect  on  the  spontaneous  contractility, 
it  reduces  the  effect  of  acetylcholine  somewhat.  Other  SH  reagents  act  sim- 
ilarly and  it  appears  that  the  response  to  acetylcholine  is  dependent  on 
SH  groups.  The  contractility  of  nonconducting  rabbit  psoas  muscle  is  block- 
ed by  300  mM  hydrogen  peroxide  after  7  min  exposure,  and  this  is  not 
reversible  with  cysteine  (Korey,  1950).  However,  little  can  be  learned  from 
concentrations  of  this  magnitude. 

An  interesting  relationship  was  discovered  by  Feinstein  et  al.  (1954),  in 
that  a  sublethal  dose  of  iodoacetate  (20  mg/kg)  and  a  20%  fatal  dose  of 
hydrogen  peroxide  (15  meq/kg)  given  together  kill  all  the  animals.  Inas- 
much as  iodoacetate  also  potentiates  the  lethality  of  X-irradiation  in  mice, 
this  was  considered  as  evidence  that  radiation  may  produce  some  of  its 
effects  by  the  release  of  hydrogen  peroxide.  It  was  noted  that  the  toxicity 
of  hydrogen  peroxide  is  markedly  increased  by  treating  the  animals  with 
azide,  a  catalase  inhibitor;  however,  hydroxylamine,  which  is  a  better  cat- 
alase  inhibitor,  does  not  augment  the  effects  of  hydrogen  peroxide. 


TETRATHIONATE 

Tetrathionate  appears  to  be  a  fairly  specific  oxidant  for  SH  groups  under 
the  proper  conditions,  but  has  been  used  very  little  in  enzyme  work.  It  was 
found  to  be  capable  of  antagonizing  cyanide  poisoning  in  dogs  at  doses  of 
500  mg/kg  (Chen  et  al.,  1934),  and  today  we  might  interpret  this  as  due  to 
methemoglobin  formation.  Tetrathionate  has  been  used  in  a  method  for 
the  determination  of  protein  methionine,  which  is  demethylated  to  homo- 
cysteine and  then  oxidized  (Baernstein,  1936).  It  has  been  used  clinically 
in  thromboangiitis  obliterans,  supposedly  for  an  effect  on  the  blood,  an  in- 
crease in  the  oxygen  capacity  being  observed  (Theis  and  Freeland,  1940). 
It  is  surprising  that  the  blood  glutathione  increases  after  injection  of  tetra- 
thionate. It  has  been  applied  occasionally  to  the  reduction  of  the  cytochrome 
components  of  the  respiratory  chain,  since  the  initial  work  of  Keilin  and 
Hartree  (1940),  the  tetrathionate  apparently  being  oxidized  to  sulfite.  It  is 
thus,  like  hydrogen  peroxide,  both  an  oxidant  and  a  reductant,  which  makes 


TETRATHIONATE  69? 

its  effects  in  complex  systems  more  difficult  to  interpret.  The  first  use  of 
tetrathionate  as  a  reagent  for  protein  SH  groups  was  by  Anson  (1941),  who 
demonstrated  that  it  would  titrate  denatured  ovalbumin,  although  the  reac- 
tion is  slower  than  with  ferricyanide,  porphyrindin,  or  p-chloromercuriben- 
zoate,  not  being  complete  in  3  min  at  neutrality.  It  has  not  been  used  ex- 
tensively for  this  purpose,  but  it  may  well  be  a  valuable  reagent  in  certain 
types  of  work;  a  detailed  description  of  the  method  is  given  by  Chinard 
and  HeUerman  (1954). 

Chemistry  and    Reaction   with   SH   Groups 

Sodium  tetrathionate  is  prepared  from  the  thiosulfate  by  oxidation  with 
iodine  in  90%  ethanol.  The  precipitate  is  purified  by  redissolving  it  in  an 
equal  weight  of  water  and  filtering  it  into  absolute  ethanol,  in  which  it  re- 
precipitates.  It  is  washed  with  ethanol  and  dried  in  vacuo.  Sodium  tetra- 
thionate crystallizes  with  2  waters  of  hydration.  When  it  is  kept  at  0°  in 
the  dark,  both  the  solid  and  0.1  M  solutions  are  stable  for  many  weeks 
(Pollock  and  Knox,  1943),  but  it  is  unstable  when  kept  under  ordinary 
conditions.  For  all  accurate  work  it  is  necessary  to  be  certain  that  it  is  free 
of  appreciable  thiosulfate  and  other  impurities,  and  it  should  be  recrystal- 
lized  as  above. 

Tetrathionate  rapidly  oxidizes  simple  thiols,  such  as  cysteine,  homocys- 
teine, and  glutathione,  according  to  the  reaction: 

2  R— SH  +  S^Og"  ^  R— S— S— R  +  2  S^Og"  +  2  H+ 

In  titrations  of  SH  groups  the  thiosulfate  is  determined  iodometricaUy. 
According  to  Baernstein  (1936),  it  is  specific  for  SH  groups  and  does  not 
react  with  other  amino  acid  groups.  It  is  not  a  strong  oxidant,  since  the 
standard  oxidation-reduction  potential  for  the  reaction 

2  §203=  ±?  8406=  +  2  e- 

is  +  0.08  V.  Although  it  has  always  been  assumed  that  tetrathionate  oxi- 
dizes SH  groups  to  disulfide,  Pihl  and  Lange  (1962)  have  obtained  evidence 
that  sulfenyl  thiosulfate  groups  may  be  formed: 

R— SH  +  8406=  ->  R— 8— 8^03=  +  8^03=  +  H+ 

Incubation  of  phosphoglyceraldehyde  dehydrogenase  with  tetrathionate-S^^ 
leads  to  the  appearance  of  S^^  bound  to  the  protein,  and  the  binding  of  each 
S^^  is  associated  with  the  disappearance  of  one  SH  group. 

One  of  the  few  thorough  kinetic  studies  of  SH  group  oxidants  was  made 
by  Goffart  and  Fischer  (1948).  It  was  shown  that  tetrathionate  oxidizes 
protein  SH  groups  more  slowly  than  cysteine  or  glutathione  (Fig.  5-1). 


698 


5.    OXIDANTS 


Furthermore,  the  reaction  is  initially  rapid  but  in  most  instances  slows 
down  suddenly,  which  is  difficult  to  explain  for  the  simple  thiols.  The  oxi- 
dation proceeds  much  faster  at  pH  7  than  at  pH  5  (see  accompanying 
tabulation). 


Time  for  complete  reaction 

Protein 

(min) 

pH  5             pH  7 

Lens  protein 

140                  30 

Ovalbumin  (denatured) 

120                  12 

Myosin  (denatured) 

140                  50 

06 
0  5 

GSH  ^_______ 

-rSTEINt 

0.4 

[          ' 

0.3 

OVAL  BUM 
(OEN. 

-^^Z^Z^ 

02 

^^-'- 

"-^^^^^^^^ 

01 

/ 

Fig.  5-1.  Rates  of  reaction  of 
10  mil/  tetrathionate  with  the 
SH  groups  of  thiols  and  proteins 
at  pH  5.  (From  Goffart  and 
Fischer,    1948.) 


Inhibition  of  Enzymes  and   Metabolism 

Succinate  dehydrogenase  is  inhibited  around  90%  by  0.1  mM  tetrathi- 
onate (Keilin  and  Hartree,  1940).  This  is  an  effect  on  the  dehydrogenase  SH 
groups  according  to  these  authors  and  PhiHps  et  al.  (1947),  who  confirmed 
the  inhibition  on  succinate  dehydrogenases  from  several  tissues.  No  inhi- 
bition of  ascorbate  oxidation,  and  hence  of  the  cytochrome  system,  is  ob- 
served even  with  10  mM.  Succinate  protects  the  enzyme;  when  tetrathionate 
is  0.5  mM,  succinate  reduces  the  inhibition  from  96%  to  39%,  and  when 
it  is  0.1  mM  from  79%  to  10%  (pigeon  breast  enzyme).  The  inhibition  is 
only  partially  reversible  with  glutathione  or  cysteine.  In  work  with  suc- 
cinate dehydrogenase  it  may  be  well  to  consider  the  possibility  that  some 
of  the  inhibition  results  from  a  competitive  action  of  the  tetrathionate,  since 
it  has  negative  charges  appropriately  separated.  Choline  dehydrogenase 
from  rat  liver  is  also  quite  sensitive  to  tetrathionate,  38%  inhibition  re- 


TETRATHIONATE  699 

suiting  from  0.2  mM  and  93%  from  0.6  mM  (Gordon  and  Quastel,  1948). 
The  only  clear-cut  demonstration  of  reaction  with  enzyme  SH  groups  is 
that  of  phosphoglyceraldehyde  dehydrogenase  (Pihl  and  Lange,  1962).  Here 
tetrathionate  inhibits  as  well  as  /)-chloromercuribenzoate,  i.e.,  when  3  moles 
of  inhibitor  are  reacted  per  mole  of  enzyme,  the  activity  is  reduced  to  zero 
in  both  cases.  Also  the  inhibition  is  fully  reversible  with  thiols.  However, 
as  mentioned  above,  the  reaction  does  not  appear  to  be  a  simple  oxidation, 
but  involves  the  formation  of  sulfenyl  thiosulfate  groups.  The  enzyme  is 
very  sensitive,  since  0.005  mM  inhibits  completely  (enzyme  =  0.0005  mM) 
within  5  min. 

In  view  of  the  potent  inhibition  of  phosphoglyceraldehyde  dehydrogenase 
one  might  anticipate  tetrathionate  to  be  a  glycolytic  inhibitor.  Goffart  and 
Fischer  (1948)  attempted  to  demonstrate  a  Lundsgaard  effect  in  muscle, 
i.e.,  a  typical  contracture  such  as  produced  by  iodoacetate  and  certain  other 
SH  reagents.  Following  injection  into  rabbits,  the  extremities  become  weak 
but  the  muscles  remain  elastic  and  the  reflexes  normal;  if  the  gastrocne- 
mius is  stimulated,  it  does  not  go  into  contracture.  Intraarterial  injection 
produces  a  temporary  contracture  (or  at  least  some  inhibition  of  relaxation). 
Injection  into  frogs  does  not  give  an  iodoacetate-like  effect  and  the  isolated 
frog  rectus  abdominis  muscle  gives  only  a  temporary  contracture-like  reac- 
tion. It  is  doubtful  if  true  contractures  are  observed,  and  in  any  case  the 
tetrathionate  concentration  must  be  quite  high.  It  is  possible  that  the  phos- 
phoglyceraldehyde dehydrogenase  in  the  muscle  is  protected  by  a  perme- 
ability barrier  to  the  doubly  charged  inhibitor,  and  by  the  presence  of  NAD 
and  substrate  on  the  enzyme.  MacLeod  (1951)  found  inhibition  of  glycolysis 
in  human  spermatozoa,  but  the  tetrathionate  concentration  was  10  mM  and 
the  inhibition  progressed  very  slowly,  so  that  even  after  3  hr  the  glycolysis 
is  not  completely  depressed  (around  50%).  The  motility  decreases  simul- 
taneously with  the  reduction  in  glycolysis.  One  of  the  pitfalls  of  reversibility 
experiments  is  well  illustrated  here,  for  when  cysteine  is  used  there  is  a  rapid 
toxic  effect  on  the  spermatozoa,  this  being  due  to  the  cystine  arising  as  the 
result  of  the  oxidation  by  tetrathionate. 

Tetrathionate  is  reduced  to  thiosulfate  by  reaction  with  SH  groups  and 
it  has  been  supposed  that  the  rapid  conversion  into  thiosulfate  in  rabbits 
and  dogs  is  due  to  this  (Gilman  et  al.,  1946).  While  this  must  be  true  in  part, 
there  is  some  evidence  for  enzyme  systems  catalyzing  this  reaction.  Thus  in 
various  bacteria  tetrathionate  is  readily  reduced,  while  in  others  no  reaction 
at  all  occurs  (Pollock  and  Knox,  1943).  Postgate  (1956)  has  isolated  cell-free 
systems  reducing  tetrathionate,  thiosulfate,  and  sulfite  from  the  anaerobe 
Desulfovibrio  desulfuricans,  the  cytochrome  system  acting  as  an  electron 
carrier  for  the  tetrathionate  reductase.  Indeed,  it  is  likely  that  tetrathio- 
nate can  be  oxidized  through  the  cytochrome  system  in  most  cells.  Thus 
tetrathionate  must  generally  be  rather  labile  in  most  biological  situations. 


700  5.    OXIDANTS 

Nephrotoxic  Action 

Intravenous  injection  of  around  0.5  g/kg  of  sodium  tetrathionate  into 
dogs  leads  to  the  development  of  anuria  within  30-60  min,  a  rapid  reduc- 
tion in  creatinine  clearance,  and  the  appearance  of  proximal  tubular  ne- 
crosis (Gilman  et  al.,  1946).  At  the  time  of  death  there  is  no  evidence  of 
toxicity,  symptomatic  or  histological,  in  any  other  tissue  from  such  min- 
imally lethal  doses;  higher  doses,  however,  can  cause  a  long-lasting  ataxia, 
and  in  rabbits  some  difficulty  in  muscular  relaxation.  Inasmuch  as  simul- 
taneous reduction  of  the  tetrathionate  to  thiosulfate  occurs,  it  was  assumed 
that  the  renal  damage  is  related  to  the  oxidation  of  SH  groups,  the  toxicity 
thus  being  related  to  that  produced  by  the  mercurials.  Nevertheless,  nephro- 
toxic doses  in  rabbits  do  not  inhibit  kidney  succinate  dehydrogenase  at  aU 
in  vivo,  and  yet  this  enzyme  is  very  sensitive  to  tetrathionate  (Philips  et  al., 
1947).  Large  doses  of  tetrathionate  (1  g/kg  of  the  sodium  salt)  in  rabbits 
lead  to  a  77%  loss  of  glutathione  in  the  kidneys,  28%  in  the  blood,  30% 
in  the  liver,  and  20%  in  muscle  after  90  min,  most  of  the  change  occurring 
within  30  min  (Goffart  and  Fischer,  1948).  These  results  might  indicate  that 
tetrathionate  has  a  greater  effect  on  renal  SH  groups  than  those  of  other 
tissues,  without  implying  that  the  toxicity  is  due  to  the  loss  of  glutathione. 


CHAPTER  6 

o-IODOSOBENZOATE 


The  most  commonly  used  oxidant  for  enzyme  SH  groups  at  present  is 
o-iodosobenzoate  because  it  is  probably  the  most  selective  for  these  groups. 
For  this  reason  it  deserves  a  somewhat  more  detailed  treatment  than  the 
other  oxidants  and  a  separate  chapter.  o-Iodosobenzoate  was  first  prepared 
by  Meyer  and  Wachter  (1892)  and  studied  biologically  by  Heinz  (1899)  in 
Germany.  The  early  interest  stemmed  from  the  use  of  iodine  and  organic 
iodine  compounds  in  superficial  infections.  Indeed,  Heinz  was  mainly  con- 
cerned with  the  administration  of  sodium  iodide  and  o-iodosobenzoate  to- 
gether so  that  by  the  oxidation  of  the  iodide  it  would  be  possible  to  form 
"nascent"  iodine  in  the  tissues.  Consequently  there  w^ere  early  investiga- 
tions on  the  antibacterial  activity  (Jahn,  1914)  and  the  effects  on  phagocy- 
tosis (Arkin,  1912).  The  initial  pharmacological  study  was  by  Loevenhart 
and  Grove  (1909,  1911)  at  the  University  of  Wisconsin,  but  the  results  did 
not  engender  much  clinical  enthusiasm  and,  inasmuch  as  the  actions  at 
that  time  were  not  related  to  any  metabolic  site  of  attack,  o-iodosobenzoate 
was  little  used  by  biochemists  until  it  was  introduced  by  Hellerman  et  al. 
(1941)  for  the  estimation  of  protein  SH  groups.  The  application  to  enzyme 
characterization  was  slow  but  during  the  past  several  years  it  has  come  to 
be  one  of  the  most  useful  SH  reagents.  It  differs  from  the  arsenicals,  the 
mercurials,  and  the  alkylating  agents  in  not  introducing  new  groups  or  side 
chains  onto  the  enzymes,  since  it  is  generally  believed  that  the  primary 
action  is  an  oxidation  of  the  SH  groups  to  the  disulfide  state.  However,  the 
use  of  o-iodosobenzoate,  like  most  SH  reagents,  in  complex  systems  or  cel- 
lular preparations  is  limited  because  of  the  number  of  components  affected 
and  the  inherent  difficulty  in  the  interpretation  of  the  results. 

CHEMISTRY 

The  structures  of  the  different  oxidation  states  of  the  iodinated  benzoates 
may  be  written  as: 

701 


702 


coo 


o-Iodobenzoate 


6.    0-IODOSOBENZOATE 

coo' 

I— O" 

o-Iodosobenzoate 


COO 


o-lodoxybenzoate 


It  is  possible  that  the  aryl  iodoso  compounds,  cp — 1+ — 0~,  can  add  a  proton 
to  form  9? — 1+ — OH  since  the  corresponding  protonated  iodoxybenzene  is 
known,  but  the  ionization  constant  is  unknown.  Indeed,  the  carboxyl  p^^ 
is  not  accurately  known,  but  is  probably  around  6.0  to  6.5,  in  contrast  to 
the  piiig  for  o-iodobenzoic  acid  (2.86),  and  o-iodosobenzoic  acid  may  be 
precipitated  from  solution  by  passing  COg  through  solutions  of  the  sodium 
salt.  o-Iodosobenzoic  acid  may  be  easily  prepared  by  oxidation  of  o-iodo- 
benzoic acid  with  permanganate,  crystallization  by  cooling,  and  recrystal- 
lization  with  COg,  and  can  be  determined  iodimetrically  (Loevenhart  and 
Grove,  1911;  Chinard  and  HeUerman,  1954).  Commercial  samples  should 
probably  be  repurified  for  accurate  work.  The  m-  and  p-iodosobenzoates 
are  also  strong  oxidizing  agents  and  might  possibly  have  certain  advantages 
over  o-iodosobenzoate  for  particular  purposes,  but  they  have  been  almost 
completely  ignored.  It  is  interesting  that  p-iodosobenzoate  reacts  like  o-iodo- 
sobenzoate with  the  SH  groups  of  L-glutamate  dehydrogenase,  but  no  de- 
tailed comparison  was  made  (Hellerman  et  al.,  1958).  The  o-iodoxybenzoate 
is  also  a  potentially  useful  reagent  but  essentially  nothing  is  known  of  its 
actions  on  proteins  or  enzymes. 

The  oxidation  of  SH  groups  by  o-iodosobenzoate  results  in  the  formation 
of  o-iodobenzoate.  Whether  the  disulfide  link  is  intra-  or  intermolecular  de- 
pends on  the  thiol  reacted;  with  cysteine  or  glutathione  it  is  obviously  in- 
termolecular, but  what  evidence  exists  for  proteins  suggests  that  intramo- 
lecular oxidation  is  dominant.  Oxidation  apparently  does  not  proceed  be- 
yond the  disulfide  stage: 


COO 


COO 


I— O 


SH 


SH 


+         R 


/' 


+       H,0 


at  pH  7,  and  under  proper  conditions  accurate  titration  of  SH  groups  can 
be  achieved.  However,  at  lower  pH's  or  in  the  presence  of  excess  of  o-iodo- 
sobenzoate, further  oxidation  to  the  sulfinate  or  sulfonate  stages  may  occur, 
and  groups  other  than  SH  may  be  attacked.  Whether  a  free  radical  mecha- 
nism is  involved  in  the  oxidation  of  SH  groups  here  is  not  known,  but  if  so, 
one  might  postulate  the  following  types  of  reaction: 


REACTION   WITH    PROTEIN   SH   GROUPS  703 

-S— S— R 
R— SH ^R-S-= '"'    >   R— S— I  — <)— coo" 


Reaction  (1)  would  involve  combination  with  another  R — S-,  while  reac- 
tions (2)  and  (3)  would  involve  additional  o-iodosobenzoate.  Compounds  of 
the  type  R — S — I — R'  are  generally  unstable,  but  on  proteins  such  link- 
ages may  occasionally  be  more  stable,  as  in  the  formatiom  of  sulfenyl  thio- 
sulfates  (page  697). 

At  neutrality,  25°,  and  1-5  loM  o-iodosobenzoate,  only  SH  groups  are 
significantly  oxidized;  cystine,  methionine,  glucose,  and  ascorbate  are  not 
oxidized  appreciably  under  these  conditions.  However,  ascorbate  is  oxi- 
dized slowly  by  o-iodosobenzoate  at  pH  4.6  when  both  are  present  at  1  mM 
in  acetate  buffer,  half-reaction  time  being  around  80  min  (Caraway  and 
Hellerman,  1953).  The  nature  of  the  buffer  is  important  inasmuch  as  it  is 
an  acid-catalyzed  reaction.  It  is  interesting  that  m-  and  p-iodosobenzoates 
oxidize  ascorbate  almost  instantaneously.  NADH  is  not  oxidized  by  o-iodo- 
sobenzoate at  pH  4.6  (Schellenberg  and  Hellerman,  1958).  The  oxidizability 
of  tyrosine  phenolic  groups  by  o-iodosobenzoate  has  not  been  thoroughly 
examined  but  there  is  no  evidence  from  work  with  proteins  that  this  reac- 
tion proceeds  readily.  The  tyrosine  groups  of  /5-amylase  seem  to  be  resistant 
to  o-iodosobenzoate  since  there  is  no  change  in  absorption  at  280  mju 
(Englard  et  al,  1951). 

REACTION  WITH  PROTEIN  SH  GROUPS 

In  order  to  titrate  selectively  protein  SH  groups  with  o-iodosobenzoate, 
it  is  necessary  to  control  the  conditions  carefuUy,  as  with  all  SH  reagents. 
It  is  usual  in  titrations  with  o-iodosobenzoate  to  add  a  slight  excess  of  the 
reagent  and  determine  the  amount  not  reduced  by  addition  of  KI  and  sub- 
sequent titration  of  the  released  Ig  with  thiosulfate  (Chinard  and  Heller- 
man, 1954).  The  test  is  best  run  at  pH  7  and  between  15°  and  25°.  The 
required  reaction  time  varies  with  the  protein  tested  but  is  usually  less 
than  30  min.  Ovalbumin  denatured  with  guanidine  is  titrated  quite  satis- 
factorily and  all  of  the  SH  groups  are  oxidized.  Only  a  fraction  of  the  SH 
groups  of  native  ovalbumin  or  other  proteins  is  oxidized. 

The  specific  oxidation  of  protein  SH  groups  to  disulfide  may  be  accom- 
panied by  changes  in  the  protein  structure,  which  could  have  important 
bearing  on  the  mechanism  of  enzyme  inhibition.  Evidence  for  such  struc- 
tural alteration  is  given  by  the  increased  water-binding  capacity  of  gels 
formed  from  serum  proteins  previously  treated  with  o-iodosobenzoate  (Jen- 
sen et  al.,  1950).  At  about  equimolar  ratios,  o-iodosobenzoate  changes  the 


704  6.    O-IODOSOBENZOATE 

nature  of  the  clots  from  soft  and  opaque  to  firm,  elastic,  and  almost  trans- 
parent, and  simultaneously  the  water  binding  increases  from  14.3  to  36.5 
g/g.  It  is  quite  possible  that  more  linear  proteins,  which  may  be  reasonably 
flexible,  can  be  altered  quite  markedly  by  the  formation  of  disulfide  bonds, 
and  that  particular  regions  on  the  surface  may  be  made  unavailable  for 
other  reactions. 

Further  evidence  for  structural  changes  induced  by  o-iodosobenzoate  is 
the  decrease  in  the  viscosity  of  G-actin  brought  about  by  2  mJf  of  the 
reagent  acting  for  30  min  at  25^  and  pH  7.8-8  (Barany  et  al.,  1962).  This 
is  interpreted  as  an  inhibition  of  polymerization.  Simultaneously  there  is  a 
decrease  in  the  ability  to  bind  Ca++,  as  shown  by  the  loss  of  G-actin-bound 
Ca^^  upon  tretment  with  o-iodosobenzoate.  Although  there  appears  to  be 
some  correlation  between  changes  in  viscosity  and  Ca++  binding  as  pro- 
duced by  various  SH  reagents,  the  mechanisms  involved  are  not  yet  under- 
stood. 

INHIBITION   OF  ENZYMES 

Most  SH  enzymes  are  inhibited  by  o-iodosobenzoate  (see  Table  6.1)  and 
in  a  few  instances  the  inhibition  is  very  marked.  Enzymes  without  SH 
groups  at  or  near  their  active  centers,  as  shown  by  failure  to  be  inhibited 
by  SH  reagents  in  general,  are  not  affected  by  o-iodosobenzoate,  except 
possibly  in  the  single  instance  where  it  has  been  claimed  to  act  as  a  com- 
petitive inhibitor  on  D-amino  acid  oxidase,  although  in  such  cases  the  con- 
centration would  usually  have  to  be  a  good  deal  higher  than  for  the  oxi- 
dation of  susceptible  SH  groups  (Frisell  et  al.,  1949)  (see  page  342).  Inas- 
much as  o-iodosobenzoate  is  used  up  in  the  reaction,  inhibition  is  of  a  titra- 
tion type  and  spontaneously  irreversible;  thus  the  degree  of  inhibition  will 
often  depend  on  the  amount  of  enzyme  present,  or  the  amount  of  some  SH 
containing  impurity  that  also  reacts  with  the  reagent.  In  most  complex 
systems,  many  enzymes  will  be  inactivated  to  varying  degrees,  and  probably 
little  specificity  is  possible.  However,  the  remarkable  sensitivity  of  creatine 
kinase  —  definite  inhibition  at  0.00001  mM  and  complete  inhibition  at 
0.00013  mM  —  might  well  make  it  possible  to  block  this  enzyme  selectively 
(Ennor  and  Rosenberg,  1954).  This  inhibition  may  explain  the  observation 
of  Bailey  and  Marsh  (1952)  that  the  fall  in  creatine  phosphate  in  muscle 
homogenates  is  almost  completely  prevented  by  o-iodosobenzoate.  It  might 
be  worthwhile  to  consider  the  use  of  o-iodosobenzoate  in  glycerinated  and 
similar  muscle  preparations  to  determine  the  role  of  creatine  kinase  in  the 
initiation  of  contraction  or  relaxation.  Phosphoglyceraldehyde  dehydro- 
genase seems  to  be  less  sensitive  to  o-iodosobenzoate  than  to  iodoacetate  or 
iodoacetamide,  so  that  a  specific  block  of  glycolysis  at  this  step  would  be 
impossible.  It  may  be  noted  that  several  dehydrogenases  are  quite  well 
inhibited  by  o-iodosobenzoate,  notably  the  xanthine,  malate,  and  aldehyde 


INHIBITION  OF  ENZYMES 


705 


IS 

H 

c 

H 

1— 1 

< 

g 

ill 

Iz; 

_o 

fn 

"-3 

n 

ei    ^~. 

o 

CO 

1? 

O 

§  s 

Q 

o    i- 

O 

c 

P5 


.§1 


a 


t^         ^ 


-«  ^ 


rt         r^ 


W         tij 


^ 

»o  ^ 

e 

fi 

^ 

2    T3 

"«3 

05 

e 

^^  c 

ce 

C 

^_^ 

^ 

>>   ^ 

c3 

O 

c 
a 
>> 

tf 

W  M 

Cm 

5 

r-< 

c: 

(M 

-H 

^-^ 

o 

»C 

^■^ 

;-l 

£ 

OS 

c 

<D 

o 

Ml 
C 

^ 

<D 

CZ3 

3 

73 

t3 
C 

^ 

^ 

C 

;^ 

a 

ce 

o 

a 

*—* 

u 

i=l 

rj 

(S 

<o 

c 

o 

ft 

ft 

o 

s 

ft 

o 

a 

b 

a 

m 

3 

iCCOOCOOOOOrZrcsOiOOt-OSt-OMGOt^fCiO 
MQO  »005»COC'n<  CCi05^^(M»CO<N-^iC00t^ 


^  00  --H  o    o 

TtfiCO'-i^OOOO'-H-HOO 


s 

Si 
H 


^ 


^ 


eS 

bS 

S 

h 

=0 

o 

o 

O 

S 

03 

S 

4^ 

_o 

^ 

*. 
^ 

s 

S 

r2 

(-1 

o 

2 

^ 
^ 

13 

-t^ 

-t^ 

s 

o 

03 

(D 

"« 

c3 

a 

e 

^ 

(^ 

pq 

H 

P5 

^ 

oq 

^ 


03       OJ 


^      03      q;) 


Ph 


3  -S 


4)  o, 

o3  -5 

-C  ft 

ft  tH 

03  -tJ 

o  » 

ft  x 


<    < 


A 

N 

ft 

a 

s 

4) 

a 

lU 

c 

M 

o 

c 

OS 

o 

cS 

cS 

S 

C 

M 

« 

4) 

n 

M 

M 

s 

O 

O 

t3 

T3 

t^ 

£^ 

>^, 

-0 

"S 

U 

T3 

TJ 

41 

-s 

"o 

>> 

^ 

-C 

fl 

o 

a; 

03 

o 

TJ 

^ 

^ 

< 

706 


6.    O-IODOSOBENZOATE 


O      lO 

CD      05 


K 


h3 


tc      CO      ^i> 


T3    -tS 

c3      cS 


c3  c6 
ft  Ph 
ft    ft 

o    o 

in  m 


ft     u 


t3 

1^ 

eS 

o 

« 

-o 

C 

o 

C 

cS 

O 

+3 

o 

c 

S 

O 

<1 

<! 

p 

ce  ::_. 


»  i^H 


■73 

-*o 

C 

<» 

e« 

Ti 

tH 

0) 

cS 

C 

bC 

cS 

C 

h-5 

W 

CO  o  lo 

rt    >0    (M 
O    ^    CD 


OOOOOO     O      O     OOO     O 


O    I— I    <M     O    ^ 

d  d  d   d  d 


;^ 

u 

g, 

CO 

< 

s 

ft 

5_ 

03 

b 

s 

ft 

^    ^ 

c? 

;j3 

Oi 

_S 

OJ 

=0 

O 

> 

^ 

a> 

M     M 

J2 

Ph_ 

^ 

3 

-T] 

W 

S    S 

cq 

P5 


■p. 

X 

o 

£ 

0) 

c3 

X! 

4) 
13 

0) 

flH 

rs 

"? 

a) 

_S 

i 

03 

"C 

r2 

>. 

'm 

"^ 

ft 

O 

o 

o 

^ 

cS 

" 

O 

C 

3 
<1 

< 

CD 

O 
,D 

o 

£ 


^    (M    IC    t- 


H 

O 

ft 

-u 

^ 

rO 

0) 

c3 

is 

£ 


INHIBITION  OF  ENZYMES 


707 


Oi 

^„^ 

§ 

^^ 

, 

Oi 

1 

tr- 
io 
Oi 

(4 

c 

(-1 
3 

(M 

<N 

fq 

,^ 

1© 
05 

05 

e 

H 

TS 

^^^ 

s..^ 

^—^ 

c 

« 

'a 

>< 
c 

c 
o 

S 

J3 

m  § 


M 


o       :::;. 


^    ^    CO 


w 


-H  Ti< 


M  S 


»5 


^ 

-O 

ki 

05 

c 
o 

CO 

o 

-2 

o3 

^ 

P5 

S 

T3 

o   ^ 

d  d  -I    ^ 


d    d    o 


lO  F-H  o    o    -^ 


O  —I 

2  <=> 

o  o 

o  o 

o    d  d    -^ 


s 

•^ 

s 

s 

CM 

c3 

o 

a 

o 

p 

5!5 

o 

43 

1 

O 
S 

O 

-t3 

s 

^ 

<4-l 

Cn    o 


pq   Ph   a, 


K5 


P5 


— :  ao 


3 


■S3® 

Ph     W     M 


m 


<1 


O)  " 


rJ3 

c3 

« 

C 

2 

o 

"3 

^ 

_C 

ai 

.^ 

>> 

£ 

c3 

<U 

-fi 

O 

Tj 

eS 

m 

O 

a 


c3 
O 

01! 

-1^ 

13 

GO 

M 

<» 

G 
O 

'-3 

ci 

4^ 

o 

o 

0^ 

o 

-C 

^ 

Lj 

t: 

O 

Q 

O 

o 

708 


6.    O-IODOSOBENZOATE 


tf 


.§1 


W 


^  g      c3 


^  IC 


r> 

lO 

. — . 

s^ 

Id 

GO 

^ 

03 

lO 

a> 

C3 

-c 

^^ 

t3 

"e 

!<i 

o  S 
M  CO 


t-    O    GO    o    o 
(N    lO     GO  lO 


O    —     -H     O     O 


m 


>H  n,  fe] 


>       :-L 


o  ^ 


PQ 


W 


O 


»o 

cs 

1— 1 

— 

,^^ 

05 

o 

+3 

05 

w. 

vS 

_c 

'O 

c 

3 

cS 

-s 

c 

-« 

ce 

3 

N 

<S> 

c 

> 

o 

c6 

fM 

fe 

^-^ 

eo 

M 

lO 

lO 

05 

Oi 

"*— ' 

^^ 

fl 

c 

« 

cS 

in 

« 

S 

J 

« 

^ 

73 

o 

T5 

C 

)— H 

C 

cS 

c3 

^ 

c 

C 

o 

"3 

O 

5 

M 

O 

(MiMO-HCOOOCOOO-H&Or-HiMO© 


O    O     -H            O 

O     <M    »0    "O 

o  o"   d  —1   d  -H 

iM  o   d   d  d  d 

d 

^ 

C 

73, 

> 

o 

eS 

'2. 

"— * 

02 

be 

~  « 


tf 


pq  >H 


c3      O     «^ 

P5   O  M 


^ 

>v 

-S3 

73. 

n3 

^ 

o 

^ 

g 

rH 

'ci 

c^ 

'o 

cS 

01 

s 

S 

s 

"ft 

03 
X 

o 
J3 

o 
u 
Xi 
o 

I 

o 

o 

o 

Xi 
c4 

ft 

o 

o 

S 

O 

>> 

0) 

>> 
a 

5 

4) 

5 

5 

s  s 

«     C 

£> 

a    a> 

01 

m 

0)    bo 

CO 

eS 

C 

2 

c3 

Sd   o 

>> 

1 

4> 

It 

Ph    1 

>> 

^ 

03 

-TS 

Q^ 

c 

4) 

CO     fl 

_c 

-is 

15 

^ 

0)      § 

cS 

o 

(-1 

GO        ^J 

-ti 

oj 

8  ^ 

'3 

s 

o 

a 

3 

o 

,3  O 

W 

fa 

fa 

fa 

O    J 

INHIBITION  OF  ENZYMES 


709 


« 

-^ 

3" 

g 

l-H 

M 

Ci 

c* 

(M 

;-! 

^H 

7^ 

O 

^^^ 

«> 

■ 

> 

C5 

■*3 

-^ 

fc^ 

£ 

4) 

O 
-1-^ 

-c 

"-' 

^ 

4) 

-^ 

J^ 

~    ir^ 

J3 

eS 

'5 

c   _ 

eS 

s 

cS 

C 

°    li 

§ 

~ 

s 

0; 

C    ■« 

"2      C 

O 

'g 

c5 

3 

-§  i 

PS 

c» 

o 

S 

P5  o 

5      - 


^    a 


►^  K 


CM 


00    -H    (M    <M    O 

(M  ec  »c  »o 


O    O    ©    -H    o 


o 
d   ^ 


o  o  -^  o  o  ^  o 


-H         o^oo^oooo 


eg 

05 


.|  ^  :2  > 

S  IS  ^  " 

g  "^  .£P  1« 

05  tf  (1^  Ph 


> 

•--    3 

— 

"o  -a 

« 

fej  S 

pq 


O 


-d 

2 

« 

c£ 

^ 

<D 

CO 

IS 

^3 

s 

a^ 

'x 

[g 

CO 

o 

g 

cS 

>-, 

S 

.£ 

o 

3 

■+-• 

03 

^ 

"x 

ClJ 

^ 

® 

o 

CO 

® 

« 

'-(3 

■1^ 

tc 

© 

<K 

s 

€i 

-*J 

-»S 

4J 

-1-2 

cS 

« 

eS 

Sd 

C 

o 

^ 

eS 

3 

« 

.3 

"o 

o 

4) 

o 

o 

o 

3 

>. 

>■. 

>> 

o 

£ 

o 

o 

o 

o 

W 

CO 

o 

3 

e 

cS 

CS 

t^. 

hfi 

o 

R 

4; 

O 

W 

M 

73 

t3 

r( 

O 

a 

X 

W 

ij 

^ 

ec 

<o 

<& 

^ 

m 

c 

« 

i-*^ 

3 

be 

2 

4) 

2 

o 

T3 

a. 

<» 

T3 

>> 

CO 

-2 

j3 

>■, 

'>^. 

"q; 

-o 

X 

>< 

'C 

1 

o 

O 

C3 

o 

1 
'33 

o 

1 

"3 
o 

kO  HH  I— I 


710 


6.    O-IODOSOBENZOATE 


^ 


P5 


*^  H 


00 

lO 

Oi 

-* 

■ — ^ 

«c 

o^ 

>^, 

F-H 

TS 

^-v 

fi^ 

hJ 

oo 

e 

Tt^ 

O 

t3 

Ol 

05 

"S 

c5 

^-^ 

^H 

ro 

M 

iti 

-tJ 

C 

s 

O 

s 

tf 

■ft 

02 

<N 

00      C 

»o 

^    9 

J 

05 

^^    o 

o 

»o 

> 

T3 

'e 

1-H 

2w 

cS 

<D 

^-^ 

'^    T3 

(-( 

c 

(-1 

tH        C 

« 

o 

lU 

_«     cS 

c 

c 

'^ 

^ 

t4 
c5 

u 

"3 

03    'O 

^ 

> 

s 

S  Q 

jH  ^    ._    -^ 


W 


^  o  o 


O    !M    (N     lO    C<l    CO 

O    ■*    <M     lO    00    O 


-H    dsl    lO    ^<     O    ^H    (M 


^OOOOOi-HOOO 


O      -H 

-H    o    d    -H 


o 


Si 

=0        9, 


%   ^ 


m  ^ 


W)    CO     Q    ;^ 


s  ? 

^^ 

5    S    a 

«    <<i 

OJ       ? 

rH 

"!s       ID      Qi 

>    s 

S-(        r^ 

O 

.£f  .5P  .Sf 

5=1 

(In   W 

S 

S  S  S 

PL|  cq 

s 

0) 

«3 

fi 

^.— N 

<o 

tiD 

c8 

bc 

C 

2 

73 

j4 

>. 

'>! 

<o 

r^ 

X 

c 

<D 

o 

'o 

T3 

^ 

(h 

OJ 
m 

^ 
^ 

® 

o3 
o 

03 

S 
o 

c3 

Is 

^  ^ 

"    «« 


;§      Q     3 

03  <;    >:^ 


INHIBITION  OF  ENZYMES 


711 


^  ^  o 

o  lo  00  05 

Oi  Oi  ^  '^ 

f— I  1— I  05  ^^ 


S    £    S  -a    2 


61]    -^     ^ 

Pn    S    g  .5,  ^    .ii 
d  K  M  P5  H?  P5 


05     05 


^3  73 


^KK 


o3  .jO  O 

CO  CC  CO 

CO  CO  CO 

^  <^  Oi 


■-H        -P       ._      ._      ._ 

a 
a 
ft 
ft 
o 


ft 

ft 


m  m  ui   '^ 


e  ^      c:     C3     CB      e 

5    .«    .«      C     -t3    -(J    -ij     .►- 
0     O     O      3      c3     cS       ~ 


M   P)    §  H  H    h3 


3     3      N      C     C     C     ^ 

o  o  o  S 


1-^       ::i 


o 

3 


e    O    ^ 

._      3     02 


Q 


ft 
O 

Ah 

13 

3 


S     W 


(N     ■* 


.-      00 

3      OO 


-H      O     — I 

odd 


CO    t^  t^  CO 


s<i    ^    oo-H    CO    000 


d  d 


-d  o 


e  ^  eS 


3      « 


2    .S;  ^  a^  « 

"  C  3  .S 

«  rt  c3  -S 

3  3  3  a3 


^^S^aiOfespH        -t^Kl^Kffi 


CO  CO  oi  -r; 

oS  cs  C8  .0 

0  a;  o  cs 

>H  >1  >H  Ch 


P5 


P5 


00 

3 

*-<i 

^ 

ki 

0 

v 

-ti 

_> 

~= 

0 

s 

c3 

M 

>-q 

e5 

£ 

m 

c3 

c« 

3 

PM 
fe 

rX 

Q 

-1-2 

:2      ^ 


<» 

>. 

CO 

'0 

Ah 

-t^ 

3 

-4J 

S 

2. 

ft 

03 

J^ 

rj 

3 

0 

Lh 

© 

<u 

^ 

D 

Ah 

Ah 

Ah 

^  =  ^     -r. 


ft 


Ah    Ah 


'-' 

cS 

0 

OJ 

0) 

<A 

05 

o3 

p— , 

3 

CJ 

CJ 

0 

c^ 

!> 

b£ 

lU 

^ 

>> 

^, 

Oi 

Is 

« 

0 

M 

b£ 

r^ 

ji 

0 

M 

"^ 

'S 

0 

^ 

ft 

CO 

t; 

0 

0 

0 

;-i 

0 

0 

73 

-3 

A 

A 

73 

—3 

A 

0 

>> 

ft 

ft 

ft 

>■. 

ft 

ft 

j3 
Ph 

T3 

to 
0 

-3 

J3 

0! 

0 

0 

A 

CD 

Ah 

Ah 

Ah 

^ 

Ah 

712 


6.    O-IODOSOBENZOATE 


P5 


o 


-^   ^^     (M 


o 

o 

03 

o 

o 

^_^ 

^^^ 

03 

Ci 

^^ , 

Tt< 

^^ 

^^ 

^H 

t^ 

»o 

o 

„— V 

"■""^ 

'^'^ 

Tt* 

a> 

o 

lO 

cS 

s 

05 

r^ 

05 

!2 

a 

05 

"S 

^— ^ 

M 

c 

'*— ' 

t-i 

i-i 

>> 

o 

lO 

o3 

a 

-t^ 

Ci 

ISZ 

>> 

02 

m 

w 

^^ 

2 

-2 

T3 

c 

s 
> 

CD 
05 

"e 

73 

t3 

S 

2 

T3 

" — ' 

t3 

« 

t3 

^ 

cS 

cS 

eg 

4) 

'2 

C 

a 

02 

cS 
T3 

02 

■ft 

s 

c6 
ft 

ft 
O 

m 

C 

i3 

X 
3 
o 

M    2 


C 

<S 

<J 

-C 

'tS 

o 

o 

cc 

c 

'> 

^ 

ce 

c3 

ft 

P^ 

C5  O  ■" 

cs  rt  .2 

1  I  2 

1-1  02  O 


lOffOOOO      OOOOOOO      (M00     005     0 


O     O    O    O    ^    <N      — 1 


C<1      d      r-H 


OOfO     OOi-HOiO^ 


<Sl  3 


1 


X 

o 


Gj 

4) 

> 

oo 

^ 

cS 

cS 

g 

^ 

0) 

^ 

o 

"Sj 

c« 

C! 

s 

o; 

cS 

^ 

A^ 

^ 

?§ 

£ 

£ 

•<s> 

o 

o 

O 

C 

C 

^ 

-t> 

cS 

as 

cS 

p^ 

Ph 

^ 

P5 

s 

s 

<a 

•^ 

"oo 

«5l 

O 

H 

■?g 

o 

<U 

►o 

§ 

V 

5^ 

«J 

O 

o 

e 

o 

1 

>> 

X 

o 

<0 

o 

73 

a> 

ft 

03 

CO 

« 

cS 

o 

-1^ 

C! 

^ 

<s 

^ 

ft 

> 

o 

3 

O 

(1 

^ 

1h 

d^ 

PM 

(D 

03 

00 

c 

^ 

<D 

3 

M 

a> 

O 

a; 

O 

-s 

2 

tf 


INHIBITION  OF  ENZYMES 


713 


05 

IC 

,—1 

to 

^— V 

Tt* 

^ — ■ 

OS 

t~ 

Oi 

a 

T3 

i-H 

lO 

i-H 

^— ^ 

Is 
"3 

05 

05 

H 

H 

-c 

C5 

6 

a 

T3 

CR 

<M 

"2 

cS 

c3 

c 

rt 

> 

P5 

o 

d 
M 

M 

ft 
ft 

S 

-Hp^OOOOOO 


■D  -7" 


g      C       CD     ^     4^ 


Si  si    ^        'Si 


(X,  pq    Oh  m   M  |ii 


M 


>. 

eS 

O 

,i3 

">. 

T3 

T5 

X 

o 

>> 

"2 

tH 

"5 

'p 

T3 

-c 

2? 

^ 

43 

£3 

^ 

a> 

a> 

S 

t3 

cS 

fl 

« 

cS 

s 

-^  .-    >• 

ft 

s 


ft    o     Si 


O    »0    CO     --H    CO     o 
O    t^    05    CO    o    oo 


-H       -H       O       O 

(M     lO     O 


-H       ^       C^ 

O     O     CO 


o    S 


cS     -^ 


^      S      cS      S      =1 

S    M    Ph    tn    Ph 


IC  — - 


C      ©     ^     J3     r=     -r- 


la:)  W  M  H  tiij        mm 


Ol 

o 

(N 

»o 

«D 

^-^ 

05 

05 

bD 

tH 

C 
O 

C 

a 

tc 

ft 

b' 

-2 

C5 

>-^ 

.^ 

O 

'q3 

CO 

o 
o 

K 

t3 

-^ 

» 

3 

OD 

eS 

rO 

>> 

fi 

'm 

ki 

O 

43    ^-..    ^      O) 


43     -!5 


43      cS     "C 


■*     10    O    C'    o 


CO 

^    o 


W5 

o 
o 

rt       -H       rt       O       I-H      O 


^  >  I 


PL, 


13 

5? 

;i^ 

£ 

0 

0 

0^ 

0 

<U 

§ 

Pm 

pq 

Q 

N 

^ 

g    43   ^    c 


.ii     43    S     S     3 


§     ft    g 
H    H    H 


o    s 
Ph 


.S  -S 


cS      e    '3    '^ 


t^      =« 


p  p  p  t:^  >  kl    Q 


si  4) 


fl 

Oj 

_o 

0 

'-3 

^ 

o3 

Pi 

-fi 

•  1-4 

3 

^,— ^ 

0 

•<si 

C 

s 

'S 

(H 

^3 

ft 

c 

0 

eS 

43 

■-3 

03 

05 
C 

_cS 

"3 

43 

Si 

4J 

'-(3 
C 

'-3 

CO 

a: 

>-, 

M 

43 

02 

-IJ 

1 

;-i 

43 
T3 

c3 
43 

43 

43 

'O 

C 
43 

« 

43 

_g 

0 

£ 

H 

a 

cS 

714  6.    O-IODOSOBENZOATE 

dehydrogenases.  At  the  present  time,  o-iodosobenzoate  is  more  useful  in  the 
study  of  pure  enzymes  as  an  indicator  of  SH  groups  than  in  cellular  systems, 
but  has  been  little  investigated  in  the  latter  and  may  possess  potentialities 
as  a  metabolic  blocker  if  applied  properly. 

Titration  of  Enzyme  SH  Groups 

Muscle  phosphoglyceraldehyde  dehydrogenase  contains  11  cysteine  resi- 
dues and  reacts  rapidly  with  1 1  moles  of  ^^-chloromercuribenzoate  per  mole 
of  enzyme.  Segal  and  Boyer  (1953)  reported  that  7.3-7.45  moles  of  o-iodo- 
sobenzoate react  with  each  mole  of  this  enzyme,  indicating  14.6-14.9  re- 
ducing groups.  Theoretically  one  would  expect  5.5  moles  of  o-iodosobenzoate 
to  be  reduced  by  each  mole  of  enzyme,  assuming  that  all  the  SH  groups 
are  oxidized  to  the  disulfide  level.  Segal  and  Boyer  thus  suggested  that  some 
of  the  SH  groups  may  be  oxidized  beyond  the  disulfide  stage.  Actually  only 
10  of  the  11  SH  groups  could  form  intramolecular  disulfide  bonds,  so  the 
extra  SH  group  must  either  remain  unoxidized,  form  a  disulfide  link  with 
another  molecule  of  the  enzyme  (which  is  unlikely),  or  be  oxidized  to  some 
state  other  than  the  disulfide.  Since  oxidation  to  the  S — 0~  or  SOg"  state 
would  require  2  molecules  of  o-iodosobenzoate  for  each  SH  group,  7  moles 
of  o-iodosobenzoate  would  react  with  each  mole  of  enzyme  if  oxidation  of 
the  extra  SH  group  occurred  in  this  way.  Rafter  (1957)  investigated  this 
problem  further  and,  under  his  conditions,  found  10-11  moles  of  o-iodoso- 
benzoate to  react  with  each  mole  of  the  enzyme,  indicating  20-22  reducing 
equivalents.  Furthermore,  the  o-iodosobenzoate-treated  enzyme  still  pos- 
sesses 30%  of  its  initial  SH  groups,  as  determined  by  reaction  with  p-chloro- 
mercuribenzoate.  These  results  point  to  reaction  of  o-iodosobenzoate  with 
groups  other  than  SH  groups,  or  to  oxidation  of  a  fraction  of  the  SH  groups 
beyond  the  disulfide  stage.  There  is  no  evidence  for  reaction  with  other 
groups  and  the  enzyme  is  completely  reactivated  by  cysteine.  Thus  one 
might  assume  that  4  SH  groups  are  oxidized  to  the  disulfide  level,  4  are 
oxidized  beyond  this,  and  3  remain  unreacted;  this  would  require  10  moles 
of  o-iodosobenzoate  per  mole  of  enzyme.  Another  possibility  is  that  some 
of  the  SH  groups  form  E — S — I — cp — C00~  residues.  It  is  interesting  that 
although  o-iodosobenzoate  abolishes  the  usual  phosphoglyceraldehyde  dehy- 
drogenase activity  and  the  arsenolysis  reaction,  it  simultaneously  increases 
phosphatase  activity  6-fold,  this  phosphatase  activity  being  dependent  on 
NAD.  The  esterolytic  activity  with  p-nitrophenylacetate  as  substrate  is  in- 
hibited completely  in  10  min  when  4-5  moles  of  o-iodosobenzoate  have  react- 
ed per  mole  of  enzyme  (Olson  and  Park,  1964).  No  substrate  protection  was 
observed.  Yeast  phosphoglyceraldehyde  dehydrogenase  contains  fewer  SH 
groups  than  the  muscle  enzyme  —  2-4  per  molecule  —  and  reacts  with  6 
molecules  of  o-iodosobenzoate,  so  that  here  too  an  anomalous  effect  is  seen. 
Barron  and  Levine  (1952)  report  11.9  SH  groups  in  yeast  alcohol  dehydro- 


INHIBITION  OF  ENZYMES 


715 


genase  by  o-iodosobenzoate  titration  and  9.3  by  an  amperometric  method, 
and  thus  the  value  with  o-iodosobenzoate  is  a  little  high  in  this  case  also. 

Kinetics  of  Inhibition 

The  rate  of  oxidation  of  simple  thiols  by  o-iodosobenzoate  is  usually  quite 
rapid,  but  protein  or  enzyme  SH  groups  vary  greatly  in  the  rapidity  with 
which  they  react  with  this  reagent.  The  inhibition  of  succinate  oxidase  by 
o-iodosobenzoate  at  0.2  uiM  and  37°  requires  about  30  min  to  become  max- 
imal (Slater,  1949).  This  is  shown  in  Fig.  1-12-12,  where  the  maximal  inhi- 
bition of  around  35%  indicates  that  insufficient  o-iodosobenzoate  was  pres- 
ent for  reaction  with  all  of  the  enzyme  and  nonenzyme  material.  At  16°  it 
is  evident  that  there  are  two  phases,  one  complete  within  10  min  and 
the  other  incomplete  after  2  hr  (Fig.  1-12-13).  Although  the  first  phase 
undoubtedly  represents  oxidation  of  SH  groups,  it  is  not  clear  if  the  slow 
reaction  is  further  oxidation  of  other  SH  groups  or  secondary  inactivation 
of  the  enzyme. 

The  inhibition  of  ribonuclease  by  o-iodosobenzoate  also  shows  two  phases 
(Figs.  6-1  and  6-2),  one  a  fairly  rapid  reaction  inhibiting  around  20%  and 


Fig.  6-1.  Rates  of  inhibition  of  ribo- 
nuclease by  o-iodosobenzoate  at  pH  7. 
The  times  indicate  the  duration  of  the 
incubation  of  the  enzyme  and  inhib- 
itor.  (From  Ledoux,  1954.) 


the  other  a  much  slower  one  that  is  linear  at  least  over  1-2  hr  (Ledoux, 
1954).  The  rate  of  inhibition  during  the  second  phase  is  dependent  on  the 
concentration  of  o-iodosobenzoate  and,  since  this  phase  starts  from  essen- 
tially the  same  degree  of  inhibition  for  each  concentration,  it  seems  that 


716 


6.    O-IODOSOBENZOATE 


the  slow  phase  is  a  further  reaction  with  enzyme  groups  rather  than  a  sec- 
ondary inactivation.  It  has  been  emphasized  several  times  with  other  en- 
zymes, e.g.  /3-amylase  (Englard  et  al.,  1951)  and  phosphoglucomutase  (Mil- 
stein,  1961),  that  the  reaction  with  o-iodosobenzoate  is  slow.  For  this  reason, 
many  of  the  results  given  in  Table  6-1  are  not  comparable,  since  different 
times  of  incubation  with  the  inhibitor  were  used  and  usually  the  times  were 
not  given.  Unless  preincubation  of  the  enzyme  with  o-iodosobenzoate  for  a 
reasonably  long  period  (at  least  30  min)  is  done,  it  is  likely  that  the  inhibi- 
tions determined  are  partial  and  do  not  accurately  represent  the  true  effect 
of  the  oxidant  on  all  of  the  enzyme  present. 


Fig.   6-2.    Rates   of  inhibition   of  ribonuclease    by 
1  raM  o-iodosobenzoate  at  various  pH's.  The  times 
indicate  the  duration  of  the  incubation  of  the  en- 
zyme and  inhibitor.   (From  Ledoux,   1954.) 


Effects  of  pH 

It  has  been  emphasized  that  in  titrations  of  protein  SH  groups  a  pH  near 
neutrality  must  be  maintained  if  specificity  is  desired;  below  a  pH  of  7  the 
oxidizing  power  of  o-iodosobenzoate  increases,  and  oxidation  of  other  pro- 
tein groups  may  occur.  Also  the  solubility  of  o-iodosobenzoate  decreases 
rapidly  below  pH  7.  One  might  expect,  therefore,  that  the  inhibition  of  en- 
zymes might  decrease  as  the  pH  is  raised  above  7,  but  just  the  opposite 
has  been  observed.  /^-Amylase  is  inhibited  more  strongly  and  more  repro- 
ducibly  at  pH  7.8  than  at  pH  7  (Englard  et  al.,  1951 ),  succinate  dehydrogen- 
ase is  inhibited  more  at  pH  7  than  at  pH  6  (Stoppani  et  al.,  1953),  and 
ribonuclease  is  inhibited  more  rapidly  and  completely  as  the  pH  is  increased 
from  7  to  10,  no  inhibition  occurring  at  pH  5  (Fig.  6-2)  (Ledoux,  1954).  The 
very  rapid  initial  oxidation  of  ribonuclease  at  pH's  above  8.5  may  be  due 


INHIBITION  OF  ENZYMES  717 

to  the  ionization  of  the  SH  groups.  Penicillinase  is  inhibited  29%  by  2  xnM 
o-iodosobenzoate  at  pH  7.4  but  not  at  all  at  pH  6  (Smith,  1963  b).  On  the 
basis  of  these  results,  one  might  conclude  that  one  should  avoid  pH's  of 
7  or  below  in  enzyme  work.  Unfortunately,  there  are  no  data  on  the  speci- 
ficity of  o-iodosobenzoate  at  higher  pH's. 

Protection  of  Enzymes  against  Inhibition   by  o-lodosobenzoate 

Protection  of  an  enzyme  by  addition  of  some  thiol  with  the  o-iodoso- 
benzoate does  not  tell  one  anything  about  the  mechanism  of  inhibition, 
since  the  inhibitor  is  simply  depleted  by  oxidation  of  the  thiol.  However, 
protection  by  substances  interacting  with  the  enzyme  active  center  pro- 
vides some  evidence  for  the  site  of  the  inhibition  by  o-iodosobenzoate.  The 
normal  substrate  of  an  enzyme  has  been  shown  frequently  to  protect  against 
o-iodosobenzoate,  if  it  is  present  during  the  incubation  of  the  enzyme  with 
the  inhibitor.  Thus,  alcohol  dehydrogenase  is  protected  by  ethanol  (Barron 
and  Levine,  1952),  fumarase  by  either  fumarate  or  malate  (Favelukes  and 
Stoppani,  1958),  D-amino  acid  oxidase  by  alanine  (Frisell  and  Hellerman, 
1957),  choline  oxidase  by  choline  (Rothschild  et  al.,  1954),  glutamate  se- 
mialdehyde  reductase  by  glutamate  semialdehyde  (Smith  and  Greenberg, 
1957),  succinate  oxidase  by  succinate  (Thorn,  1959),  and  homogentisicase 
by  homogentisate  (Tokuyama,  1959).  In  some  instances  the  protection  may 
be  very  marked;  e.g.,  fumarase  is  completely  protected  against  0.5  vaM 
o-iodosobenzoate  by  25  mM  fumarate,  and  a  58%  inhibition  of  alcohol  dehy- 
drogenase is  reduced  to  6.5%  by  ethanol.  Coenzymes  can  likewise  protect 
in  certain  instances:  aldehyde  dehydrogenase  is  protected  by  NAD  and 
NADP  (Stoppani  and  Milstein,  1957  a,b),  alcohol  dehydrogenase  is  protect- 
ed by  NAD  (Barron  and  Levine,  1952),  and  D-amino  acid  oxidase  is  pro- 
tected by  FAD  (Frisell  and  Hellerman,  1957).  It  is  interesting  that  the  dif- 
ferent aldehyde  dehydrogenases  are  protected  to  different  degrees  by  their 
coenzymes.  The  K+-activated  yeast  enzyme  is  protected  by  both  NAD  and 
NADP,  as  well  as  by  acetaldehyde,  whereas  the  NAD-linked  liver  enzyme 
is  protected  only  by  NAD  and  not  by  NADP.  These  observations  may  be 
taken  to  mean  that  the  inhibition  is  the  result  of  a  reaction  of  o-iodoso- 
benzoate with  SH  groups  at  or  near  the  active  center,  and  that  when  the 
active  center  is  covered  by  substrate  or  coenzyme  the  oxidant  is  unable  to 
attack  these  SH  groups.  Such  effects  must  be  taken  into  account  when 
o-iodosobenzoate  is  used  in  cellular  preparations.  In  one  instance  the  effect 
of  substrate  is  abnormal.  The  lactate  dehydrogenase  of  Propionibacterium 
pentosaceum  is  inhibited  more  readily  in  the  presence  of  lactate  than  in  its 
absence  (see  accompanying  tabulation)  (Molinari  and  Lara,  1960).  This 
was  explained  by  assuming  that  lactate  increases  the  fraction  of  free  SH 
groups,  suggesting  that  the  SH  groups  may  be  involved  in  the  electron 
transport. 


718  6.    O-IODOSOBENZOATE 


o-Iodosobenzoate 

% 

Inhibition 

(mM) 

No  lactate 

Lactate  50  mM 

1 

5 

30 

2 

2 

36 

5 

36 

54 

Reactivation  of  o-lodosobenzoate  Inhibition 

If  the  only  action  of  o-iodosobenzoate  is  the  oxidation  of  SH  groups  to 
disulfide  bonds,  one  might  expect  some  reversal  of  the  inhibition  by  thiols, 
and  this  has  been  observed  with  certain  enzymes.  The  inhibition  of  D-amino 
acid  oxidase  is  reversed  completely  by  cysteine  (Rocca  and  Ghiretti,  1958) 
but  in  most  cases  only  partial  reactivation  is  possible,  for  example,  succinate 
oxidase  by  dimercaprol  (Thorn,  1959)  and  by  glutathione  (Slater,  1949), 
amylo-l,6-glucosidase  by  glutathione  (Earner  and  Schliselfeld,  1956),  thre- 
onine aldolase  by  dimercaprol  (Karasek  and  Greenberg,  1957),  and  alcohol 
dehydrogenase  by  glutathione  (Barron  and  Levine,  1952).  It  is  difficult  to 
interpret  partial  reactivation  since  failure  to  reverse  the  inhibition  com- 
pletely may  be  due  to  a  variety  of  factors.  No  reactivation  by  thiols  has 
been  reported  for  a  few  enzymes:  Acid  phosphatase  cannot  be  reactivated 
by  cysteine  or  thioglycolate  (Tsuboi  and  Hudson,  1955  b),  nor  xanthine 
oxidase  by  cysteine  (Harris  and  Hellerman,  1956),  nor  /^-amylase  by  gluta- 
thione or  dimercaprol  (Englard  et  al.,  1951),  nor  5-hydroxytryptophan  de- 
carboxylase by  thiols  (Buzard  and  Nytch,  1957).  However,  these  failures 
cannot  be  immediately  attributed  to  other  actions  of  o-iodosobenzoate  and 
perhaps  the  most  likely  explanation  is  progressive  secondary  inactivation 
consequent  to  the  protein  distortion  induced  by  disulfide  bond  formation. 
Some  failures  might  also  be  due  to  attempting  reactivation  in  the  presence 
of  oxygen,  which  can  often  oxidize  the  thiols  to  disulfides,  which  in  turn 
can  inhibit  the  enzyme,  as  pointed  out  by  Slater  (1949). 

Variation  of  the   Inhibition   with  the  Substrate   Used 

The  degree  of  inhibition  of  lipase  by  1  mM  o-iodosobenzoate  varies  with 
the  substrate  (Singer,  1948;  Singer  and  Hofstee,  1948  a).  The  inhibition  by 
p-chloromercuribenzoate  is  also  dependent  on  the  substrate  used  and  Singer 
postulated  that  the  mercurial  is  bound  near  the  substrate  site  so  that  it 
sterically  interferes  with  the  binding  of  the  substrates,  the  interference  be- 
ing greater  the  larger  the  substrate  molecule.  However,  how  this  explanation 
could  apply  to  o-iodosobenzoate  is  not  clear,  since  the  simple  formation  of 


INHIBITION  OF  ENZYMES  719 

a  disulfide  link  near  the  substrate  site  would  not  obviously  produce  a  steric 
effect.  Inasmuch  as  the  general  problem  of  dependence  of  inhibition  on  the 
substrate  will  come  up  several  times  with  other  inhibitors,  it  will  be  well  to 
suggest  some  of  the  possible  mechanisms  by  which  such  effects  can  arise. 

Substrate  %  Inhibition  by  o-iodosobenzoate   (1  m.M) 

Triacetin  50 

Tripropionin  67 

MonobutjTin  74 

Tributyrin  89 


(1)  It  is  conceivable  that  the  formation  of  a  disulfide  structure  can  dis- 
tort the  enzyme  structure  at  or  near  the  substrate  site  so  that  inhibition 
will  result,  and  this  inhibition  might  not  be  the  same  for  each  substrate, 
because  of  either  steric  factors  or  changes  in  the  spatial  position  of  the  en- 
zyme groups  involved  in  the  hydrolysis.  (2)  The  SH  reagent  might  not  pri- 
marily react  with  the  enzyme  but  with  the  substrate,  as  suggested  by  Wills 
(1960)  for  the  inhibition  of  pacreatic  lipase  by  p-chloromercuribenzoate. 
Since  this  enzyme  is  not  inhibited  by  o-iodosobenzoate,  a  relation  with  sub- 
strate cannot  be  established.  Wills  believes  that  the  mercurial  is  adsorbed 
onto  the  glyceride-water  interface  and,  in  order  to  examine  this  possibility, 
shook  tributyrin  with  10  nxM  p-chloromercuribenzoate,  washed  it,  and  then 
used  this  as  a  substrate;  marked  inhibition  was  noted,  indicating  a  rather 
strong  affinity  of  the  glyceride  for  the  mercurial.  However,  again  this  ex- 
planation would  not  seem  to  hold  for  o-iodosobenzoate,  since  it  should  in- 
hibit the  pancreatic  lipase  as  well  as  the  wheat  germ  lipase  (with  which 
Singer  worked)  if  the  substrates  are  altered.  Also  it  is  not  too  surprising 
that  a  molecule  like  p-chloromercuribenzoate  would  adsorb  at  an  oil-water 
interface,  since  it  possesses  polar  and  nonpolar  regions,  but  o-iodosobenzoate 
has  polar  groups  at  either  end.  (3)  A  group,  such  as  — S — I — (p — COO",  is 
actually  added  to  the  enzyme  near  the  substrate  site  and  interferes  steri- 
cally  as  p-chloromercuribenzoate  may  do. 

Another  explanation  involves  the  basic  kinetics  of  such  inhibitions.  It 
seems  to  have  been  generally  assumed  that  when  an  irreversible  inhibitor 
reduces  the  affinities  of  each  substrate  of  an  enzyme  equally,  the  inhibi- 
tions will  all  be  the  same,  which  can  readily  be  shown  not  to  be  true.  Let 
us  assume  that  Kj„  is  equal  to  K^,  the  true  dissociation  constant  of  the  ES 
complex,  so  that  the  uninhibited  rate  is  given  by: 

F^(S) 
(S)  +  K, 


720  6.    O-IODOSOBENZOATE 

If  only  the  affinities  of  the  substrates  are  altered  by  the  inhibitor,  we  may 
write  for  the  inhibited  rate: 


(S)  +  aK, 


where  a  is  a  factor  indicating  the  magnitude  of  the  effect  of  the  inhibitor 
on  the  substrate  binding  (a  >  1).  The  inhibition  is  then  given  by: 

a  -  1 

(6-1) 


(S')  +  a 


where  (S')  is  the  specific  concentration  of  the  substrate,  (^)IK,.  Thus,  even 
though  a  is  the  same  for  each  substrate,  the  inhibition  will  vary  with  (S'). 
Superficially  it  might  appear  that  an  irreversible  inhibitor  reducing  sub- 
strate binding  would  produce  inhibitions  independent  of  the  substrate  con- 
centration, but  such  is  not  the  case.  Some  of  the  confusion  arises  from  as- 
sociating this  type  of  inhibition  with  noncompetitive  inhibition,  which  it 
is  not  in  any  sense.  The  inhibitions  by  o-iodosobenzoate,  and  probably  most 
SH  reagents,  are  usually  competitive,  as  shown  by  the  protection  afforded 
by  the  substrate  when  it  is  present  during  the  incubation  with  the  inhibitor, 
and  K,  is  altered  rather  than  k^,  the  rate  constant  for  the  breakdown  of  the 
ES  complex  into  products.  It  may  be  noted  that  even  though  Kj„  is  not  Kg, 
but  the  more  complex  {k_i  +  li2)jk-^,  a  similar  expression  for  the  inhibition 
will  be  found,  and  the  substrate  concentration  will  play  a  role  in  the  degree 
of  inhibition  produced.  Furthermore,  if  the  inhibition  actually  is  noncom- 
petitive and  k^  is  altered  rather  than  K„  it  can  easily  be  shown  that  the 
inhibition  is  given  by: 

'"^  -  (6-2) 


^k. 


A:_i[(S')  +  l] 

where  /?  is  the  factor  by  which  k^  is  changed  by  the  inhibitor  (/?<!).  Here 
the  variation  of  the  inhibition  with  the  specific  concentration  of  the  sub- 
strate is  different  than  in  the  previous  case,  in  that  the  inhibition  rises  as 
(S')  increases,  as  long  as  (S')  does  not  greatly  differ  from  unity.  Indeed,  at 
high  substrate  concentrations,  t  =  1  —  /5,  the  usually  expected  purely  non- 
competitive inhibition  and,  likewise,  if  iii,,,  =  K^.,  ^  =  1  —  /?. 

It  is  consequently  not  necessary  to  assume  some  complex  mechanism  in- 
volving steric  factors  when  the  inhibition  is  found  to  vary  with  the  sub- 
strate used,  unless  the  specific  concentrations  of  all  the  substrates  are  kept 
equal.  It  was  stated  by  Singer  (1948)  in  his  study  of  lipase  that  the  sub- 
strate concentration  was  chosen  so  as  to  "just  saturate  the  enzyme  and 
thereby  give  optimal  activity."  However,  calculation  of  the  values  of  (S') 


INHIBITION   OF  METABOLISM  721 

from  the  -flT^'s  given  by  Singer  and  Hofstee  (1948  b)  shows  that,  for  the 
substrates  used  with  o-iodosobenzoate,  (S')  varies  from  4  to  23  at  least  and, 
furthermore,  the  variation  of  the  inhibition  with  (S')  is  as  one  would  expect 
from  Eq.  6-1,  i.e.,  it  decreases  with  increasing  (S').  Thus  the  results  with 
o-iodosobenzoate  can  be  explained  quite  simply.  However,  it  is  not  implied 
that  this  will  explain  all  of  the  results  obtained  by  Singer,  and  it  is  quite 
possible  that  with  p-chloromercuribenzoate,  where  a  bulky  group  is  added 
to  the  enzyme,  steric  factors  also  play  a  role.  The  purpose  of  the  foregoing 
treatment  is  to  indicate  the  importance  of  keeping  (S')  constant  when  com- 
paring inhibitions  with  different  substrates. 

INHIBITION   OF  METABOLISM 

Very  little  quantitative  work  has  been  done  on  the  effects  of  o-iodoso- 
benzoate on  glycolysis,  respiration,  the  tricarboxylate  cycle,  or  other  me- 
tabolic pathways,  so  that  the  following  is  not  so  informative  as  indicative 
of  possibly  interesting  experiments  to  be  done.  From  Table  6-1  it  is  evident 
that  few  glycolytic  enzymes  have  been  tested  and  these  few  are  not  particu- 
larly sensitive  to  o-iodosobenzoate.  Only  one  investigation  of  glycolysis 
in  vivo  has  apparently  been  reported,  that  of  Harting  (1947),  who  found 
o-iodosobenzoate  at  1  mM  to  stimulate  scallop  muscle  anaerobic  glycolysis, 
as  does  p-chloromercuribenzoate.  This  may  not  be  due  to  direct  action  on 
the  glycolytic  system,  but  to  some  effect  on  the  muscle  membranes  facilitat- 
ing glucose  entry.  Glycolysis  in  muscle  homogenates  is  definitely  inhibited 
by  4  milf  o-iodosobenzoate  (Bailey  and  Marsh,  1952).  The  changes  in  the 
pH  and  phosphate  fractions  with  time  are  modified,  as  shown  in  Table  6-2. 
The  fall  in  pH  is  quite  strongly  inhibited  and  the  normal  decrease  in  ATP 
is  accelerated,  presumably  by  inhibiting  ATP  formation.  The  effects  on 
fructose- 1,6-diP  are  interesting;  in  the  control  there  is  an  initial  accumula- 
tion followed  by  a  fall  to  low  levels  —  o-iodosobenzoate  blocks  the  accu- 
mulation partially,  but  what  does  accumulate  remains,  indicating  some  in- 
hibition of  aldolase  or  phosphoglyceraldehyde  dehydrogenase.  Part  of  the 
depression  of  the  early  accumulation  may  be  due  to  the  low  levels  of  ATP, 
but  it  is  likely  that  some  inhibition  is  exerted  on  the  enzymes  forming  fruc- 
tose-l,6-diP,  perhaps  hexokinase.  The  prevention  of  the  fall  in  creatine-P 
is  undoubtedly  due  to  the  potent  inhibition  of  the  transfer  of  the  phosphate 
to  form  ATP.  The  minor  accumulation  of  phosphoglyceraldehyde  produced 
by  o-iodosobenzoate  may  also  point  to  some  block  of  the  dehydrogenase, 
not  unlikey  at  this  rather  high  concentration. 

One  might  expect  o-iodosobenzoate  to  inhibit  respiration  fairly  strongly 
since  several  dehydrogenases  and  cycle  enzymes  are  quite  sensitive.  The 
respiration  of  sea  urchin  spermatozoa  is  depressed  almost  completely  by 
0.3-1  milf  o-iodosobenzoate,  although  lower  concentrations  around  0.1  txlM 


722  6.    O-IODOSOBENZOATE 

Table  6-2 
Effects  of  o-Iodosobenzoate  on  Glycolysis  in  Muscle  Homogenates" 


Conditions  ''^™^  Zl  pH        A  ATP         A  CrP        J  FrPP    A  TrioseP 

(mm) 


Controls 

3 

-0.28 

-14 

-27 

+27 

+  3 

10 

-0.76 

-19 

-29 

+  10 

+  4 

30 

-1.03 

-23 

-29 

-  2 

+  3 

60 

-1.04 

-23 

-29 

-  3 

+  3 

o-Iodosobenzoate 

3 

-0.14 

-19 

-   2 

+   9 

+  5 

(4  mM) 

10 

-0.26 

-20 

-   5 

+  13 

+   7 

30 

-0.35 

-20 

-10 

+  13 

+  7 

60 

-0.38 

-20 

-11 

+  13 

+   7 

"  The  values  for  the  phosphate  fractions  are  changes  in  the  per  cents  of  the  total 
acid-soluble  phosphorus.  (From  Bailey  and  Marsh,  1952.) 

may  stimulate,  a  phenomenon  seen  with  other  SH  reagents  (HgClg,  js-chloro- 
mercuribenzoate,  arsenite,  and  iodoacetamide)  (Barron  et  al.,  1948).  The 
respiration  of  Ehrlich  ascites  tumor  cells  is  inhibited  50%  by  0.35  mM 
o-iodosobenzoate,  0.1  mM  inhibiting  11%  and  1  mM  93%  (Shacter,  1957). 
Thus  the  susceptibility  of  respiration  is  confirmed  but  there  are  no  data 
for  locating  the  principal  sites  of  action. 

The  binding  of  K+  in  liver  mitochondria  is  believed  by  Gamble  (1957)  to 
be  related  to  the  sites  for  oxidative  phosphorylation,  although  it  is  not 
directly  dependent  on  ATP.  The  evidence  comes  from  the  ability  of  2,4- 
dinitrophenol  to  lower  mitochondrial  K+  markedly.  o-Iodosobenzoate  at 
0.03  mM  produces  effects  similar  to  2,4-dinitrophenol,  which  does  not  nec- 
essarily imply  an  uncoupling  action  of  the  o-iodosobenzoate,  but  indicates 
some  effect  on  the  electron  transport  chain.  Scott  and  Gamble  (1961)  have 
found  mercurials  to  stimulate  the  K+  exchange  rate  of  mitochondria  and 
simultaneously  to  reduce  the  bound  K+.  These  effects  are  also  produced  by 
o-iodosobenzoate:  the  exchange  rate  is  doubled  by  0.08  mM,  the  mitochon- 
drial K+  is  half  reduced  by  0.15  mM,  and  oxidative  phosphorylation  is  50% 
inhibited  by  0.08  mM.  These  potent  actions  of  o-iodosobenzoate  point  to 
important  effects  on  mitochondrial  oxidative  systems  that  apparently  play 
a  role  in  the  depression  of  respiration. 


EFFECTS    ON    ANIMAL    TISSUE    FUNCTIONS  723 

EFFECTS  ON  ANIMAL  TISSUE   FUNCTIONS 

The  injection  of  o-iodosobenzoate  into  animals  or  its  application  to  skeletal 
muscle  preparations  does  not  produce  rigor  so  readily  as  does  iodoacetate. 
In  the  whole  animal,  indeed,  the  actions  on  muscle  seem  to  be  of  little  im- 
portance, and  the  paralysis  sometimes  seen  is  more  likely  explained  by  a 
central  effect.  Applied  directly  to  isolated  frog  muscle  in  reasonably  high 
concentration,  o-iodosobenzoate  can  lead  to  a  loss  of  excitability  and  the 
development  of  contracture,  whereas  o-iodobenzoate,  although  it  depresses 
excitability  somewhat,  does  not  induce  contracture  (Jahn,  1914).  The  turtle 
biceps  muscle  is  also  slowly  and  irreversibly  shortened  by  o-iodosobenzoate 
(and  iodoacetamide)  at  concentrations  much  higher  than  would  be  reached 
in  vivo  (Pisanty,  1948).  If  the  mechanism  of  iodoacetate  in  contracture  is  a 
block  of  glycolysis,  o-iodosobenzoate  does  not  seem  to  share  this  selectivity, 
which  confirms  what  little  is  known  from  the  results  on  enzymes  and  me- 
tabolism (see  page  721).  The  ability  of  myosin  to  associate  with  actin  and 
to  split  ATP  depends  on  SH  groups  and  is  inhibited  by  o-iodosobenzoate  as 
well  as  by  other  SH  reagents  (Bailey  and  Perry,  1947),  and  the  binding  of 
Ca++  by  G-actin  is  depressed  parallel  to  the  reduction  in  polymerization 
by  o-iodosobenzoate  (Barany  et  al.,  1962).  In  these  respects,  o-iodosoben- 
zoate is  more  potent  and  rapidly  acting  than  iodoacetamide,  and  such  ef- 
fects may  play  a  role  in  the  contractures  observed  at  high  concentration, 
although  a  metabolic  site  of  action  is  not  excluded.  The  contractile  response 
to  ATP  by  nonconducting  psoas  muscle  fibers  is  abolished  by  0.5-1  mM 
o-iodosobenzoate,  and  this  is  reversible  if  the  fibers  are  incubated  for  90- 
120  min  in  10  mM  cysteine  (Korey,  1950). 

The  heart  appears  to  be  more  sensitive  to  o-iodosobenzoate  than  is  skele- 
tal muscle.  In  the  initial  pharmacological  study  by  Loevenhart  and  Grove 
(1911),  intravenous  injection  into  rabbits,  cats,  and  dogs  was  found  to  pro- 
duce a  rapid  fall  in  the  blood  pressure,  little  change  in  the  cardiac  rate,  and 
a  decrease  in  cardiac  output  with  dilation  of  the  heart.  o-Iodoxybenzoate 
acts  very  similarly  but  o-iodobenzoate  is  inactive,  indicating  that  the  oxidiz- 
ing activity  is  essential.  Minimal  effects  are  given  in  the  cat  by  13.2  mg 
(50  //moles),  so  that  the  total  concentration  is  probably  around  0.5  mM. 
However,  inasmuch  as  Jahn  (1914)  showed  that  blood  appreciably  reduces 
the  action  of  o-iodosobenzoate  —  due  to  reaction  with  hemoglobin,  other 
proteins,  and  glutathione  —  the  concentration  of  free  o-iodosobenzoate  is 
undoubtedly  much  less.  Jahn  also  showed  that  the  perfused  frog  heart  is 
depressed  by  as  little  as  0.038  mM  o-iodosobenzoate  and  that  0.38  mM 
causes  a  prolonged  depression  of  the  amplitude,  although  not  standstill  or 
contracture.  The  results  of  Mendez  (1946)  and  Mendez  and  Peralta  (1947) 
on  the  frog  heart  differ  from  those  of  Jahn,  in  that  concentrations  of  0.2- 
0.4  mM  were  found  to  cause  an  increase  in  the  contractile  amplitude,  and 
0.83  mM  to  produce  systolic  standstill  within  20  min.  Furthermore,  the 


724  6.    O-IODOSOBENZOATE 

rate  always  increases  up  to  the  final  failure,  whereas  Jahn  observed  only 
slowing.  No  conduction  disturbances  were  noted.  Since  partial  reversal  can 
be  achieved  by  lengthy  perfusion  with  o-iodosobenzoate-free  medium,  in 
tissues  some  reduction  of  disulfide  groups  may  occur.  The  dog  heart-lung 
preparation  is  quite  resistant  to  o-iodosobenzoate,  100  mg  producing  no  ef- 
fect, although  high  doses  increase  the  venous  pressure,  presumably  by  ini- 
tiating cardiac  failure  (Mendez  and  Pisanty,  1949). 

The  effects  of  o-iodosobenzoate  on  smooth  muscles  are  at  least  superficially 
similar  to  those  on  skeletal  and  heart  muscle  (Alanis,  1948).  Isolated  rabbit 
intestine  and  the  uterus  in  several  species  are  put  into  a  form  of  contrac- 
ture, although  this  is  eventually  followed  by  relaxation  and  loss  of  aU  rhyth- 
mic activity.  These  actions  are  similar  to  those  of  iodoacetamide  and  ar- 
senicals. 

Although  the  effects  of  o-iodosobenzoate  in  the  whole  animal  indicate 
marked  effects  on  the  central  nervous  system,  no  analysis  of  this  has  been 
made  so  that  sites  and  mechanisms  are  completely  unknown.  Frog  nerve 
axons  are  unaffected  by  1  mM  o-iodosobenzoate  as  measured  by  excitability 
and  conduction  (Jahn,  1914).  but  the  central  actions  are  undoubtedly  on 
synaptic  mechanisms.  Neuroblastic  damage  has  been  found  in  developing 
mice  and  rats  after  injections  of  o-iodosobenzoate,  as  with  other  SH  re- 
agents, but  this  probably  relates  more  to  growth  and  differentiation  than 
function  (Hicks,  1953). 


EFFECTS  IN  WHOLE  ANIMALS 

The  earliest  study  of  o-iodosobenzoate  by  Heinz  (1899)  is  not  ve»y  illu- 
minating since  he  administered  potassium  iodide  simultaneously  to  generate 
"nascent"  iodine.  However,  he  showed  it  to  be  irritant  to  the  eye,  the  gas- 
tric mucosa,  and  the  peritoneum,  and  that  this  action  seems  to  be  due  to 
something  other  than  its  acidic  properties.  Loevenhart  and  Grove  (1911) 
confirmed  the  inflammatory  action  in  the  eye  and  subcutaneously,  and  show- 
ed that  intraperitoneal  injections  can  be  fatal  as  a  result  of  the  congestion 
produced.  Whether  this  is  related  in  any  way  to  the  vesicant  activity  of 
many  SH  reagents  is  not  known.  On  the  other  hand,  Bernheim  et  al.  (1932) 
found  that  injection  into  rabbits  of  o-iodosobenzoate  inhibits  conjunctival 
edema  induced  by  mustard  oil.  However,  this  could  well  be  a  nonspecific 
action,  since  o-iodobenzoate  and  benzoate  are  somewhat  active  (as  the  am- 
monium salts),  and  might  well  be  mediated  through  the  adrenal  cortex. 

Loevenhart  and  Grove  (1909,  1911)  investigated  the  pharmacological 
properties  of  o-iodosobenzoate  and  related  compounds  because  they  believed 
that  the  oxygen  of  this  substance  is  physiologically  active  and  can  be  used 
by  the  tissues;  e.g.,  o-iodosobenzoate  alone  does  not  oxidize  phenolphthalin 
to  phenolphthalein,  but  does  if  some  serum  is  present,  this  being  interpreted 


EFFECTS    IN    WHOLE    ANIMALS  725 

as  an  action  mediated  by  peroxidase,  the  o-iodosobenzoate  acting  like  hy- 
drogen peroxide  —  furthermore,  the  taste  of  o-iodosobenzoate  is  almost 
exactly  like  hydrogen  peroxide.  Injection  of  10-20  //moles  of  o-iodosoben- 
zoate into  animals  causes  an  immediate  and  marked  depression  of  the  res- 
piration usually  lasting  2-3  min,  from  which  recovery  occurs  spontaneously. 
o-Iodoxybenzoate  is  somewhat  more  potent  but  o-iodobenzoate  is  inactive. 
Higher  doses  are  required  to  elicit  the  circulatory  depression  described  above 
and  the  apnea  is  not  secondary  to  the  fall  in  blood  pressure.  Antagonism 
between  o-iodosobenzoate  and  cyanide  on  the  respiration  (the  latter  stim- 
ulates respiration)  is  also  observed  and  felt  to  support  the  concept  that 
o-iodosobenzoate  acts  by  giving  up  its  active  oxygen.  Jahn  (1914)  observed 
rather  nonspecific  toxic  effects  in  frogs,  followed  by  a  slowly  developing 
paralysis  and  loss  of  reflexes,  death  occurring  when  reflex  activity  has  drop- 
ped to  zero  and  cardiac  failure  is  evident.  o-Iodobenzoate  is  less  than  one 
tenth  as  toxic.  The  relative  inactivity  of  o-iodobenzoate  in  all  of  these  stud- 
ies makes  it  very  unlikely  that  any  of  the  actions  of  o-iodosobenzoate  are 
due  to  the  former  compound,  which  undoubtedly  is  formed  in  the  tissues. 
Jahn  postulated  an  enzyme  that  splits  the  iodine  from  o-iodobenzoate  since 
he  found  both  organic  and  inorganic  iodine  in  the  urine  after  o-iodosoben- 
zoate, the  product  presumably  being  salicylate. 

Very  interesting  effects  on  the  blood  are  observed  following  intravenous 
infusion  of  0.5  millimole  of  o-iodosobenzoate  into  rabbits  (Loevenhart  and 
Grove,  1911).  Over  a  period  of  3  days  there  is  a  slight  depression  of  the 
erythrocytes  (around  15%)  and  negligible  effects  on  coagulation  mechanisms 
but  there  appears  early  a  very  marked  leucocytosis,  this  being  confined 
almost  entirely  to  the  polymorphonuclears,  which  increase  from  2,160  to 
11,362  in  24  hr.  It  is  not  known  if  this  stems  from  a  reaction  with  SH  groups 
or  an  action  on  some  metabolic  system. 

One  factor  which  must  be  taken  into  account  in  considering  the  effects 
of  any  SH  reagent  on  the  whole  animal  is  the  possible  release  of  active 
substances.  Thus  o-iodosobenzoate  at  fairly  low  concentrations  (0.1  nxM) 
releases  catecholamines  from  the  isolated  chromaffine  granules  of  the  adrenal 
medulla  (D'lorio,  1957).  This  was  thought  to  be  an  effect  on  the  SH  groups 
located  in  the  granule  membranes,  but  there  is  no  evidence  for  any  mecha- 
nism. On  the  other  hand,  the  release  of  histamine  from  rat  peritoneal  mast 
cells  by  Compound  48/80  is  inhibited  by  o-iodosobenzoate,  and  presumably 
histamine  is  not  released  by  o-iodosobenzoate  alone  (VanArsdel  and  Bray, 
1961). 

The  intravenous  lethal  dose  in  rabbits  is  150-200  mg/kg  (0.57-0.76  milli- 
mole/kg)  and  such  values  have  generally  been  found  in  most  animals.  Dr. 
Loevenhart  courageously  ingested  a  total  of  1.3  g  within  5.5  hr  without 
the  slightest  effect.  The  lethal  dose  of  iodosobenzene  is  the  same  as  that  of 
o-iodosobenzoate,  indicating  that  the  carboxylate  group  is  not  essential  for 


726  6.    O-IODOSOBENZOATE 

the  toxicity  (Luzzato  and  Satta,  1910).  Probably  the  only  useful  role  for 
the  carboxylate  group  is  to  increase  the  solubility. 

EFFECTS   ON   SEA   URCHIN    EGG   DEVELOPMENT 

The  effects  of  0.66  mM  o-iodosobenzoate  in  sea  water  on  the  development 
of  Arbacia  eggs  was  studied  by  Runnstrom  and  Kriszat  (1952).  It  was  found 
that  fertilization  and  cleavage  proceed  quite  normally  up  to  the  blastula 
stage  (perhaps  with  a  slight  delay),  but  after  6  hr  the  controls  are  hatched 
whereas  the  treated  ones  are  not.  After  20  hr  the  controls  are  bilateral  early 
plutei,  but  80-90%  of  the  treated  larvae  are  still  within  their  membranes, 
the  formation  of  the  entoderm  being  suppressed  in  these.  The  animal  re- 
gion is  characterized  by  a  high  cylindrical  region  of  epithelium  carrying  a 
ciliary  tuft,  whereas  the  cells  at  the  vegetal  pole  are  flattened.  The  treated 
larvae  contain  no  pigment  and  the  pigment  initially  present  has  disappear- 
ed. This  effect  of  animalization  of  the  larvae  can  be  brought  about  by  other 
enzyme  inhibitors  (iodoacetate,  parapyruvate,  etc.)  and  has  been  confirm- 
ed for  o-iodosobenzoate  by  Ranzi  (1955).  If  the  larvae  after  6  hr  exposure 
to  o-iodosobenzoate  are  removed  to  normal  sea  water,  some  recovery  occurs 
and  fairly  normal  plutei  may  be  formed,  although  the  arms  are  lacking  and 
the  archenteron  shows  no  differentiation.  The  general  conclusion  was  that 
oxidation  of  certain  SH  groups  suppresses  primarily  the  differentiation  of 
the  entomesoderm. 

A  more  detailed  study  of  the  earliest  stages  of  Arbacia  egg  development 
was  made  by  Monroy  and  Runnstrom  (1952).  The  high  concentration  of 
2.64  mM  o-iodosobenzoate  does  not  prevent  the  fertilization  reaction  or  the 
formation  of  the  fertilization  membrane,  but  the  membrane  is  somewhat 
thicker  and  more  refractile  than  normally.  At  80  min  the  controls  are  in  2- 
and  4-ceU  stages  with  the  membrane  unchanged,  whereas  the  treated  eggs 
are  all  in  the  2-cell  stage  with  conspicuous  membranes.  One  hour  later  three 
fourths  of  the  treated  eggs  are  cytolyzed  with  escape  of  pigment.  The  mem- 
brane thickening  and  the  escape  of  pigment  seem  to  be  correlated.  If  the 
eggs  are  first  centrifuged,  thickening  of  the  membrane  occurs  only  at  the 
pole  where  the  pigment  is  located.  Thus  the  membrane  changes  do  not  ap- 
pear to  be  due  to  a  direct  action  of  the  o-iodosobenzoate.  Possibly  the  na- 
ture of  the  membrane  and  its  later  changes  during  development  depend  on 
substances  formed  in  the  egg  and  metabolic  inhibitors  interfere  in  the  pro- 
duction or  action  of  these  substances. 

The  exposure  of  Paracentrotus  lividus  eggs  to  0.35-0.7  mM  o-iodosoben- 
zoate does  not  affect  subsequent  fertilization  or  suppress  cleavage,  although 
hatching  is  prevented  (Hagstrom,  1963),  confirming  the  earlier  results  of 
Runnstrom  and  Kriszat  (1952)  on  Arbacia  eggs.  There  are,  nevertheless, 
differences  in  the  response.  First,  cleavage  is  somewhat  accelerated:  The 


EFFECTS    ON    BACTERIA    AND    VIRUSES  727 

controls  at  135  min  after  fertilization  are  2%  in  the  2 -cell  stage,  50%  in  the 
4-cell  stage,  and  48%  in  the  8-cell  stage,  whereas  those  treated  with  o-iodo- 
sobenzoate  are  8%  in  the  4-cell  stage  and  92%  in  the  8-cell  stage.  Second, 
there  is  no  obvious  disturbance  in  development,  e.g.,  no  evidence  of  animal- 
ization,  and  the  ciliated  embryos  inside  their  membranes  appear  to  be  nor- 
mally active.  Differentiation  in  Paracentrotus  is  thus  less  susceptible  than  in 
Arbacia  to  o-iodosobenzoate.  Higher  concentrations  of  o-iodosobenzoate  may 
produce  other  effects  on  eggs  but  whether  these  actions  are  mediated  through 
SH  group  oxidation  is  not  known.  The  eggs  of  Hemicentrotus  pulcherrimus 
and  Urechis  unicinctus  elevated  the  fertilization  membrane  when  incubated 
for  10  min  in  10  raM  o-iodosobenzoate  at  pH  4  and  then  returned  to  normal 
sea  water  (Isaka  and  Aikawa,  1963).  It  was  suggested  that  the  vitelline  and 
plasma  membranes  are  connected  by  hydrogen  bonds  and  that  o-iodosoben- 
zoate and  other  SH  reagents  react  with  SH  groups  in  the  plasma  membrane, 
weakening  these  bonds  and  allowing  separation  of  the  membranes.  Move- 
ments during  cleavage  have  been  supposed  to  involve  contractile  proteins 
as  in  muscle,  and  threads  formed  from  fibrous  proteins  obtained  from  Hemi- 
centrotus eggs  contract  when  metal  ions  (e.g.,  Mg++,  Cu++,  Cd++,  etc.)  are 
added  (Sakai,  1962).  This  contraction  is  blocked  by  5  rsxM  o-iodosobenzoate 
and  high  concentrations  of  other  SH  reagents,  indicating  that  SH  groups 
are  necessary. 


EFFECTS  ON  BACTERIA  AND  VIRUSES 

The  early  interest  in  the  antibacterial  actions  of  iodine  led  Arkin  (1911), 
in  connection  with  the  pharmacological  studies  of  Loevenhart  and  Grove 
at  Wisconsin,  to  investigate  the  effects  of  o-iodosobenzoate  and  related 
compounds  on  various  bacteria.  Eberthella  typhosa,  E.  coli,  S.  aureus,  and  B. 
pyocyaneus  are  all  killed  by  exposures  of  24  hr  to  1  mM  at  37°,  not  surpris- 
ingly. o-Iodoxybenzoate  is  even  more  potent,  but  o-iodobenzoate  does  not 
kill  even  at  10  mM.  Jahn  (1914)  found  the  growth  of  E.  coli  to  be  inhibited 
by  0.38  mM  o-iodosobenzoate,  but  not  by  38  mM  o-iodobenzoate,  indicat- 
ing the  importance  of  the  oxidative  action  and  confirming  the  results  in 
animals.  Chinard  (1942)  considered  the  possibility  of  using  o-iodosoben- 
zoate locally  in  infected  wounds.  He  observed  marked  inhibition  of  the 
growth  of  E.  coli  at  0.02  mM  with  eventual  death  of  the  bacteria  in  72  hr, 
and  death  of  hemolytic  streptococci  at  0.38  mif.  If  the  o-iodosobenzoate  is 
injected  with  these  streptococci  subcutaneously  into  mice,  no  infections  are 
seen,  but  all  the  control  mice  die.  The  flagellar  activity  of  B.  brevis  is  well 
inhibited  by  2  mM  o-iodosobenzoate  at  30  sec  and  maximally  at  5  min 
(De  Robertis  and  Peluffo,  1951).  Yeast  is  more  resistant,  since  it  requires 
3.8  mM  to  inhibit  the  growth  50%  (Loveless  et  al.,  1954).  No  analyses  at 
all  have  been  made  of  the  sites  or  mechanisms  of  action.  It  is  likely  that 


728  6.    O-IODOSOBENZOATE 

the  activity  against  bacteria  will  be  strongly  influenced  by  the  media  used 
and  the  other  conditions;  most  of  the  media  for  pathogens  contain  substances 
readily  reacting  with  o-iodosobenzoate.  Phagocytosis  of  staphylococci  and 
streptococci  by  human  leucocytes  is  stimulated  by  o-iodosobenzoate,  but 
this  is  indirect  since  it  occurs  only  in  the  presence  of  serum,  and  is  perhaps 
an  activation  of  serum  opsonin  (Arkin,  1912). 

The  psittacosis  virus  is  30-75%  inactivated  by  exposure  to  0.1  roM  o-iodo- 
sobenzoate for  1  hr  at  37°,  only  p-chloromercuribenzoate  of  aU  the  agents 
tested  being  more  potent  (Burney  and  Golub,  1948).  In  addition,  it  is  the 
most  effective  substance  in  reducing  viral  growth  in  chick  embryo  cultures 
without  inhibiting  culture  growth.  The  selectivity  is  probably  not  great 
enough  to  warrant  clinical  interest. 


CHAPTER  7 

MERCURIALS 


The  mercurials  occupy  a  rather  special  niche  in  the  subject  of  enzyme 
inhibition;  they  are  very  useful  for  demonstrating  the  presence  and  impor- 
tance of  SH  groups  in  enzyme  reactions,  but  apparently  lack  specificity 
toward  particular  enzymes  or  classes  of  enzymes.  Since  so  many  enzymes 
contain  reactive  SH  groups  at  or  near  the  active  center,  the  mercurials 
would  seem  to  inhibit  more  enzymes  than  they  leave  unaffected.  When  a 
mercurial  acts  on  living  cells,  one  cannot  state  which  enzymes  are  affected 
most  readily.  In  other  words,  they  are  reasonably  specific  with  regard  to 
the  molecular  group  attacked,  but  quite  nonspecific  at  the  enzyme  or  cel- 
lular levels.  The  mercurials  wiU,  in  addition,  react  with  nonenzymic  proteins 
and  may  modify  complex  systems  by  mechanisms  unrelated  to  metabolism. 
The  mercurials  are  thus  at  present  generally  useless  as  tools  to  study  the 
relationship  of  a  particular  enzyme  to  the  over-all  metabolism,  growth,  or 
function  of  a  cell  or  organism.  Nevertheless,  with  judicious  use,  they  may 
give  some  insight  into  the  broader  metabolic  basis  of  function,  as  in  certain 
studies  of  gastric  acid  secretion,  renal  transport,  and  mitosis.  Their  primary 
use,  however,  is  the  detection  and  titration  of  SH  groups  on  enzymes. 
They  are  often  stated  to  be  the  most  specific  SH  reagents;  this  may  be 
questioned,  but  without  doubt  they  are  among  the  most  reactive  reagents 
and  seldom  does  one  find  SH  groups  resistant  to  the  mercurials  and  capable 
of  reacting  with  other  SH  reagents.  Like  all  inhibitors,  they  are  valuable 
only  when  used  in  the  proper  system.  It  is  always  tempting  to  use  inhib- 
itors such  as  the  mercurials  which  will  almost  always  produce  definite 
effects,  but  unfortunately  the  results  usually  cannot  be  interpreted  satis- 
factorily. We  shall  emphasize  the  quantitative  side  of  mercurial  action  and 
the  inhibitions  of  pure  enzymes,  discussing  only  briefly  effects  observed  on 
complex  systems,  inasmuch  as  little  useful  information  can  be  derived 
from  this  latter  work. 

The  medical  use  of  the  mercurials  can  be  traced  back  for  over  3000  years, 
although  their  modern  therapeutic  applications  began  with  the  rediscovery 
of  the  diuretic  action  of  mercurous  chloride  in  1849  (since  then  this  action 
has  been  rediscovered  several  times),  the  demonstration  of  the  antiseptic 

729 


730  7.  MERCURIALS 

action  of  mercuria  chloride  by  Koch  in  1881,  and  the  introduction  of  or- 
ganic mercurials  for  diuresis,  antisepsis,  and  other  chemotherapeutic  pur- 
poses from  1900  to  1920.  The  marked  toxicity  of  inorganic  mercury  was 
recognized  in  antiquity  and  became  a  more  critical  problem  over  400  years 
ago,  especially  in  processes  such  as  fur  felting  for  hats  and  more  recently 
in  the  widespread  use  of  mercurials  as  plant  fungicides  for  various  rots  and 
rusts.  There  was  a  good  deal  of  experimentation  and  speculation  on  the 
nature  of  mercurial  antisepsis  between  1900  and  1940,  but  little  of  this  is 
pertinent  to  our  present  purposes.  The  early  work  was  much  concerned 
with  the  examination  of  the  validity  of  certain  vague  concepts,  such  as  the 
Arndt-Schulz  law  (which  states  that  drugs  stimulate  in  low  concentration 
and  inhibit  in  high  concentration),  oligodynamic  action,  and  the  Ostwald 
adsorption  theory.  Despite  the  fact  that  the  combination  of  mercurials  with 
thiols,  e.g.  cysteine,  has  been  known  since  1875  at  least,  investigations  on 
the  metabolic  effects  and  enzyme  inhibition  are  very  sparse  before  1930. 
Actually  the  mercurials  have  been  intensively  used  by  biochemists  for  the 
characterization  of  enzymes  for  only  the  past  several  years.  Of  the  some 
1350  publications  on  the  effects  of  mercurials  on  isolated  enzymes,  only  4% 
were  issued  prior  to  1950,  16%  from  1950  to  1956,  and  80%  from  1956 
through  1964.  By  the  time  this  volume  goes  to  press,  approximately  half 
of  the  publications  on  this  aspect  of  the  mercurials  will  have  appeared  after 
1960.  These  figures  indicate  essentially  that  each  newly  isolated  enzyme  is 
subjected  to  one  or  more  mercurials  for  the  purpose  of  detecting  SH  groups. 
One  of  the  major  aims  of  this  chapter  is  to  attempt  to  determine  the  va- 
lidity and  usefulness  of  such  determinations. 


CHEMICAL  PROPERTIES 

The  most  commonly  used  inorganic  mercury  compound  in  inhibition  work 
is  mercuric  chloride  (HgCla).  Some  fundamental  properties  of  the  Hg++  ion 
and  its  halides  are  summarized  in  Table  7-1.  It  may  be  noted  that  although 
the  linearity  of  HgXg  molecules  is  established  and  the  configuration  of 
certain  HgX4  complexes  appears  to  be  tetrahedral,  the  nature  of  the  HgClg" 
and  HgCl4=  ions  is  not  clear  and  a  planar  arrangement  is  possible.  The 
aqueous  solubility  of  HgClg  increases  with  the  concentration  of  NaCl,  KCl, 
or  other  halide  present;  thus  the  solubility  of  HgClg  in  Krebs-Ringer  me- 
dium is  around  12.5  and  in  sea  water  around  27  g/100  ml  (Barnes  and 
Stanbury,  1948).  The  deficiency  in  the  ionic  character  of  HgClg  is  indicated 
by  the  high  solubility  in  ethanol  (26.3  g/100  ml)  and  even  in  ether  (4.55 
g/100  ml).  Indeed,  HgClg  has  been  said  to  be  reasonably  lipid-soluble,  a 
fact  of  some  importance  in  considering  the  distribution  in  the  tissues.  HgBrg 
and  Hglg  are  much  less  soluble  than  HgClg  in  water  and  seem  to  have  no 
advantages  over  HgClg  in  enzyme  studies. 


CHEMICAL   PROPERTIES 


731 


Table  7-1 
Some  Properties  of  Mercury,  the  Mercuric  Ion,  and  the  Mercuric  Halides 


Radii 
Hg  atom 
Hg++  ion 

Bond  lengths 
Hg-Cl 
Hg-Br 
Hg-I 

Bond  ionic  character 
Hg-Cl  in  HgCIs, 

Bond  types 
HgCl, 
HgClr 

Electronegativity 
Hg(II) 

Solubility  in  water  (g/100  ml  solution] 
HgCl, 

Solubility  product 
HgCl, 
K,^=  (Hg++)(C1-)^ 

pH  of  saturated  solution 
HgCl, 

Redox  equilibrium 
Hg++  +  Hg(I)  ±^  Hg,++ 
K  =  (Hg,++)/(Hg++) 

Redox  potentials  {E\r,o) 
Hg2++  ±^  2  Hg++  +  2e- 
2  Hg     ±5  Hg2++  +  2e- 
Hg        ±?  Hg++    +  2e- 


1.59  A 
0.66  A 


2.20  A 
2.40  A 
2.55  A 


28% 

Linear  {sp) 
Tetrahedral   (sp') 

1.9 


6.8  (25°) 

8.9  (37.5°) 

1.06  X  10-"  (25°) 
2.95  X  10-i»   (37.5°) 


4.7  (25°) 


129.2 


-  0.92  V 

-  0.79  V 

-  0.85  V 


732  7.  MERCURIALS 

Equilibria   between    Hg++  and    Halide   Ions 

In  aqueous  solution  HgClg  does  not  dissociate  simply  into  Hg++  and  CI" 
ions,  but  forms  a  series  of  complexes,  the  relative  concentrations  of  which 
depend  on  the  CI"  concentration  and  the  pH.  The  following  species  are  the 
most  important:  Hg++,  HgCl+,  HgClg,  HgClg",  and  HgCl4=.  This  applies  to 
acid  solutions  where  hydrolysis  and  hydroxyl  complexes  can  be  ignored 
(see  next  section).  Higher  CI  complexes  with  Hg++  can  be  neglected  in 
biological  work,  as  can  univalent  Hgg^  +  and  its  complexes  (since  no  equilib- 
rium with  metallic  mercury  occurs).  Sillen  and  his  collaborators  in  Stock- 
holm have  summarized  their  extensive  investigation  of  the  halide  complexes 
of  mercury  (Sillen,  1949)  and  we  shall  follow  their  values  for  the  equilib- 
rium constants  (it  should  be  noted  that  their  work  was  done  at  25°  so  that 
small  corrections  should  be  applied  for  solutions  at  other  temperatures). 
However,  we  shall  differ  in  two  ways  from  Sillen  in  the  expression  of  the 
constants.  In  the  first  place,  we  shall  use  dissociation  rather  than  association 
constants,  in  conformity  to  the  usage  throughout  this  book.  In  the  second 
place,  we  shall  indicate  the  individual  dissociations  by  ^'s  and  the  cumu- 
lative dissociations  by  /5's,  in  conformity  with  the  usual  terminology  in 
metal-ligand  complexes  and  chelates  (Bjerrum  et  al.,  1957).  The  fundamental 
dissociations  and  their  constants  can  be  formulated  as  in  Table  7-2.  The  tight 
binding  of  the  first  two  Cl~  ions  is  evident,  but  the  next  two  are  bound  only 
weakly,  due  perhaps  to  the  change  in  bond  configuration  and  the  increasing 
negativity;  that  the  latter  is  not  a  major  factor  is  indicated  by  the  similar 
behavior  of  the  ammonia  complexes.  The  constants  for  the  Br"  and  I" 
complexes  are  much  less  than  for  CI",  i.e.,  the  former  ions  are  more  tightly 
bound  to  Hg++,  but  such  equilibria  are  seldom  of  importance  in  biological 
systems. 

The  relative  concentrations  of  the  various  complexes  depend  in  simple 
solutions  mainly  on  the  Cl~  concentration.  The  fractions  of  the  total  mer- 
cury in  particular  complexes  may  be  calculated  from  the  following  equa- 
tions: 


(7-1) 


where 


(Hg++)/(Hg,)     = 
(HgCl+)/(Hg,)   = 
(HgCl,)/(Hg,)    = 
(HgCl3-)/(Hg,)  = 
(HgClr)/(Hg,)  = 

1/A 

(Cl-)/iS.A 

(C1-)V/3.A 

(C1")VM 
(C1-)V/J4A 

(C1-)          (Cl-)= 

(Cl~)^ 

A  =  1  +  ^-—  + — - —  H —  + 


(Cb)^ 


i5i  /3,  /33  ^4 


The  CI"  concentration  varies  over  a  wide  range  in  the  media  used.  In  isolated 
enzyme  work  it  may  be  very  low  (unless  KCl  or  NaCl  is  added);  it  is  102  mM 


CHEMICAL    PROPERTIES 


733 


o 


o 


o 


Q 


^- 
II 

c 


O 


>. 


W 


Q 

^ 

w 

It 

it 

it 

it 

1 

5 

o 

1 

o 

1 

3 

+ 

+ 

+ 

+ 

+ 
+ 

be 

1 

o 

1 
a 

H    c 


734 


7.  MERCURIALS 


in  serum,  126  mM  in  Krebs-Ringer  bicarbonate  medium,  143  mM  in  Tyrode 
solution,  154  mM  in  physiological  saline,  and  515  vaM  in  sea  water.  The 
distribution  between  species  of  complexes  for  three  situations  (low,  moder- 
ate, and  high  Cl~)  is  shown  in  Table  7-3,  and  the  distribution  over  a  com- 
plete spectrum  of  Cl~  concentrations  is  illustrated  in  Fig.  7-1.  It  so  happens 


Fig.  7-1.  Curves  showing  the  distribution  of  the  different  chloride  com- 
plexes of  Hg++  with  Cl^   concentration.    (From   Sillen,   1949.) 

that,  in  most  media  used  in  ceU  and  tissue  preparations,  the  concentrations 
of  HgClg,  HgClg",  and  HgCl4=  are  roughly  equal,  whereas  in  sea  water  the 
predominant  form  is  HgCl4=.  In  the  media  for  isolated  enzyme  study,  in 
which  Cl~  is  often  low,  the  predominant  form  may  be  HgClg  or  even  HgCl+. 

Table  7-3 

Distribution  of  Mercuric  Chloride  Complexes  as  Fractions 
OF  the  Total  Mercury  in  Media  of  Different  CI"  Concentration 


Fraction 


(C1-)  =  1  ml/ 


Krebs-Ringer  medium 
(C1-)  =  126  mM 


Sea  water 
(C1-)  =  515  mM 


(Hg++)/(Hg() 

6.03  X  10-8 

1.26  X  10-12 

9.8  X  10-is 

(HgCl+)/(Hg,) 

3.31  X  10-" 

8.68  X  10-' 

2.8  X  10-« 

(HgCl,)/(Hg,) 

0.9925 

0.331 

0.0428 

(HgCl3-)/(Hg,) 

7.09  X  10-3 

0.296 

0.156 

(HgCIr)/(Hg,) 

7.09  X  10-5 

0.373 

0.801 

It  has  often  been  assumed  in  the  past  that  the  mercuric  ion  Hg++  is  the 
predominant  form  or  the  active  inhibitor,  but  it  is  now  realized  that  this 
is  not  the  case.  The  importance  of  such  complexes  in  inhibition  studies  is 
2-fold.  In  the  first  place,  the  equilibria  for  the  binding  of  mercury  to  SH 


CHEMICAL   PROPERTIES 


735 


groups  are  modified,  since  one  may  say  that  the  Cl~  ions  are  competing 
with  the  SH  groups  for  Hg++;  this  affects  the  dissociation  constants  for 
R — S — Hg+  and  R — S — Hg — S — R  complexes  (see  page  740).  In  the  sec- 
ond place,  the  penetration  of  the  inhibitor  into  cells  will  depend  on  the  rel- 
ative concentrations  of  these  complexes.  A  few  investigators  have  realized 
the  implications  of  such  complexes  and  have  attempted  to  take  into  ac- 
count the  equilibria  under  their  experimental  conditions.  Jowett  and  Brooks 
(1928)  calculated  the  relative  concentrations  of  the  complexes  in  a  0.2  milf 
solution  of  HgClg  in  Locke's  medium  in  a  study  of  the  effects  of  HgClg  on 
tissue  glycolysis  and  respiration,  and  concluded  that  the  dominant  pene- 
trating species  is  HgClg,  although  they  were  uncertain  as  to  the  form  ef- 
fective on  the  enzymes.  Barnes  and  Stanbury  (1948)  realized  that  Hg++  is 
extremely  low  in  sea  water  in  their  investigation  of  the  toxic  actions  of 
HgClg  on  a  copepod,  and  assumed  that  the  prevalent  species  were  HgClg" 


Hg" 

56 

.10" 

10 

1  Bl 

10'' 

zs 

,0-'^ 

70 

10"" 

9.8.10"'^ 

HgCI 

3.1 

10' 

5 

50 

10-^ 

1.4 

10-^ 

12 

10"^ 

28>I0'^ 

ID 

- 

, 

08 

- 

~~~^ 

-sHgClj 

^^ 

0.6 

- 

\ 

\ 

/^ 

0  4 

- 

\ 

>/ 

0.2 

FRACTION 
OF 

- 

HgClj 

-^"gcij 

^ 

-^ 

^ 

^ 

^::^ 

TOTAL 

1 

______—- 

100 


1000 

mM 


Fig.  7-2.  Fraction  of  Hg  in  various  forms  in  acid  medium  with  varying  Cl" 

concentration.  The  figures  at  the  top  give  the  concentrations  of  Hg++  and 

HgCl+  at  selected  Cl~  concentrations. 


and  HgCl4=.  Green  and  Neurath  (1953)  likewise  discounted  the  importance 
of  Hg++  in  the  inhibition  of  trypsin  in  a  medium  containing  10  mikf  CI". 
In  order  to  facilitate  estimation  of  the  relative  concentrations  of  the  com- 
plexes in  a  narrower  range  of  Cl~  concentration  as  commonly  used  in  inhi- 
bition studies.  Fig.  7-2  is  presented.  However,  before  considering  these  com- 
plexes further,  it  will  be  necessary  to  discuss  their  so-called  hydrolysis  in  a 
pH  range  around  neutrality. 


736  7.   MERCURIALS 

Equilibria  between  Hg++  and  Hydroxyl  Ions 

It  has  generally  been  assumed  that  the  Hg++  ion  is  hydrated  and  that 
this  ionizes  according  to  the  following  equations: 


Hg(H,0),++      ^  HgOH(H,0)+  +  H+ 

P^a, 

=  3.70 

HgOH(H20)+  ^  Hg(OH),          +  H- 

VK., 

=  2.60 

This  is  essentially  saying  that  the  hydrated  ion  is  a  dibasic  acid  (Hietanen 
and  Sillen,  1952).  It  is  evident  that  at  pH's  near  neutrality,  Hg(0H)2  will 
be  the  predominant  form.  For  our  purpose  and  comparison  of  these  equilib- 
ria with  those  for  Cl~,  it  might  be  better  to  express  the  reactions  as  simple 
complexing  with  OH"  ions.  Thus  p^oH  ~  10.3,  and  p^qh  —  11-4: 

(Hg++)   (OH  ) 

Hg+-      +  OH-  ±.  HgOH+  ZoH,  =       ^jrnl.^       =  ^-Z^".  =  ^  ^  1^"" 

*  {HgOH  +  )  ' 

HgOH-  +  OH-  ^  Hg(OH),       ^OH,  =  ^^^h^qh^^^  =  ^'^'^"^  =  *  ""  ^^"" 

which  may  be  compared  to  p^C^]  =  6.74,  and  p^^  =  6.48  (Table  7-2). 
Now  in  a  Cb-free  medium,  even  at  pH  5,  the  ratio  [Hg(0H)2]/(Hg++)  will 
be  5000,  and  at  pH  7  will  be  50,000,000,  so  that  Hg++  will  be  negligible. 
Although  the  affinity  of  Hg++  for  OH"  is  greater  than  for  CI",  when  CI"  is 
present  in  appreciable  concentration  (e.g.  10-150  mM)  it  will  compete  ef- 
fectively for  the  Hg++  ion  since  at  neutrality  its  concentration  will  be  10^ 
to  10®  times  that  of  OH".  Therefore  one  would  predict  that,  in  the  usual 
media  for  inhibition  studies,  the  Cl~  complexes  will  predominate  over  the 
OH"  complexes  although,  particularly  as  the  pH  is  increased  above  7,  it 
is  clear  that  complexes  of  the  type  HgCl(OH),  HgCl(0H)2",  HgCl2(0H)-, 
and  HgCl3(0H)=  may  contribute  significantly  to  the  total  population.  The 
data  given  by  Sneed  and  Brasted  (1955)  allow  one  to  calculate  the  constants 
for  the  following  equilibria: 

HgCl-     +  OH-  ^  HgCl(OH)  ^"$^pinwfi"^  =  3.2xl0-« 

[HgCl(OH)] 

HgCl3-   +  OH-  ^  HgCl3(0H)=  ^S?.r)nMfi^  =  ^"^ ^  1^"* 

[HgCl3(0H)=] 

The  affinities  of  the  CI"  complexes  for  OH"  are  thus  of  the  same  order  of 
magnitude,  and  less  than  for  the  Hg++  ion.  At  pH  7,  (HgCl+)/HgCl(OH)  = 
0.32  and  (HgCl3")/HgCl3(OH)=  -  0.5,  so  it  is  seen  that  these  OH"  com- 
plexes are  indeed  significant.  The  importance  of  these  OH"  and  mixed  com- 
plexes for  inhibition  studies  is,  of  course,  the  same  as  that  of  the  CI"  com- 
plexes, but  the  concentration  of  Hg++  wiD  be  even  less  than  calculated  in 
the  previous  section. 


CHEMICAL    PROPERTIES  737 

Complexes  of  Hg++  with   Various   Ligands 

Most  metal  ions,  including  Hg++,  form  strongly  ionic  covalent  bonds  with 
ligand  atoms  capable  of  donating  electron  pairs,  both  Cl~  and  0H~  being 
simple  examples  of  this.  We  would  expect  that  Hg++  might  complex  readily 
with  a  variety  of  substances,  many  of  which  occasionally  occur  in  media 
used  for  inhibition  studies.  It  is  usually  stated  that  Hg++  reacts  with  SH 
groups  selectively  and  that  other  groups  on  proteins  seldom  contribute  to 
the  binding;  it  is  necessary  to  look  into  this  matter  quantitatively,  and  ob- 
tain some  idea  of  the  relative  affinities  of  the  various  groups  for  Hg++. 
Complexes  of  Hg++  with  ammonia  are  well  known  so  that  binding  to  amino 
groups  might  be  predicted  and,  since  some  interaction  with  carboxylate 
groups  is  likely,  it  may  be  anticipated  that  amino  acids  would  provide  ef- 
fective ligands.  Indirect  evidence  for  such  complexes  was  obtained  by  Salle 
and  Ginoza  (1943)  by  showing  that  several  amino  acids  reduce  the  bacte- 
ricidal activity  of  HgClg.  The  minimal  lethal  concentration  of  HgClg  is  in- 
creased 6  times  by  glycine,  aspartate,  glutamate,  arginine,  and  lysine  at 
67  inM,  and  120  times  by  cysteine.  This  indicates  appreciable  complexing 
with  amino  acids  under  physiological  conditions,  although  the  reaction  with 
the  SH  group  of  cysteine  is  evidently  stronger  than  with  other  groups. 
Haarmann  (1943  a,b)  claimed  that  whereas  1  equivalent  of  Hg  is  bound  to 
certain  amino  acids  at  pH  7,  as  much  as  4  to  8  equivalents  may  be  bound 
at  pH  11,  some  loosely  and  some  tightly.  A  definitive  investigation  was 
made  by  Perkins  (1952,  1953)  and  a  number  of  stability  constants  were 
determined.  Two  major  complexes  were  assumed,  probably  with  the  follow- 
ing structures: 

OC— O  OC— O  N 

H, 

Complex  I  Complex  II 

The  composite  constant,  ^2  =  K^K^,  where  K^  and  K^^  are  defined  by  the 
following  equilibria: 


Hg++      +  AA-  ±s  HgAA+  K, 

HgAA+  +  AA-  ±^  Hg(AA)2         K, 


(Hg++)    (AA-) 
(HgAA+) 

(HgAA+)  (AA-) 
Hg(AA), 


was  determined  in  each  case,  and  these  values  are  given  in  Table  7-4  along 
with  the  dissociation  constants  for  a  number  of  ligand  complexes.  The  form 
of  the  amino  acid  necessary  for  chelation  with  Hg++  is  "OOC — E — NHg 


738 


7.  MERCURIALS 


Table  7-4 
Dissociation  Constants  for  Various  Mercuric  Complexes" 


Ligand 

Pi^i 

P^. 

VP2 

Reference  * 

Inorganic  ions 

ci- 

6.74 

6.48 

13.22 

(1) 

Br- 

9.05 

8.28 

17.33 

(1) 

I- 

12.87 

10.95 

23.82 

(1) 

OH- 

10.3 

11.4 

21.7 

(2) 

CN- 

— 

— 

34.7 

(3) 

SCN- 

— 

— 

17.4 

(3) 

Pyrophosphate 

— 

— 

17.45 

(4) 

Nitrogenous  compounds 

Methylamine 

8.6 

9.3 

17.9 

(5) 

Triethylamine 

7.8 

7.8 

15.6 

(5) 

1,2-Diaininopropane 

— 

— 

23.5 

(5) 

1 ,2,3-Triaminopropane 

19.6 

— 

— 

(5) 

Ethylenediamine 

— 

— 

23.4 

(5) 

Ethanolamine 

8.5 

8.8 

17.3 

(5) 

Diethanolamine 

7.8 

7.8 

15.6 

(5) 

Triethanolamine 

6.9 

6.2 

13.1 

(5) 

2,2'-DiaminodiethyIamine 

21.8 

— 

— 

(5) 

Triethylenetetramine 

25.3 

— 

— 

(6) 

Ammonia 

8.8 

8.7 

17.5 

(7) 

Pyridine 

5.1 

4.9 

10.0 

(3) 

Piperidine 

8.7 

8.7 

17.4 

(5) 

Imidazole 

— 

— 

16.7 

(8) 

Ethylenediaminediacetate 

9.75 

6.05 

15.8 

(5) 

Ethylenediaminetetraacetate 

22.1 

— 

— 

(5) 

Hexamethylenediaminetetraacetate 

21.4 

— 

— 

(5) 

Amino  acids 

Glycine 

10.3 

8.9 

19.2 
18.2 

(5) 
(9) 

Glycylglycine 

— 

— 

12.4 

(9) 

Alanine 

— 

— 

18.4 

(9) 

Leucine 

— 

— 

17.5 

(9) 

Proline 

— 

— 

20.5 

(9) 

CHEMICAL   PROPERTIES  739 

Table  7-4  (continued) 


Ligand 

pK, 

P^2 

P^2 

Reference  * 

Serine 

_^ 

_ 

17.5 

(10) 

Tyrosine 

— 

— 

17.1 

(10) 

Arginine 

— 

— 

17.4 

(10) 

Histidine 

— 

— 

21.2 

(8) 

Methionine 

6.52 

4.93 

11.45 

(11) 

Ethionine 

7.25 

5.92 

13.17 

(11) 

Cysteine 

14.21 

— 

— 

(11) 

<S-Methylcysteine 

7.20 

5.81 

13.01 

(11) 

Purines  and  pyridimidines 

Adenine 

— 

— 

11.5 

(12) 

Adenosine 

— 

— 

8.5 

(12) 

Thymine 

— 

— 

21.2 

(12) 

Thymidine 

— 

— 

21.2 

(12) 

Cytosine 

— 

— 

10.9 

(12) 

Miscellaneous 

Acetate 

4.0 

— 

— 

(13) 

Cyclohexene 

4.3 

— 

— 

(5) 

Penicillamine 

16.15 

— 

— 

(11) 

"  (^2  is  the  cumulative  dissociation  constant  for  HgLj,  and  is  K^^K^ 
*  References: 

(1)  Sillen  (1949). 

(2)  Hietanen  and  Sillen  (1952). 

(3)  Simpson  (1961). 

(4)  Yamane  and  Davidson  (1960). 

(5)  Bjerrum  et  al.  (1957). 

(6)  Chaberek  and  Martell   (1959). 

(7)  Bjerrum  (1941). 

(8)  Brooks  and  Davidson  (1960). 

(9)  Perkins  (1952). 

(10)  Perkins   (1953). 

(11)  Lenz  and  Martell  (1964). 

(12)  Ferreira  et  al.  (1961). 

(13)  Gurd  and  WUcox  (1956). 


740  7.   MERCURIALS 

and,  in  the  calculations  of  the  constants,  only  that  fraction  of  the  amino 
acid  at  the  pH  used  was  considered. 

The  fact  that  simple  amines  complex  with  Hg++  to  approximately  the 
same  degree  as  the  amino  acids  indicates  that  the  amino  group  is  the  im- 
portant ligand,  the  carboxylate  group  perhaps  contributing  slightly  to  the 
stability.  Ring  nitrogen  atoms  are  probably  not  as  effective  as  amino  groups. 
Hg++  reacts  with  both  the  amino  and  imidazole  groups  of  histidine,  but 
more  tightly  with  the  former,  the  pj^'s  being  10.6  and  7.5,  respectively 
(Simpson,  1961).  The  effect  of  the  pH  on  the  stability  of  these  complexes 
is  well  illustrated  by  the  constants  for  the  following  equilibria  with  histidine 
(Brooks  and  Davidson,  1960): 

Hg++  +  2  hist-  ±5  Hg  (hist)2  p/ff^  =  21.2 

Hg++  +  hist-  +  H-hist  i?  Hg  (hist)  (H-hist)+  Pi92  =  18.4 

Hg++  +  2  H-hist  15  Hg  (H-hist)2++  pp^  =  15.0 

where  hist  designates  the  ~00C — R — NHg  form  and  H-hist  the  OCC — 
R — -NH3+  form.  These  complexes  with  histidine  were  assumed  to  be  linear 
and  it  was  claimed  that  chelation  must  play  only  a  small  role  in  Hg++ 
complexes  due  to  the  tendency  of  Hg++  to  form  linear  complexes.  This 
brings  up  an  interesting  point  of  importance  in  understanding  the  reactions 
of  HgClg  with  proteins  and  enzymes.  Certainly  some  of  the  most  stable 
complexes  of  Hg++  —  as  with  1,2,3-triaminopropane,  triethylenetetramine, 
and  EDTA  —  must  be  chelates  and  nonlinear,  and  it  is  also  well  known 
that  Hg++  reacts  with  dimercaprol  (BAL)  to  form  a  ring  with  the  two  SH 
groups.  Whether  chelation  is  or  is  not  important  in  any  case  probably  de- 
pends on  several  factors,  such  as  the  spatial  arrangement  of  the  ligand 
groups,  the  intrinsic  affinity  of  the  Hg++  for  these  groups,  and  the  entropy 
changes  accompanying  the  formation  of  the  complex.  Certainly  the  third 
and  fourth  ligands  generally  add  to  the  HgLg  complex  much  less  readily 
than  the  first  two,  as  we  have  seen  for  CI".  This  is  also  true  for  ammonia, 
the  successive  constants  being  given  by  pifj  =  8.8,  piiTg  =  8.7,  pK-^  =  1.0, 
and  TpK^  =  0.78.  Thus  the  formation  of  HgLg  and  UgL^  type  of  complexes 
must  involve  some  additional  factors  increasing  the  stability. 

In  any  event  it  is  clear  that  the  complexes  of  Hg++  with  amino  acids 
and  many  other  compounds  are  stable  enough  so  that,  when  these  sub- 
stances are  present  in  the  media  used  for  the  study  of  inhibition,  a  signifi- 
cant fraction  of  the  Hg  may  be  in  the  form  of  such  complexes.  For  example, 
it  may  be  calculated  for  a  solution  containing  1  mM  glycine  and  a  total 
concentration  of  Hg  of  0.1  mM  that  (Hg++)  =  6.3  x  lO-i^  M.  If  the  Cl" 
concentration  is  appreciable,  this  will  reduce  the  binding  to  these  other 
ligands.  One  may  visualize  the  situation  somewhat  as  follows.  In  most 
media  there  will  be  several  substances  —  Cl~,  OH",  buffers,  amino  acids, 
substrates,  etc.  —  capable  of  complexing  with  Hg++.  The  end  result  will 


CHEMICAL    PROPERTIES  741 

be  that  the  Hg++  concentration  will  be  extremely  low,  the  Hg  being  parti- 
tioned between  numerous  complexes  of  different  types,  each  one  reducing 
to  some  extent  the  reaction  of  Hg  with  enzymes.  It  must  be  clearly  under- 
stood that  when  inhibition  is  stated  below  to  be  by  HgCla  or  Hg(N03)2  or 
Hg  acetate,  it  refers  only  to  what  was  added  and  not  to  the  dominant 
form  present  or  the  active  inhibitor. 

Complexes  of  Hg++  with  Nucleotides  and  Nucleic  Acids 

Complexes  between  Hg++  and  certain  purines  and  pyrimidines,  especially 
thymine,  are  quite  stable  (Table  7-4)  (Katz,  1962),  and  complexes  with 
phosphates  are  probably  formed  readily;  thus,  one  would  expect  nucleotides 
and  nucleic  acids  also  to  bind  Hg++  rather  well.  Inagaki  (1940)  found  var- 
ious nucleotides,  such  as  AMP,  GMP,  and  IMP,  to  be  precipitated  by  mer- 
curic compounds,  but  unfortunately  no  further  work  has  been  done  on  these 
complexes.  One  would  like  to  know  the  nature  and  extent  of  the  interactions 
of  mercurials  with  ATP,  NAD,  FAD,  and  related  substances.  However, 
the  studies  of  the  reactions  between  Hg++  and  nucleic  acids  have  recently 
been  accelerated,  and  it  is  obvious  that  the  results  could  be  very  important 
in  understanding  the  effects  of  mercurials  on  cellular  growth  and  prolifera- 
tion. Hg++  has  been  found  to  complex  with  nucleic  acids  from  thymus 
(Katz,  1952),  plants  (Trim,  1959),  pneumococci  (Dove  and  Yamane,  1960), 
and  tobacco  mosaic  virus  (Katz  and  Santilli,  1962  b).  The  general  effects 
on  the  nucleic  acids  may  be  summarized  briefly  as  follows:  a  decrease  in 
viscosity;  an  increase  in  turbidity,  sedimentation  constant,  aggregation, 
and  the  dimer :  monomer  ratio;  an  increase  in  the  flexibility  of  the  chains 
with  the  assumption  of  a  more  compact  configuration;  and  a  change  in  the 
ultraviolet  absorption  spectra  (Katz,  1952;  Thomas,  1954;  Yamane  and 
Davidson,  1961).  It  was  originally  believed  that  the  complexing  is  with  the 
phosphate  groups,  but  the  nature  of  the  absorption  spectrum  changes,  the 
stoichiometry  of  the  reactions,  and  the  release  of  H+  indicate  that  the  bases 
are  the  sites  of  binding,  the  Hg:base  combining  ratio  being  1  :  2  in  most 
cases,  Hg  apparently  bridging  the  double  strands  of  the  DNA  helix  (Katz, 
1962).  The  single-stranded  tobacco  mosaic  virus  RNA,  however,  gives  a 
combining  ratio  of  1  :  1,  as  expected  (Katz  and  Santilli,  1962  a).  It  may  be 
noted  that  the  combining  ratio  is  1  :  2  for  guanine  oligoribonucleotides,  such 
as  GpCxpG  (Lipsett,  1964).  These  complexes  are  usually  completely  reversi- 
ble upon  adding  various  Hg++  complexers  (Cl~,  cyanide,  EDTA,  or  thiols), 
and  indeed  the  pneumococcal  transforming  DNA  after  demercuration  retains 
aU  of  its  activity  (Dove  and  Yamane,  1960),  and  the  tobacco  mosaic  virus 
after  removal  of  the  Hg++  regains  its  infectivity'  (Singer  and  Fraenkel- 
Conrat,  1962),  these  observations  indicating  that  the  original  configurations 
of  the  nucleic  acids  can  be  restored  despite  the  apparently  marked  struc- 
tural modifications  occurring  during  reaction  with  Hg++. 


742  7.  MERCURIALS 

Organic  Mercurials 

It  will  be  convenient  to  discuss  some  of  the  general  properties  of  the  or- 
ganic mercurials  before  coming  to  the  important  problem  of  the  reaction  of 
mercurials  with  SH  groups.  Many  organic  mercurials  were  developed  for 
chemotherapy,  disinfection,  and  diuretic  activity  and,  although  some  of 
these  have  been  occasionally  used  in  inhibition  work,  mercurial  inhibitors 
are  generally  simpler  structurally.  The  chemical  properties  and  structure- 
action  relationships  will  be  taken  up  later  for  the  antiseptics  (page  970) 
and  diuretics  (page  917).  The  aryl  mercurials  (such  as  p-chloromercuri- 
benzoate)  were  introduced  as  enzyme  inhibitors  by  Hellerman  (1937)  and 
the  alkyl  mercurials  (such  as  methylmercuric  chloride)  as  protein  reactants 
by  W.  L.  Hughes  (1950).  The  accompanying  formulas  are  for  the  ions  which 


V\        A^Hg  H3C-Hg 


Phenylmercuric  ion  Methylmercuric  ion 


^-Mercuribenzoate  ion  /'-Mercuriphenylsulfonate  ion 

eventually  complex  with  SH  groups  and  other  ligands.  They  are  used  as 
chlorides,  hydroxides,  nitrates,  or  acetates  but,  once  they  have  been  added 
to  the  medium,  the  ions  as  formulated  complex  with  various  ligands  which 
may  be  present,  essentially  as  Hg-^+  does.  Thus  it  is  correct  to  speak  of 
phenylmercuric  acetate  or  p-chloromercuribenzoate  as  the  mercurial  used, 
but  this  terminology  does  not  give  an  accurate  representation  of  the  forms 
present  in  solution.  For  example,  it  makes  no  difference  in  the  final  inhi- 
bition whether  one  uses  p-chloromercuribenzoic  acid  or  sodium  p-hydro- 
xymercuribenzoate,  the  two  forms  of  the  p-mercuribenzoate  ion  commonly 
available.  For  the  sake  of  compression,  we  shall  use  the  abbreviations  shown 
in  the  following  tabulation  in  the  remainder  of  this  chapter. 


Ion  Abbreviation 

Phenylmercuric  PM 

Methylmercuric  MM 

p-Mercuribenzoate  p-MB 

p-Mercuriphenylsulfonate  p-MPS 


CHEMICAL   PROPERTIES  743 

It  will  be  useful  to  summarize  briefly  some  of  the  important  differences 
between  the  organic  mercurials  and  HgClg. 

(a)  Functionality.  HgClg  is  bifunctional  in  the  sense  that  it  can  react 
with  two  ligands  to  form  L-Hg-L  complexes,  whereas  the  organic  mercu- 
rials are  monofunctional  in  that  they  can  react  with  only  one  ligand  to 
give  R-Hg-L.  Hg++  can  also  form  cyclic  mercaptides  with  two  adjacent 
SH  groups  but  the  organic  mercurials  cannot.  These  differences  are  often 
very  important  in  the  reactions  with  thiols  and  enzymes,  and  in  fact  one 
of  the  major  reasons  for  the  preference  of  many  investigators  for  the  or- 
ganic mercurials,  especially  in  the  quantitative  titration  of  SH  groups,  is 
their  monofunctional  nature. 

(b)  Aqueous  solubility.  The  organic  mercurials  are  less  soluble  than  HgClg 
and  occasionally  this  has  created  a  problem  if  higher  concentrations  are 
required.  However,  in  most  titration  or  inhibition  studies,  the  concentra- 
tion required  is  seldom  over  1  mM,  and  this  can  be  easily  attained  in  most 
cases.  Phenylmercuric  chloride  is  soluble  to  the  extent  of  only  0.16  toM  in 
distiUed  water,  but  a  good  deal  more  soluble  in  salt  solutions,  and  the  in- 
troduction of  anionic  groups  increases  the  solubility.  It  is  well  known  that 
23-chloromercuribenzoic  acid,  which  for  many  years  was  the  only  commonly 
used  mercurial,  does  not  readily  go  into  solution  at  neutral  pH.  It  is  there- 
fore usual,  to  dissolve  it  in  dilute  KOH  or  NaOH  solutions  (0.01-0.05  M) 
and  adjust  to  the  desired  pH  with  HCL*  However,  the  sodium  salt  is  now 
commercially  available  and  dissolves  readily.  The  phenylsulfonic  acid  is 
also  more  soluble  than  the  benzoic  acid  derivative.  The  solubility  will  be 
determined,  as  with  HgClj,  by  the  concentrations  of  various  complexing 
substances  in  the  medium;  thus  the  solubility  is  reasonably  high  in  most 
physiological  media  containing  over  100  mM  Cl~,  or  various  other  ions 
such  as  pyrophosphate  or  sulfate,  concentrations  around  10  mM  of  mer- 
curial being  readily  obtained. 

(c)  Lipid  solubility.  The  unsubstituted  alkyl  and  aryl  mercurials  are  more 
soluble  than  the  Hg++  ion  and  its  complexes  in  lipids.  Hughes  (1957)  has 
estimated  that  the  simple  alkyl  mercurials  are  around  100  times  more  sol- 
uble in  lipids  than  in  water.  This  property  wiU  presumably  allow  the  organ- 
ic mercurials  to  penetrate  more  readily  than  inorganic  mercury  into  cells 
and  tissues,  and  evidence  for  this  is  provided  by  the  greater  central  nervous 
system  toxicity  of  the  organic  mercurials  (page  951).  In  this  connection, 

*  Although  I  have  no  definite  evidence  that  strongly  alkaline  media  are  detrimental 
to  p-MB,  I  would  prefer  not  to  use  1  M  NaOH  solution  to  dissolve  the  material,  as 
has  been  done  by  some  (e.g.,  Snodgrass  et  al.,  1960),  since,  as  we  shall  see,  the  C — Hg 
bond  is  weak  and  dissociation  is  a  possibiltiy,  and,  furthermore,  such  strongly  alkaline 
solutions  are  not  necessary.  For  most  work  it  is  satisfactory  to  dissolve  p-chloromer- 
curibenzoic  acid  in  0.02  M  KOH  or  NaOH  at  approximately  2  mg/ml  or  7.4  mM. 


744  7.  MERCURIALS 

if  maximal  penetration  is  desired,  it  is  probably  best  to  use  the  alkyl  or 
the  unsubstituted-phenylmercurials,  since  the  C00~  and  SO3"  groups  will 
reduce  the  permeability. 

(d)  Configuration.  Although  HgClg  is  linear,  as  shown  by  Raman  spectra 
and  electron  diffraction,  the  organic  mercurials  for  some  reason  are  ap- 
parently not.  Dipole  moment  studies  (e.g.,  for  phenylmercuric  chloride, 
ju  =  2.99)  suggest  that  the  angle  of  the  C — Hg — X  bonds  is  around  130° 
or  higher  (Sipos  et  al.,  1955).  The  moment  is  directed  as  foUows: 

+       - 
R  -  Hg  -  X 

(e)  Molecular  size.  The  organic  mercurials  are,  of  course,  larger  than  the 
simple  complexes  of  Hg++.  This  may  be  of  importance  in  the  reaction  with 
the  SH  groups  of  proteins  and  enzymes,  since  steric  factors  may  impede  the 
approach  of  the  mercurial  to  SH  groups  not  exposed  on  the  surface.  A  re- 
duction in  volume  was  one  of  the  reasons  for  the  introduction  of  the  alkyl 
mercurials  by  W.  L.  Hughes  (1950).  In  addition,  penetration  into  cells  will 
depend  to  some  extent  on  the  molecular  size.  Other  factors  will  be  discussed 
relative  to  enzyme  inhibition. 

(f )  Affinities  for  ligands.  The  organic  mercurial  ions  in  solution  tend  to 
complex  with  various  ligands  in  the  same  way  as  the  Hg++  ion,  forming 
complexes  of  the  type  R-Hg-L.  The  affinities  seem  to  be  somewhat  less 
for  the  organic  mercurial  ions  than  for  Hg++,  although  very  few  have  been 
studied.  Simpson  (1961)  gave  the  dissociation  constants  for  MM  complexes 
(Table  7-5),  and  generally  the  pii's  are  around  1.3  units  less  than  for  the 
Hg++  complexes.  Nevertheless,  the  affinities  are  of  sufficient  magnitude  so 
that  at  pH  7  there  is  perhaps  500  times  as  much  MM-OH  as  MM+,  and  if 
much  Cl~  is  present  there  may  be  100  times  as  much  MM-Cl  as  MM-OH 
(Hughes,  1957).  Rowland  (1952)  determined  the  equilibrium  constant, 
ii:=(RHg-OH)  (H+)  (Cl-)/(RHg-Cl),  for  a  variety  of  diuretic  mercurials, 
and  found  a  mean  value  for  ])K  of  9.9,  so  that  at  pH  7  and  (Ch)  =  100  mM 
the  ratio  (RHg-Cl)/(RHg-OH)  is  around  100  as  for  MM.  Ledoux  (1953) 
reported  the  interaction  of  p-MB  with  nucleic  acid  and  from  the  spectral 
changes  assumed  a  complex  to  be  formed  with  the  carbonyl  group  of  pyri- 
midines.  Various  complexes  of  PM  and  amino  acids  were  obtained  by  Smalt 
et  al.  (1957)  but  were  claimed  to  dissociate  rather  readily.  The  remarkably 
tight  complex  of  PM  and  thyroxine  has  been  investigated  by  Frieden  and 
Naile  (1954).  One  must  thus  assume  that  in  solution  in  most  physiological 
media  the  organic  mercurials  will  exist  in  a  variety  of  complexes,  and  that 
this  will  be  an  important  factor  in  determining  the  degree  of  reaction  with 
proteins  and  enzyme  SH  groups. 

One  characteristic  of  the  organic  mercurials  is  the  weakness  of  the  C — Hg 
bond,  the  energy  of  which  is  only  15-19  kcal/mole  (CottreU,  1954),  so  that 


CHEMICAL   PROPERTIES  745 

Table  7-5 
Dissociation  Constants  for  Complexes  with  Methylmekcuric  Ion  " 

Ligand  pK 

CI-  5.45 

Br-  6.7 

I-  8.7 

OH-  9.5 

CN-  14.2 

SCN-  6.1 

Acetate  3 . 6 

Phenolate  6.5 

HEDTA^  6.2 

Ammonia  8.4 

Pyridine  .4.8 

Imidazole  7 . 3 

Histidine  (NHj  group)  8.8 

Histidine  (imidazole  group)  6.4 

"  From  Simpson  (1961). 

inorganic  Hg  may  be  split  off  more  readily  than  is  usually  supposed.  The 
exchange  reaction: 

Hg2«3Cl2  +  -OOC— 97— Hg+    ±?   HgCl^  +  -OOC— (p— Hg2«^  + 

is  fairly  fast,  the  rate  constant  being  5.4  liters  mole^^sec"^  at  25°  with  an 
activation  energy  of  12  kcal/mole  (Cerfontain  and  van  Aken,  1956),  and 
this  indicates  the  instability  of  the  C — Hg  bond.  It  is  possible  to  produce 
Hg203-labeled  p-MB  by  this  reaction.  This  problem  has  assumed  a  good 
deal  of  importance  in  diuretic  action  and  will  be  discussed  more  fully  in 
this  connection. 

The  synthesis  of  p-MB  has  been  described  by  Whitmore  and  Woodward 
(1941).  It  may  be  purified  by  repeated  solutions  in  dilute  NaOH  and  preci- 
pitations with  excess  HCl  (Boyer,  1954).  For  the  accurate  titration  of  SH 
groups  it  is  suggested  that  the  purity  of  the  2>-MB  be  checked  by  iodometric 
titration  or  spectrophotometrically  by  absorption  measurement  at  232  m// 
at  pH  7  (£.v  =  1-69  X  10*)  or  234  m//  at  pH  4.6  (f^  =  1-74  x  10*).  The 
stability  of  39-MB  solutions  has  not  been  determined  quantitatively,  but 
Cunningham  et  al.  (1957)  found  that  heating  to  80o-82o  for  90  min  with 
various  buffers  and  at  different  pH's  does  not  destroy  more  than  2-4%, 
and  MacDonnell  et  al.  (1951)  stated  that  solutions  are  stable  for  a  month 
at  room  temperature.  Nevertheless,  I  would  advise  making  solutions  daily 
for  accurate  work. 


746  7.  MERCURIALS 

Reactions  of  the  Mercurials  with  SH  Groups 

The  various  types  of  mercaptide  which  can  be  formed  are  indicated  in 
the  accompanying  tabulation.  The  complexes  formed  under  particular  cir- 


Monofunctional 

Bifunctional 

organic  mercurials 

inorganic  mercurials 

R'-Hg 

Hg-^ 

Monothiols 

R-S-Hg— R' 

R-S— Hg' 

R— SH 

R— S-Hg-S— R 

<;> 

Dithiols 

S-Hg 

"■SH 

S-Hg-R' 
R 
^S— Hg— R' 

S-Hg 

S-Hg-S^ 
R                           R 
^S-Hg— S^ 

(—R— S-Hg-S— ),, 

cumstances  will  depend  on  the  relative  concentrations  of  thiol  and  mercurial, 
the  presence  of  ligands  capable  of  complexing  with  the  mercurials,  the  spa- 
tial arrangement  of  the  SH  groups,  the  pH,  and  the  nature  of  the  R  and  R' 
groups.  Monofunctional  mercurials,  such  as  p-MB  and  PM,  react  with 
cysteine,  glutathione,  and  2-mercaptoethanol  in  1  :  1  molar  ratio  to  form 
R — S — Hg — R'  type  of  complexes  (Benesch  and  Benesch,  1952;  Hoch  and 
Vallee,  1960).  These  reactions  can  be  conveniently  followed  polarographi- 
cally.  Reactions  with  dithiols  are  more  complex.  The  dimercaptide  formed 
from  PM  and  dimercaprol  (BAL)  is  insoluble,  but  if  hydrophilic  groups 
occur  on  the  mercurial  the  product  is  usually  soluble.  However,  the  in- 
stability of  the  C — Hg  bond  may  allow  further  reaction  to  form  the  cyclic 
mercaptide,  as  occurs  with  mersalyl  (Benesch  and  Benesch,  1952).  This 
reaction,  where  R  =  — CH2CONH — 9? — OCHgCOO",  leads  to  the  splitting 


CHp— OH  CHp— OH 


I  /OCH3 

CH— S-Hg-CHaCH^   _  +  H 

I  ^ 

I  /OCH3 

CH2— S— Hg— CH2CH 

^R" 


Hg— CH2CH(OCH3)R 

CH2=:CHR" 

CH.OH 


CHEMICAL   PROPERTIES  747 

off  of  one  mersalyl  and  the  formation  of  inorganic  mercury  from  the  other, 
essentially  the  reverse  of  the  reaction  whereby  the  mercurial  diuretics  are 
synthesized  (the  oxymercuration  of  alkenes).  Such  reactions  presumably  do 
not  occur  with  the  simpler  organic  mercurials.  It  has  also  been  shown  that 
diphenethjTiyl  mercury  and  glutathione  react  to  form  GS — Hg — SG  and 
phenylacetylene  (Tanaka,  1961).  Mercurials  of  the  type  R2Hg  might  be 
expected  to  be  unreactive  with  SH  groups;  inasmuch  as  they  are  quite 
toxic,  Webb  et  al.  (1950)  studied  their  reactions  and  found  that,  although 
most  thiols  are  not  attacked,  dithiozne  is  reacted  as  follows: 

R— Hg— R  +  R'— SH  ->  R— Hg— S— R'  +  RH 

This  type  of  cleavage  of  the  C — Hg  bond  occurs  at  physiological  temperature 
and  pH. 

The  primary  products  of  the  reactions  between  HgClg  and  cysteine,  gluta- 
thione, thioglycolate,  and  other  monothiols  are  the  dimercaptides  of  the 
type  R — S — Hg — S — R,  and  it  is  difficult  to  study  the  initial  formation  of 
a  monomercaptide  R — S — Hg+  (Shinohara,  1935;  Stricks  and  Kolthoff, 
1953;  Stricks  et  al.,  1954).  The  reactions  between  HgClg  and  dithiols  are 
complex  and  several  types  of  mercaptide  may  be  formed,  as  indicated  in 
the  tabulation  above.  The  occurrence  of  cyclic  mercaptides  and  polymer- 
captide  linear  complexes  will  depend  mainly  on  the  spatial  configuration  of 
the  SH  groups.  The  pH  apparently  plays  some  role  in  determining  the  na- 
ture of  the  complexes,  since  as  the  pH  increases  above  2.5,  more  of  the 
forms  Hg2(SG)2  and  Hg3(SG)2  appear  (Kolthoff  et  al.,  1954). 

We  next  turn  to  the  problem  of  the  stability  of  mercaptides  and  it  is  im- 
portant to  establish  the  dissociation  constants  for  the  fundamental  com- 
plexes formed  in  simple  reactions,  such  as  the  following: 

R— S-  +  Hg++  ±^  R— S— Hg+ 
R— S-  +  R'— Hg+  ±5  R— S— Hg— R' 

Table  7-6  shows  a  few  of  the  recently  determined  constants  for  Hg++  and 
MM.  If  one  assumes  that  -pK^  and  p^g  ^I's  similar  in  magnitude,  which  is 
reasonable,  it  is  seen  that  pJ^j  (which  applies  to  the  reactions  above)  lies 
between  20  and  22,  with  a  mean  value  of  21.3.  This  range  may  thus  be 
taken  provisionally  as  indicating  the  usual  affinity  between  mercurials  and 
thiols  in  the  absence  of  competing  protons  and  ligands.  Comparison  of  these 
values  with  those  in  Tables  7-4  and  7-5  shows  that  the  affinity  of  mercurials 
for  SH  groups  is  far  greater  than  for  any  other  single  ligands,  and  that,  in  a 
mixture  of  thiols  and  various  other  complexing  ligands,  a  mercurial  will  be 
predominantly  associated  with  the  thiols.  The  variations  of  pK  with  the 
temperature  and  the  ionization  of  auxiliary  groups  on  the  thiol  are  shown, 
as  summarized  from  the  studies  of  Stricks  and  Kolthoff  (Table  7-7).  Proto- 


748 


7.  MERCURIALS 


Table  7-6 
Some  Dissociation  Constants  for  Simple  Mercaptides' 


Mercurial 

Thiol 

pA', 

Vth 

Reference 

Hg++ 

Cysteine^ 

20.1 

_ 

Simpson  (1961) 

20.5 

— 

Perkins  (1953) 

— 

43.57 

Stricks  and  Kolthoff  (1953) 

Glutathione- 

— 

41.58 

Stricks  and  Kolthoff  (1953) 

Thioglycolate~ 

— 

43.82 

Stricks  et  al.  (1954) 

Methyl-Hg+ 

Cysteine" 

15.7 

— 

Simpson  (1961) 

Human  seralbumin 

22.0 

— 

Hughes  (1957) 

Bovine  seralbumin 

22.6 

— 

Hughes  (1957) 

Bovine  HbOj 

22.1 

— 

Hughes  (1957) 

Bovine  HbCO 

22.6 

— 

Hughes  (1957) 

"  The  p/^i's  for  the  mercaptides  with  seralbumin  and  hemoglobin  have  been  recal- 
culated (page  755),  assuming  pK„  for  the  SH  groups  to  be  8.7  and  the  pK  for  complex- 
ing  with  I  to  be  8.7.  These  values  should  not  be  considered  as  accurate  because  of 
the   assumptions   involved. 

Table  7-7 

Dissociation  Constants  for  Mercaptides  with  Hg++  Illustrating 
THE   Effects   of  pH  and  Temperature 


Cysteine 


Constant" 

12° 

25° 

pA'i 

41.82 

40.25 

pK, 

45.27 

43.60 

P^3 

45.40 

43.57 

pA% 

7.39 

7.10 

VK, 

10.72 

10.48 

Glutathione 
12° 


42.29 
43.54 
43.47 

7.97 
9.28 


25° 


40.96 

41.92 

41.58 

7.85 

9.15 


Thioglycolate 


12° 


45.66 
45.85 
45.52 


25° 


44.31 
44.33 
43.82 


"  The  constants  are  defined  as  follows: 


K^ 


(Hg++)  (R)^ 
(HgR2++) 


K,= 


(H+)  (HgR,+  ) 
(HgR,H++)  ' 


(Hg++)  (R)  (R-) 
(Hgi22+) 

(H+)  (HgR,) 
(HgR,H+) 


K.  = 


(Hg++)  (R-)^ 
(HgR,) 


where  R  =  +H3N — X — S~  and  R~  =  N2H — X — S  for  cysteine  and  glutathione,  and 
R  =  HOOC— X— S-  and  R-  =  "OOC— X— S"  for  thioglycolate.  (From  Stricks  and 
Kolthoff,  1953;  Stricks  et  al,  1954.) 


CHEMICAL   PROPERTIES  749 

nation  of  an  amino  group  on  cysteine  or  glutathione  reduces  somewhat  the 
affinity  of  the  thiol  for  Hg++,  as  might  be  anticipated,  while  ionization  of 
the  thioglycolate  carboxyl  groups  has  little  effect.  With  the  constants  in 
Table  7-7  it  is  possible  to  predict  the  relative  concentrations  of  the  various 
species  present;  when  complexing  ligands  are  in  significant  concentration, 
appropriate  corrections  must  be  made  (page  737).  The  variations  with  tem- 
perature allow  the  calculation  of  certain  basic  thermodynamic  parameters. 
For  the  formation  of  the  mercaptides,  /iF  is  —  55  to  —  59  kcal/mole  and 
the  entropy  changes  are  positive  and  usually  rather  large;  for  the  equilibria 
expressed  by  y>Ki  (see  legend  in  Table  7-7),  AS  is  +27  cal/deg  for  cysteine, 
-|-  54  cal/deg  for  glutathione,  and  +  68  cal/deg  for  thioglycolate. 

We  must  now  inquire  into  the  effect  of  pH  on  mercaptide  formation  and 
particularly  consider  the  reactions  with  the  SH  group  and  the  ionized  S^ 
group.  Taking  the  two  following  equilibria: 


R— Hg+  +  R'— S-  ±^  R— Hg— S— R'  K^ 

R— Hg+  +  R'— SH  ±^  R— Hg— S— R'  +  H+         K^„ 


(R— Hg^)(R--S^) 
(R— Hg— S— R') 

(R-Hg+)  (R--SH) 
(R_Hg-S-R')(H+) 


and  the  ionization  of  SH  —  i.e.,  K,,  -  (H+)  (R'— S-)/(R'— SH)  —  it  is 
easy  to  show  that: 

pATg  =  pK^^  +  pK,  (7-2) 

The  p^^  for  SH  groups  varies  from  7  to  10,  and  in  Volume  I  a  mean  value 
of  8.7  was  assumed  for  protein  SH  groups.  In  any  event,  pKf-  and  p^sh 
will  differ  quite  markedly.  This  is,  of  course,  essentiallj-  a  competition  be- 
tween H+  ions  and  the  mercurial  for  the  S~  group.  At  physiological  pH,  SH 
will  predominate  over  S~,  and  the  apparent  pK  for  mercaptide  dissociation 
will  be  smaller  than  those  given  in  Tables  7-6  and  7-7. 

Just  as  H+  competes  with  the  mercurial  for  the  S"  groups,  so  various 
complexing  ligands  may  compete  with  the  S~  group  for  the  mercurial. 
Despite  the  fact  that  the  p/^'s  for  thiols  are  much  greater  than  for  most 
other  ligands,  a  very  significant  effect  on  the  equilibrium  may  be  exerted. 
Let  us  write  for  the  usual  reaction  of  mercaptide  formation  in  physiological 
media: 

R— Hg— X  +  R'— SH  ±i  R— Hg— S— R'  +  X"  +  H+ 

where  X  represents  some  ligand  such  as  CI"  or  0H~.  The  equilibrium  is 
given  by: 

(R— Hg— X)  (R'— SH) 


K 


(R_Hg-S-R')  (X-)  (H+) 


750  7.  MERCURIALS 

If  we  designate  the  equilibrium  with  X  by  ^^=(R— Hg+)  (X-)/(R— Hg— X), 
this  taken  in  conjunction  with  the  expressions  for  K^,  K„,  and  K  leads  to: 

pA'  =  pAg  -  pA„  -  vK,  (7-3) 

Thus  in  any  experimental  situation  the  observable  p^  will  be  less  than  the 
true  p/Cg  for  the  reaction  of  R — Hg+  and  S~  by  the  sum  of  ^K„  and  \>K^, 
each  of  these  expressing  the  competitions  involved.  Thus  in  the  work  of 
Hughes  (1957)  on  the  reaction  of  human  seralbumin  with  MM,  a  p^  of  4.6 
was  found;  if  piif^  =  8.7  and  ^^K^  —  8.7  (Table  7-5),  one  may  calculate  p^g 
to  be  22.0,  X  being  I"  in  this  case. 

The  rates  of  mercaptide  formation  increase  with  pH  as  would  be  expected 
if  the  reactive  form  of  the  thiol  is  the  ionized  R' — S~.  It  is  not  so  easy  to 
decide  on  the  reactive  forms  of  the  mercurial.  It  seems  unlikely  that  the 
Hg++  or  R — Hg+  ions  are  the  only  reactive  species  because  of  their  ex- 
tremely low  concentrations  in  most  physiological  media,  and  it  is  possible 
that  the  S^  group  makes  a  sideways  attack  on  the  Hg  atom  utilizing  a 
pair  of  the  six  .s  electron  pairs  to  displace  the  X~  ligand. 

It  is  generally  considered  that  mercurials  do  not  react  with  disulfide 
(S— S)  bonds,  and  there  is  sufficient  evidence  that  this  is  true  for  many 
proteins  and  enzymes  at  physiological  conditions.  However,  Cunningham  et 
al.  (1957)  have  shown  that  p-MB  catalyzes  the  splitting  of  S — S  bonds  in 
cystine,  insulin,  and  ribonuclease  at  pH  7  if  incubation  is  carried  out  at  80°. 
At  this  temperature,  this  may  well  be  a  matter  of  equilibria  between  S — S 
and  SH  groups,  i.e.,  between  native  and  denatured  forms  of  the  proteins, 
with  p-MB  shifting  the  equilibria  by  reacting  with  the  SH  groups.  It  is 
quite  possible  that  in  certain  enzymes  the  S — S  groups  exist  in  a  state 
where  reaction  with  mercurials  is  significant,  and  such  a  reaction  should 
not  be  completely  ignored. 

Certain  metabolically  important  cofactors,  such  as  coenzyme  A  and  li- 
poate,  are  thiols,  and  it  is  of  some  interest  to  inquire  into  whether  the 
mercurials  react  readily  with  them.  Surprisingly  little  quantitative  work 
has  been  done  and  most  of  the  evidence  is  indirect.  For  example,  Galston 
et  al.  (1955)  found  that  p-MB  increases  the  yield  of  peroxidase,  catalase, 
and  tyrosinase  in  plant  breis  when  added  to  the  preparation  medium; 
since  coenzyme  A  inactivates  these  enzymes,  it  was  assumed  that  p-MB 
protects  the  enzymes  by  forming  a  mercaptide  with  the  coenzyme  A.  In 
coenzyme  A-deficient  rats,  the  toxicity  of  mercurials  is  increased,  and  the 
mercurials  inhibit  the  coenzyme  A-dependent  acetylation  of  sulfanilamide 
(Leuschner  et  al.,  1957).  Mersalyl  and  HgClg  reduce  the  coenzyme  A  level 
of  yeast  25%  at  22.5  mM  and  0.3  mM,  respectively  (Estler  et  al,  1960). 
Sanner  and  Pihl  (1962)  followed  the  reaction  between  2?-MB  and  coenzyme 
A  by  changes  in  the  absorption  at  255  m/^  and  showed  that  the  thiol  could 
be  titrated  by  the  mercurial.  Turning  to  DL-a-lipoate,  one  finds  that  its 


REACTIONS   WITH   PROTEINS  751 

administration  prevents  mercurial  poisoning  in  mice,  and  that  it  reduces 
the  inhibition  of  pyruvate  oxidase  by  HgClg  at  high  concentrations  (Gru- 
nert,  1960;  Grunert  and  Rohdenburg,  1960).  It  is  interesting  that  with 
lower  concentrations  of  HgCla  (1.3  mM),  lipoate  increases  the  inhibition  of 
pyruvate  oxidation  in  intact  cells  of  Streptococcus  faecalis,  the  mercaptide 
possibly  entering  the  cells  more  readily  than  the  complexes  of  Hg++.  It 
would  thus  appear  that  mercurials  react  with  coenzyme  A  and  lipoate, 
but  how  rapidly  and  how  tightly  are  not  known. 

Finally,  we  consider  the  problem  of  the  splitting  of  thioesters  by  the 
mercurials.  Sachs  (1921)  showed  that  acetylthioethyl  esters  are  rapidly 
split  by  HgClg  to  acetate  and  mercurothioethanol,  and  in  general  all  acyl 
mercaptans  seemed  to  behave  in  this  way.  Thus  Lynen  et  al.  (1951)  in 
their  early  studies  on  active  acetate  and  coenzyme  A  investigated  acetyl- 
CoA  and  found  it  to  be  split  by  approximately  100  niM  Hg  acetate,  a  result 
of  questionable  significance  in  physiological  work.  Stern  (1956)  studied  ace- 
toacetyl-CoA,  a  thioester  of  importance  in  lipid  metabolism  and  possessing 
a  strong  absorption  maximum  around  303  m//,  and  found  that  HgClg  at 
concentrations  higher  than  0.001  mM  produces  a  rapid  decrease  in  this 
absorption,  0.1-0.2  mM  completely  abolishing  it,  this  corresponding  to 
about  a  1  :  1  molar  ratio  between  Hg  and  thioester.  The  following  reaction 
sequence  was  suggested: 

Hg++  +  AcAc— S— CoA-  ->  Hg— AcAc— S— CoA+ 
Hg— AcAc— S— CoA+  +  H2O  ->  Hg++  +  AcAc"  +  HS— CoA 
Hg++  +  HS— CoA  ->  Hg— S— CoA+  +  H+ 

The  initial  reaction  is  apparently  with  the  enolate  group  of  the  acetoacetyl 
radical.  Stern  believes  that  this  reaction,  occurring  at  such  low  concentra- 
tions of  Hg++  and  so  rapidly,  may  well  be  of  great  importance  in  the  effects 
of  the  mercurials  on  metabolism.  However,  Gibson  et  al.  (1958)  reported 
that  p-MB  does  not  react  with  or  split  succinyl-CoA  at  a  significant  rate, 
and  Sanner  and  Pihl  (1962)  found  a  variety  of  thioesters  to  be  resistant  to 
p-MB  (e.g.,  acetyl-CoA,  succinyl-CoA,  and  benzoyl-CoA).  From  this  limited 
work  one  might  conclude  that  HgClj  can  split  some  thioesters  but  that  ^^-MB 
cannot.  But  Vagelos  and  Earl  (1959)  found  that,  in  contrast  to  most  thio- 
esters, malonyl  semialdehyde  pantetheine  reacts  readily  with  p-MB,  and 
proposed  that  /5-carbonyl  thioesters  may  be  susceptible.  The  possible  role 
of  these  reactions  in  metabolic  inhibitions  is  at  present  unknown. 

REACTIONS   WITH    PROTEINS 

It  was  believed  in  years  past  that  mercury,  in  common  with  other  heavy 
metals,  is  adsorbed  onto  proteins,  denaturing  and  precipitating  them,  but 
recent  work  has  shown  that  under  appropriate  conditions  stoichiometric 


752  7.  MERCURIALS 

combinations  of  mercury  with  proteins  occur,  and  that  denaturation  and 
precipitation  are  by  no  means  a  general  phenomenon.  These  definite  com- 
plexes in  most  cases  are  formed  through  the  SH  groups  of  the  proteins,  and 
methods  whereby  these  groups  may  be  titrated  quantitatively  with  the 
mercurials  have  been  devised.  Some  may  feel  that  a  discussion  of  the  com- 
plexes of  mercury  with  the  proteins  is  out  of  place  in  a  book  on  metabolic 
inhibition,  but  actually  much  can  be  learned  from  the  thorough  and  illu- 
minating investigations  on  mercaptalbumin,  hemoglobin,  and  other  pro- 
teins reported  in  the  past  few  years.  One  must  also  realize  that  in  any 
system  containing  nonenzymic  proteins,  particularly  cellular  preparations, 
reaction  of  mercury  with  these  proteins  not  only  may  have  definite  effects 
on  the  enzyme  inhibition,  but  may  be  responsible,  at  least  in  part,  for  me- 
tabolic or  functional  changes.* 

Protein  Groups  Reacting  with  Mercurials 

The  mercurials  react  rapidly  with  certain  free  and  exposed  protein  SH 
groups,  more  slowly  with  others,  and  not  at  all  with  some  which  are  pre- 
sumably buried  within  the  protein  structure  or  otherwise  sterically  unavail- 
able (page  643).  Many  SH  groups  react  only  after  denaturation  of  the  pro- 
tein, a  process  which  apparently  exposes  them  for  attack  by  the  mercurials. 
Since  mercurials  often  initiate  denaturation,  they  may  produce  a  progressive 
unloosening  of  the  protein  structure  and  themselves  make  available  SH 
groups  originally  unreactive,  the  process  continuing  until  it  is  irreversible, 
the  reformation  of  the  normal  configuration  not  being  possible  when  the 
mercurial  is  removed.  Much  work  with  proteins  and  enzymes,  to  be  dis- 
cussed later,  provides  evidence  that  SH  groups  are  the  primary  site  of 
mercurial  binding,  and  often  the  only  site  under  certain  conditions.  The 
problem  to  be  considered  here  is  whether  protein  groups  other  than  SH 
can  under  any  circumstances  contribute  to  the  binding. 

Examination  of  the  constants  for  the  complexing  of  mercurials  with  SH 
groups  and  with  groups  normally  present  in  proteins  leads  to  the  immediate 

*  There  is  some  inherent  and  unavoidable  difficulty  and  ambiguity  in  the  terminology 
of  the  mercurials.  If  one  uses  the  inorganic  HgClj,  how  should  the  inhibitor  be  desig- 
nated —  as  HgCU,  Hg,  Hg++,  Hg",  or  otherwise?  We  have  seen  that  in  most,  if  not 
all,  media  the  mercuric  ion  will  exist  in  a  variety  of  complexes,  probably  of  different 
reactivities,  so  that  it  is  impossible  to  designate  the  situation  accurately.  Similar 
problems  arise  with  other  heavy  metal  ions,  e.g.,  copper  and  zinc.  We  shall,  therefore, 
arbitrarily  designate  inorganic  divalent  mercury  as  Hg++,  without  implying  that 
this  is  either  a  predominant  form  or  an  active  form.  It  is  fundamentally  and  finally 
the  Hg++  ion  which  complexes  and  reacts  with  the  various  substances  present,  so  that 
when  this  is  so  written  it  must  be  understood  that  all  of  the  complexes  are  implied. 
The  designation  of  the  organic  mercurials  as  79-MB,  PM,  etc.,  similarly  is  noncommittal 
with  respect  to  the  forms  present  or  active. 


REACTIONS   WITH   PROTEINS  753 

conclusion  that  no  single  group  can  compete  very  effectively  with  the  SH 
groups  for  the  mercurials.  Indeed,  these  other  groups  probably  have  their 
orbitals  occupied  by  competing  with  the  various  ligands  present  in  the  me- 
dia. However,  there  are  at  least  three  factors  which  must  be  taken  into 
consideration.  (1)  Certain  fortuitous  arrangements  of  two  or  more  non-SH 
groups,  perhaps  allowing  chelation  of  the  mercurial,  can  increase  the  affin- 
ity markedly,  as  seen  in  Table  7-4.  It  is  quite  possible  that  occasionally 
such  situations  occur  on  protein  surfaces,  although  generally  the  opportu- 
nities for  successful  chelation  must  be  rare;  e.g.,  the  binding  to  glycylglycine 
is  less  than  to  glycine,  and  increase  in  the  length  of  the  polypeptide  chain 
wiU  probably  reduce  affinity  except  in  very  special  cases.  (2)  The  most 
reactive  SH  group  or  groups  on  a  particular  protein  may  not  happen  to 
have  a  strong  affinity  for  a  mercurial,  due  to  steric  factors  or  an  unfavorable 
electric  field,  so  that  non-SH  groups  can  compete  more  effectively.  Although 
there  is  little  quantitative  evidence,  one  gets  the  impression  that  usually  the 
SH  groups  of  proteins  do  not  bind  most  mercurials  as  tightly  as  do  the  SH 
groups  of  simple  thiols,  such  as  cysteine  or  glutathione,  especially  since 
one  can  often  remove  a  mercurial  from  a  protein  quite  readily  by  adding 
one  of  these  thiols.  (3)  As  pointed  out  by  W.  L.  Hughes  (1950),  mercurials 
will  complex  with  non-SH  groups  when  the  SH  groups  become  saturated, 
or  actually  before  in  many  cases.  Since  excess  mercurial  is  often  present, 
especially  in  enzyme  studies,  such  secondary  complexes  must  be  considered, 
even  though  the  SH  groups  are  reacted  first. 

On  the  experimental  side,  it  has  been  observed  that  some  proteins  bind 
more  Hg++  than  corresponds  to  the  SH  content  and  that  some  of  this  is 
relatively  weakly  bound.  Haarmann  (1943  a)  found  that  with  increase  in 
the  pH  progressively  more  Hg++  is  bound  to  various  proteins,  although 
only  a  fraction  is  really  tightly  attached  to  the  protein,  and  postulated  that 
CONH  groups  might  bind  Hg"+.  More  recently,  Perkins  (1958,  1961)  re- 
ported the  binding  of  104-130  g-atoms  of  Hg/10^  g  seralbumin  and,  follow- 
ing treatment  with  bromoacetate  (blocking  SH  and  amino  groups),  the 
binding  increased  to  190  g-atoms/10^  g  protein  at  pH  5.5.  The  SH  groups 
could  have  accounted  for  only  1  g-atom/10^  g  protein  and  Perkins  felt  that 
numerically  the  only  possible  binding  sites  are  the  C00~  groups.  This  is 
a  situation  in  which  there  is  excess  Hg++  present  for  reaction  with  non-SH 
groups  and,  inasmuch  as  the  dissociation  constants  are  not  known,  it  is 
impossible  to  compare  the  affinities  for  the  various  groups.  Nevertheless, 
these  results  conclusively  establish  the  non-SH  binding  of  mercurials  under 
certain  conditions,  and  it  would  be  well  to  bear  this  in  mind  in  enzyme 
inhibition  work. 

Examples  of  Reactions  with  Specific  Proteins 

(A)  Ovalbumin.  Although  not  much  work  has  been  done  with  this  pro- 
tein, the  results  illustrate  some  of  the  problems  one  encounters.  Anson 


754 


7.  MERCURIALS 


(1941)  found  it  difficult  to  determine  if  p-MB  reacts  with  ovalbumin  or 
not,  since  ferricyanide  and  nitroprusside  cannot  be  used  to  detect  these 
SH  groups  (they  do  not  react  with  ovalbumin)  and  iodine  oxidizes  the 
group  whether  free  or  combined  with  p-MB.  Using  an  indirect  method  in- 
volving the  determination  of  the  binding  of  p-MB  to  cysteine  in  the  pres- 
ence of  ovalbumin,  Anson  claimed  that  either  p-MB  does  not  react  at  all 
with  the  ovalbumin  SH  groups,  or,  if  so,  the  binding  is  much  less  strong 
than  with  cysteine.  If  the  ovalbumin  is  denatured,  demonstrable  binding 
of  the  mercurial  occurs.  W.  L.  Hughes  (1950)  reported  that  MM  reacts 
with  ovalbumin  very  slowly,  the  reaction  requiring  a  day  to  come  to  equi- 
librium. However,  MacDonnell  et  al.  (1951)  obtained  a  crystalline  derivative 
of  ovalbumin  treated  with  59-MB  by  adjusting  the  pH  to  4.7  and  adding 
ammonium  sulfate  to  opalescence.  Three  of  the  four  SH  groups  of  oval- 
bumin react  and  the  fourth  reacts  only  after  denaturation.  The  complex 
is  very  stable  and  is  not  dissociated  by  cysteine  or  prolonged  dialysis,  al- 
though 6-day  dialysis  against  cysteine  at  pH  7.9  dissociates  about  75%  of 


EQUIVALENTS 


AMOUNT    OF     MERCURIAL     ADDED 


Fig.  7-3.  Theoretical  curves  for  the  reac- 
tion of  a  mercurial  with  a  protein,  such  as 
ovalbumin  with  three  reactive  SH  groups. 
See  text  for  explanation. 


the  mercurial.  In  such  cases  as  this,  if  one  plots  the  amount  of  mercurial 
combined,  or  the  disappearance  of  SH  groups  from  the  protein,  against  the 
amount  of  mercurial  added,  one  may  obtain  different  curves  (Fig.  7-3). 
If  the  reactions  with  the  free  SH  groups  are  equivalent,  a  linear  relation- 
ship to  saturation  will  be  obtained  (curves  A  and  B),  but  if  interaction  be- 
tween the  SH  groups  occurs  (i.e.,  if  the  binding  of  a  mercurial  reduces  the 
binding  of  the  next)  or  the  SH  groups  combine  with  the  mercurial  with 
different  affinities,  a  curve  concave  downward  will  be  obtained  (curve  C). 
If  no  further  reaction  with  protein  groups  occurs  after  saturation  of  the 
SH  groups,  the  curve  will  be  horizontal  (curve  A),  but  if  other  groups  with 


REACTIONS   WITH   PROTEINS  755 

less  affinity  for  the  mercurial  continue  to  react,  a  sloping  or  curved  line  will 
be  obtained  (curve  B).  It  is  important  to  construct  such  curves  whenever 
possible  in  order  to  understand  the  binding  characteristics. 

(B)  Hemoglobin.  Crystalline  human  oxyhemoglobin  reacts  with  2  moles 
of  MM  per  mole  of  protein  but  the  rate  is  rather  slow  at  pH  7.5  (W.  L. 
Hughes,  1950).  The  hemoglobins  of  other  species  may  contain  either  two 
or  four  SH  groups  that  react  readily.  Green  et  al.  (1954)  crystallized  a  de- 
rivative of  horse  hemoglobin  in  which  two  SH  groups  had  been  reacted 
with  p-MB  and  showed  that,  although  the  crystals  are  isomorphous  with 
normal  hemoglobin,  the  X-ray  diffraction  pattern  is  somewhat  different. 
Ingram  (1955)  established  that  Hg+^,  MM,  and  p-MB  all  combine  readily 
with  four  SH  groups  of  horse  hemoglobin  in  the  native  state,  and  with  six 
in  the  denatured  form;  ox  and  human  hemoglobins  are  similar  but  the 
latter  presents  eight  SH  groups  when  denatured.  In  native  hemoglobin, 
Hg++  probably  reacts  with  two  SH  groups  simultaneously;  at  least  2  equiv- 
alents of  Hg+^  reduce  the  free  SH  groups  to  zero.  However,  p-MB  like- 
wise blocks  two  SH  groups  and  since  reaction  of  a  molecule  of  p-MB  with 
two  SH  groups  is  impossible,  it  is  likely  that  a  pair  of  SH  groups  is  so  close 


Fig.  7-4.  The  reactions  of  the  SH  group 
pairs  on  hemoglobin  with  Hg++  and  p-MB, 
according  to  the  concept  of  Ingram  (19.55). 

that  each  group  cannot  react  with  the  large  mercurial.  It  is  difficult  to  say 
in  the  case  of  Hg++  if  the  effect  is  steric  or  due  to  the  formation  of  a  bridge 
between  the  two  SH  groups  in  a  pair,  but  the  latter  mechanism  is  favored. 
The  situation  as  represented  by  Ingram  is  shown  in  Fig.  7-4.  It  is  interesting 
that  Hg++  and  p-MB  compete  for  the  SH  groups,  and  that  the  former  is 
bound  more  tightly,  probably  due  to  the  reaction  with  two  of  the  SH  groups. 
Since  the  earliest  studies  of  hemoglobin  SH  groups,  there  has  been  diffi- 
culty in  establishing  the  exact  number  of  reactive  and  unreactive  groups, 
due  to  the  fact  that  the  nature  of  the  reactions  and  the  stoichiometry  were 
uncertain.  The  details  are  not  pertinent  to  our  purpose  and  have  been  well 
reviewed  by  Huisman  (1959).  Further  complications  have  arisen  in  the  re- 
cognition that  different  types  of  hemoglobin  react  differently  with  the  mer- 
curials and  that  the  pH  is  an  important  factor.  Murayama  (1958)  claimed 


756  7.  MERCURIALS 

that  adult  hemoglobin  contains  two  SH  pairs,  sickle  cell  hemoglobin  three 
SH  pairs,  and  hemoglobin  C  no  pairs,  while  Huisman  obtained  quite  differ- 
ent results  indicating  more  reactive  SH  groups  than  were  found  by  most 
previous  workers.  At  pH  7,  adult  hemoglobin  reacts  with  3  p-MB  molecules 
and  fetal  hemoglobin  with  2,  whereas  at  pH  4.6  the  values  are  6  and  4, 
respectively,  the  lowering  of  the  pH  presumably  altering  the  configuration 
of  the  protein.  Studies  of  mercaptalbumin  show  that  Hg++  can  induce  the 
formation  of  dimers  of  the  protein,  but  Allison  and  Cecil  (1958)  believe 
that  only  monomers  occur  in  the  case  of  hemoglobin,  and  found  that  Hg++ 
and  PM  give  the  same  titer,  while  the  results  of  Cecil  and  Snow  (1962)  are 
more  in  accord  with  those  of  Ingram,  2.2  reactive  SH  groups  of  a  total  of 
6  in  adult  hemoglobin  being  detected,  the  3.8  sluggish  SH  groups  reacting 
differently  with  Hg++,  PM,  and  p-MPS.  These  studies  of  hemoglobin  not 
only  show  the  influence  of  many  factors  on  the  interaction  of  proteins  with 
mercurials,  but  point  out  the  difficulties  of  SH  titration  of  even  relatively 
simple  proteins. 

Reaction  of  hemoglobin  with  mercurials  brings  about  striking  changes  in 
the  characteristics  of  oxygenation:  The  affinity  of  hemoglobin  for  Og  may 
be  unaffected  or  increased,  but  the  interactions  between  the  heme  groups 
are  reduced  or  abolished  by  all  the  mercurials  (A.  F.  Riggs,  1952;  Wolbach 
and  Riggs,  1955;  Riggs  and  Wolbach,  1956;  A.  Riggs,  1959).  Mersalyl  and 
Hg++  at  pH  6.8  increase  the  affinity  for  0^  quite  markedly,  Pjq  decreasing 
to  about  one  third  of  normal,  but  p-MB  and  MM  do  not  alter  the  affinity. 
AU  these  mercurials  lower  the  interaction  constant  n  from  around  2.9  to  1 
(or  near  1),  the  latter  corresponding  to  complete  loss  of  heme-heme  interac- 
tion. Glutathione  completely  reverses  these  effects.  It  is  very  interesting 
that  the  maximal  effects  are  produced  at  ratios  close  to  2  moles  of  mer- 
curial to  1  mole  of  hemoglobin;  however,  as  the  amount  of  mersalyl  is  in- 
creased, the  changes  in  heme  interaction  and  Og  affinity  progressively  dis- 
appear, so  that  at  a  ratio  of  15-16  moles  of  mercurial  per  mole  of  hemoglobin 
there  is  no  longer  an  effect.  This  curious  reversal  is  unexplained.  The  ob- 
servation may  have  some  bearing  on  the  use  of  mercurials  for  enzyme  inhi- 
bition, and  Riggs  and  Wolbach  (1956)  state,  "Our  observations  suggest 
that  the  attempt  to  inhibit  an  enzyme  with  only  a  single  high  concentration 
of  mercurial  may  lead  to  spurious  conclusions."  I  know  of  no  example  in 
which  enzyme  inhibition  is  lost  at  higher  mercurial  concentrations,  but  in 
any  case  it  would  presumably  be  a  rare  phenomenon.  Occasionally  one  finds 
stimulation  of  enzyme  activity  at  low  mercurial  concentrations  and  this 
reverses  to  inhibition  as  the  concentration  is  increased,  but  it  is  not  known 
if  this  has  any  relation  to  the  above  reversal.  Mercuration  of  oxyhemoglobin 
increases  the  rate  constants  for  the  dissociation  of  Og  from  three  of  the 
hemes,  but  decreases  the  rate  constant  for  the  dissociation  of  the  last  Og, 
this  being  in  fair  accord  with  the  effects  on  the  Og  dissociation  curve  ob- 


REACTIONS   WITH   PROTEINS  757 

served  by  Riggs  (Gibson  and  Houghton,  1955).  Oxygenation  of  hemoglobin 
facilitates  reaction  of  the  SH  groups  with  A^-ethylmaleimide  and  iodoaceta- 
mide  but  not  with  p-MB  (Benesch  and  Benesch,  1962).  The  mechanisms 
by  which  these  effects  are  produced  are  not  clear,  but  hypotheses  have  been 
offered  based  on  the  spatial  arrangement  of  the  SH  groups  and  the  hemes. 
Riggs  (1959)  considers  the  hemoglobin  molecule  to  consist  of  two  halves, 
each  with  a  pair  of  reactive  SH  groups  and  a  pair  of  hemes,  the  SH  groups 
perhaps  lying  between  the  hemes.  Any  mercurial  which  can  form  a  bridge 
between  the  2  SH  groups  of  a  pair  —  such  as  Hg++,  or  mersalyl  if  the 
C — Hg  bond  is  ruptured  and  inorganic  Hg++  is  released  —  increases  the 
affinity  of  the  hemes  for  Og,  whereas  mercurials  reacting  only  with  a  single 
SH  group  —  such  as  p-MB  and  MM  —  do  not  have  this  effect.  Any  interac- 
tion with  the  SH  groups  reduces  the  interaction  between  the  hemes,  prob- 
ably by  bringing  about  reversible  structural  changes  in  the  protein  config- 
uration. On  the  other  hand,  Klotz  and  Klotz  (1959)  favor  a  mechanism 
involving  disturbances  in  the  water  structure  around  and  between  the 
hemes.  Whatever  the  explanation,  the  bearing  on  the  effects  of  mercurials 
on  enzyme  active  centers  by  reaction  with  adjacent  SH  groups  is  obvious. 

(C)  Mercaptalbumin.  Mercaptalbumin  is  one  fraction  of  the  serum  al- 
bumins containing  a  single  reactive  SH  group  whereas  the  other  albumins 
contain  none.  It  was  isolated  as  the  crystalline  mercury  salt  by  Hughes 
(1947)  and  its  reactions  have  been  studied  in  detail,  so  that  it  has  become 
the  classic  example  of  protein-mercurial  interaction  (W.  L.  Hughes,  1950). 
The  three  major  reactions  maj^  be  represented  as  follows: 

Reaction  1:     alb— SH  +  RgX^  ±5  alb— S— HgX  +  H+  4-  X" 

Reaction  2:     alb— S— HgX  +  alb— SH  ±^  alb— S— Hg— S— alb  -f-  H+  -^  X- 

Reaction  3:     alb— S— Hg— S— alb  +  RgX^  ^  2  alb— S— HgX 

where  alb  indicates  mercaptalbumin  and  X  some  ligand  (e.g.  Cl~).  The 
first  reaction  is  mercaptide  formation,  the  second  dimerization,  and  the 
third  dissociation  of  the  dimer  by  excess  HgXg.  Formation  of  the  dimer 
increases  the  turbidity  and,  if  some  ethanol  is  added,  crystals  form.  These 
crystals  are  colorless  diamond-shaped  orthorhombic  plates,  containing  chan- 
nels or  enclosures  of  fair  size,  with  liquid  within  them,  and  permeable  to 
various  salts,  sugars,  and  dyes  (Low  and  Weichel,  1951).  Hughes  and  Dint- 
zis  (1964)  have  described  procedures  for  crystallizing  the  dimers  from  etha- 
nol-water  mixtures  at  low  temperatures.  Viscosity  and  sedimentation  stud- 
ies (Low  1952)  led  to  the  representation  of  the  dimer  as  in  A  of  Fig.  7-5, 
while  the  results  of  X-ray  diffraction  study  are  compatible  also  with  struc- 
ture B.  The  lengths  of  the  dimer  would  be  around  140-150  A,  monomer 
mercaptalbumin  being  of  molecular  weight  66,000.  The  structure  is  inde- 
pendent of  the  smaller  ions  making  up  the  crystal;  e.g.,  the  dimer  will 


758 


7.  MERCURIALS 


form  crystals  with  Hgl^g,  the  interactions  being  purely  electrostatic  and 
not  involving  SH  groups  (Lewin,  1951). 

Reaction  1  is  quite  rapid,  but  reaction  2  is  slow  because  it  involves  two 
large  molecules  of  similar  charge.  The  dimerization  requires  about  25  min 
for  half-reaction  and  2  hr  for  equilibrium  when  Hg++  is  mixed  with  mer- 


~D 


Fig.  7-5.  Possible  forms  of  the  Hg-mercaptalbumin  dimer. 
The  small  solid  circle  represents  Hg++.  (From  Low,  1952.) 

captalbumin  in  a  0.5  :  1  ratio  (Edelhoch  et  al.,  1953).  Dimerization  is  an 
endothermic  reaction,  rise  in  the  temperature  favoring  formation  of  the 
dimer,  AH^  being  about  7  kcal/mole.  The  temperature  effect  indicates 
an  activation  energy  for  dimerization  of  17-21  kcal/mole.  The  constant 
for  the  equilibrium  (alb— S— HgCl)  (alb— SH)/(alb— S— Hg— S— alb)  was 
found  to  be  3.2  X  10^^^  at  pH  4.5  and  25°.  Reaction  1  is  reversible  by  sub- 
stances forming  stable  compounds  with  Hg++,  and  reaction  2  is  reversible 
by  ligands  forming  HgX„  complexes.  Dimerization  is  also  reversed  by  reac- 


AMOUNT 
OF  DIMER 
(TURBIDITY) 


0  0.5 

MOLE     Hg/MOLE     PROTEIN 


Fig.  7-6.  The  formation  and  dissociation 
of  the  Hg-mercaptalbumin  dimer  as  the 
molar  ratio  of  Hg++  to  protein  is  in- 
creased.   (From    Edelhoch   et  al.,    1953.) 


tion  3  (Fig.  7-6),  Hg++  competing  with  mercaptalbumin  for  the  alb — S — Hg 
monomer.  The  dissociation  of  the  dimer  by  cysteine  is  very  rapid  (appar- 
ently within  a  few  seconds)  and  yet  the  dimer  is  not  split  nor  the  Hg++ 
dissociated  readily  by  dialysis,  so  that  this  would  seem  to  be  one  of  those 
interesting  situations  in  which  a  complexer  appears  to  take  the  metal  from 
the  protein  rather  than  merely  combining  with  free  metal  ions  (Straessle, 


REACTIONS   WITH   PROTEINS  759 

1954).  The  disulfide  dimer,  alb — S — S — alb,  however,  is  dissociated  by  cys- 
teine very  slowly.  The  dimerization  is  accomparied  by  an  increase  in  levoro- 
tation,  and  this  implies  that  the  mercaptalbumin  molecule  undergoes  some 
unfolding  in  the  region  of  the  reactive  SH  groups  as  a  necessary  prelude 
to  dimerization;  this  may  be  thought  of  as  a  partial  denaturation,  adding 
one  more  item  of  evidence  for  configurational  changes  induced  by  mercuri- 
als (Kay  and  Marsh,  1959). 

Organic  mercurials  such  as  p-MB,  PM,  and  MM  react  in  a  1  :  1  ratio  with 
mercaptalbumin  and,  of  course,  no  dimer  is  formed.  The  equilibrium  con- 
stant for  the  reaction: 

alb— SH  +  CH3— Hgl  ±5  alb— S— Hg— CH3  +  H+  +  Cl- 

has  been  found  by  \V.  L.  Hughes  (1950)  to  be  3.5  x  lO^^  (p^  =  4  45)^ 
from  which  the  value  for  the  dissociation  constant  of  the  mercaptide  in 
Table  7-6  was  calculated.  On  the  other  hand,  bifunctional  organic  mercuri- 
als, such  as: 

CH-O 

XHg-CH2— HC  CH— CH2— HgX 

O— CH, 


can  link  two  mercaptalbumin  molecules  together  (Straessle,  1951;  Edsall 
et  al.,  1954).  The  pA"  for  the  equilibrium  (dimer )/(monomer+)  (alb")  is  18.2 
at  pH  4.75  and  25°,  the  corresponding  p^  for  the  Hg++  dimer  equilibrium 
being  13.5.*  This  difference  of  some  4.7  p/C  units  between  the  two  dimers 
is  undoubtedly  due  to  the  fact  that  the  mercaptalbumin  molecules  must 
approach  about  10  A  closer  in  the  Hg++  dimer  than  in  the  bifunctional 
mercurial  dimer,  and  the  steric  and  electrostatic  factors  could  easily  ac- 
count for  the  some  6.7  kcal/mole  difference. 

The  importance  of  these  results  with  mercaptalbumin  for  inhibition  stud- 
ies with  the  mercurials  is  clear.  First,  the  possibility  of  the  dimerization 
of  certain  enzymes  by  Hg++  leads  to  the  concept  that  inhibition  may  oc- 
casionally result  through  steric  sequestration  of  the  active  sites  and  not 
necessarily  through  the  reaction  of  SH  groups  at  the  active  sites.  It  is  also 
possible  that  occlusion  of  active  sites  could  occur  by  linking  a  nonenzyme 
protein  to  the  protein  through  a  Hg  bridge.  It  is  known  that  Hg++  often 
produces  an  increase  in  turbidity  of  enzyme  solutions  and  even  precipita- 
tion, although  this  can  be  due  to  other  factors  as  well.  Second,  this  reason- 
ably well  understood  and  quantitatively  investigated  system  provides  a 
model  on  which  the  effects  of  ligands,  pH,  temperature,  and  other  factors 

*  Gurd  and  Wilcox  (1956)  give  values  of  17.2  and  12.6  for  the  p/C's,  respectively, 
due  to  different  assumptions  regarding  the  ligand  constants. 


760  7.  MERCURIALS 

may  be  better,  appreciated.  Third,  it  presents  one  clear  instance  in  which 
mercurials  react  specifically  with  protein  SH  groups,  since  there  is  no  evi- 
dence that  other  groups  are  even  involved  in  the  mercuration  of  mercaptal- 
bumin.  Last,  it  is  significant  in  the  use  of  mercurials  in  whole  animals  that 
the  serum  contains  sufiicient  mercaptalbumin  to  bind  most,  if  not  all,  the 
mercurial  present,  a  factor  that  must  be  considered  in  the  penetration,  dis- 
tribution, and  actions  of  the  mercurials  in  animals. 

Effect  of  pH 

The  competition  between  H+  and  the  mercurial  for  the  S"  group  and  the 
effect  this  has  on  the  over-all  equilibrium  have  been  discussed.  On  the  basis 
of  only  the  ionization  of  the  SH  group,  one  would  predict  that  mercurials 
would  react  more  rapidly  and  more  completely  at  higher  pH's  (particularly 
above  p^^'s  of  the  SH  groups).  However,  there  are  many  other  factors 
which  may  be  important.  Actually,  it  has  generally  been  observed  that  the 
rate  of  reaction  of  p-MB  with  proteins  is  decreased  with  a  rise  in  the  pH. 
Both  the  extent  and  rate  of  reaction  of  p-MB  with  ovalbumin  are  affected 
by  pH:  At  pH  4.6,  4  moles  of  p-MB  react  rapidly  with  1  mole  of  protein, 
whereas  at  pH  7,  only  3.2  moles  of  p-MB  react  in  24  hr  (Boyer,  1954). 
Reduction  of  the  rate  with  increasing  pH  has  also  been  reported  for  /5- 
lactoglobulin  and  3-PGrDH  (Boyer  and  Segal,  1954),  it  being  much  faster 
at  4.6  than  at  7.  The  work  of  Huisman  (1959)  with  hemoglobin  illustrates 
an  important  point;  the  rate  and  extent  of  reaction  may  be  influenced  dif- 
ferently by  pH,  inasmuch  as  the  rate  of  mercaptide  formation  is  faster  be- 
tween 7  and  11.2  than  at  4.6,  but  more  SH  groups  are  reactive  at  4.6.  One 
also  recalls  that  p-MB  is  dissociated  more  rapidly  from  ovalbumin  at  pH 
7.9  than  in  the  acid  pH  range  (MacDonnell  et  al.,  1951).  Another  factor 
often  overlooked  is  the  effect  of  pH  on  the  secondary  denaturation  of  the 
protein  following  mercuration.  The  rate  of  thermal  denaturation  of  seral- 
bumin is  increased  13.2-fold  at  pH  3.6,  while  at  pH  7  there  is  no  denatura- 
tion by  Hg++  at  1.85  mM,  while  denaturation  of /?-lactoglobulin  is  increased 
89-fold  at  pH  3.6  and  not  at  all  at  pH  7,  this  indicating  that  an  acid  pH 
favors  secondary  configurational  changes  resulting  from  binding  of  the  Hg++ 
(Stauff  and  Uhlein,  1958).  The  reaction  of  mercurials  with  non-SH  groups 
is  also  pH-dependent,  since  the  complexing  of  the  azomercurial  studied  by 
Horowitz  and  Klotz  (1956)  with  glycine  is  maximal  between  pH  6.5  and 
9.5;  at  low  pH's  the  +H3N — CH2 — C00~  form  of  glycine  dominates  and 
is  less  reactive,  while  at  high  pH's  there  is  competition  by  0H~. 

R-Hg-00C-CH3  +  H2N-CH2-COO-  ±?  R-Hg-OOC— CH2-NH2  +  CH3-COO- 

The  following  factors,  in  addition  to  the  ionization  of  SH  groups,  must 
thus  be  considered.  ( 1 )  The  pH  may  vary  the  number  of  reactive  SH  groups 


REACTIONS  WITH  PROTEINS  761 

or  their  individual  reactivity,  by  effects  on  the  structure  of  the  protein,  or 
on  the  association  of  subunits.  (2)  The  pH  will  determine  the  over-all  charge 
on  the  protein;  e.g.,  with  increased  pH  the  protein  will  become  more  neg- 
atively charged  and  possibly  repel  negatively  charged  mercurials,  such  as 
p-MB,  p-PMS,  or  the  higher  Ch  complexes  with  Hg++.  (3)  Rise  in  pH  will 
increase  the  OH"  concentration  and  this  ion  will  compete  with  the  SH 
groups  for  the  mercurial.  (4)  The  pH  may  determine  the  degree  of  hydrogen 
bonding  of  SH  groups  and  thus  their  reactivity  with  mercurials.  It  is  likely 
that  the  dependence  on  the  pH  will  depend  on  the  mercurial  used,  but  in- 
sufficient data  are  available  for  comparisons. 

A  final  effect  of  pH  involves  dimerization  where  it  occurs.  The  rate  and 
degree  of  dimerization  in  the  presence  of  Hg++  will  depend  on  the  total 
protein  charge,  being  maximal  at  the  isoelectric  point,  all  else  being  equal. 
Straessle  (1951)  reported  that  the  dimerization  of  mercaptalbumin  with  a 
bifunctional  mercurial  is  slower  at  pH  6  than  at  4.75,  and  Edelhoch  et  al. 
(1953)  found  the  rate  of  dimerization  with  Hg++  to  be  increased  60  times 
when  the  pH  is  decreased  from  6  to  4.75,  and  doubled  with  further  de- 
crease to  4.25.  It  was  calculated  that  a  charge  of  9  charge  units  would  ac- 
count for  this,  and  titration  data  indicated  a  change  of  10  units  over  this 
pH  range,  so  the  electrostatic  mechanism  seems  to  be  correct. 

Effects  of  Mercurials  on   Protein  Structure  and  Properties 

The  importance  of  secondary  changes  in  protein  structure  upon  reaction 
with  a  mercurial  cannot  be  overemphasized  in  studies  of  enzyme  inhibition 
and  its  reversibility,  but  unfortunately  little  exact  information  is  available. 
Configurational  changes  have  been  postulated  to  explain  certain  results, 
such  as  have  already  been  mentioned  in  regard  to  mercaptalbumin  (page 
757)  and  hemoglobin  (page  755),  and  additional  examples  wiU  be  presented 
in  connection  with  enzyme  inhibition,  but  in  most  instances  the  evidence  is 
indirect  and  tenuous.  Nevertheless  most  investigators  agree,  I  believe,  in 
accepting  that  such  changes  occur  in  certain  cases;  the  problems  are  the 
nature  of  the  changes  and  the  mechanisms  by  which  they  are  induced. 

Higher  concentrations  of  Hg++  and  most  organic  mercurials  decrease  the 
solubility  of  proteins,  and  may  precipitate  or  coagulate  them.  This  gave 
rise  to  the  early  concept  of  the  mercurials  as  denaturing  agents.  However, 
it  would  appear  that  the  primary  effect  is  seldom  denaturation  (in  the  sense 
of  disruption  of  the  polypeptide  chain  structure),  and  that  the  altered  prop- 
erties of  the  protein  are  more  directly  related  to  modification  of  side-chain 
groups  and  the  introduction  of  new  groups.  Prolonged  contact  of  proteins 
with  mercurials  occasionally  leads  to  true  denaturation  as  a  secondary  reac- 
tion, but  complete  reversibility  can  usually  be  achieved  by  removing  the 
mercurial;  this  indicates  that  if  structural  changes  occur  they  are  probably 
localized,  and  that  the  normal  configuration  can  be  restored.  Such  direct 


762  7.  MERCURIALS 

structural  effects  should  be  distinguished  from  preferential  reaction  of  mer- 
curials with  denatured  protein  in  cases  in  which  native  and  denatured  pro- 
tein exist  in  equilibrium  under  nonphysiological  conditions  (e.g.,  at  low  pH's 
or  high  temperatures)  (Habeeb,  1960).  They  should  also  be  distinguished 
from  changes  brought  about  in  proteins  evident  after  precipitation.  The 
thermal  coagulation  of  serum  proteins  is  enhanced  by  p-MB  and  the  coagula 
are  firmer,  more  elastic,  and  more  transparent,  the  water  bound  to  the  clot 
being  around  4  times  greater  (Jensen,  et  al.,  1950),  but  this  does  not  pro- 
vide evidence  that  protein  configuration  before  coagulation  is  altered  by 
the  mercurial.  High  Hg++  concentrations  weaken  keratin  fibers  so  that  they 
break  under  less  tension  (Hoare  and  Speakman,  1963).  This  would  be  ex- 
pected if  interchain  disulfide  bonds  are  disrupted.  Changes  in  gross  protein 
properties  seldom  provide  information  on  the  more  important  and  subtler 
localized  modifications  which  are  believed  to  occur. 

In  view  of  the  significance  of  configurational  changes  in  enzyme  inhibi- 
tion, and  in  the  belief  that  more  examples  will  be  postulated  and  established 
in  the  coming  years,  we  may  summarize  some  of  the  possible  mechanisms 
by  which  such  effects  can  be  brought  about.  (1)  In  those  cases  in  which 
there  are  equilibria  between  SH  and  S — S  groups,  or  where  there  is  a  cyclic 
oxidation  and  reduction,  and  in  which  the  S — S  bonds  contribute  to  the 
stability  of  a  local  configuration,  mercaptide  formation  may  loosen  the 
structure.  (2)  The  SH  groups  themselves  may  contribute  to  the  stability, 
perhaps  by  hydrogen  bonding  or  the  binding  of  cofactors,  so  that  mercura- 
tion  may  again  enhance  dissolution  of  the  native  structure.  (3)  The  intro- 
duction of  a  charged  group,  such  as  occurs  with  p-MB  or  p-MPS,  will  alter 
the  local  electric  field,  and  this  may  favor  instability.  (4)  Hg++  and  or- 
ganic mercurials  which  dissociate  to  form  Hg++  can  bind  to  two  groups 
simultaneously  and  thereby  distort  protein  configuration.  (5)  Reaction  of 
mercurials  with  non-SH  groups,  especially  N-  and  0-containing  groups, 
may  reduce  hydrogen  bonding  between  polypeptide  chains.  It  is  not  ne- 
cessary that  the  affinity  for  these  groups  be  especially  high,  and  there  is 
some  indirect  evidence  that  it  is  often  the  excess  mercurial,  above  that 
required  to  saturate  the  SH  groups,  which  is  responsible  for  denaturation. 
There  is  increasing  reason  for  believing  that  proteins  are  not  rigid  structures 
but  often  exhibit  a  fair  degree  of  flexibility  (page  1-199),  so  that  it  is  reason- 
able that  reversible  modifications  of  the  structure  may  be  fairly  easily  in- 
duced. 

Estimation  and  Titration  of  Protein  SH  Groups  with  Mercurials 

The  older  methods  for  the  determination  of  SH  groups,  using  nitroprus- 
side,  ferricyanide,  iodine,  or  other  reagents,  are  now  considered  to  be  gen- 
erally unreliable  when  applied  to  proteins,  due  mainly  to  lack  of  specifi- 
city, and,  in  addition,  these  methods  are  often  rather  laborious.  Ampero- 


REACTIONS    WITH   PROTEINS  763 

metric  titration  with  Ag+  is  still  commonly  used  and  is  often  useful  when 
combined  with  mercurial  titration;  one  should  consult  Leach  (1960)  for  a 
discussion  of  some  of  the  difficulties  of  this  method.  Amperometric  titra- 
tions at  a  rotating  platinum  electrode  using  Hg++  or  MM  have  been  shown 
to  be  accurate  and  reliable  in  some  cases  (Saroff  and  Mark,  1953;  Kolthoff 
et  al.,  1954;  Leach,  1960),  but  these  technique  have  not  yet  been  extensively 
applied  to  enzymes.  Leach  has  listed  the  requirements  for  an  ideal  SH  re- 
agent for  titrations:  It  should  (1)  be  specific  for  SH  groups,  (2)  be  highly 
reactive,  (3)  have  a  small  molecular  size,  (4)  be  preferably  monofunctional, 
(5)  be  devoid  of  charge  or  other  reactive  groups  so  that  all  protein  SH 
groups  are  equally  favored  despite  their  different  environments,  (6)  be  sol- 
uble, stable,  and  reactive  over  a  range  of  pH,  and  (7)  show  well-defined 
reduction  steps  for  amperometric  use.  MM  fulfills  most  of  these  criteria  and 
perhaps  it  has  been  neglected  in  enzyme  work.  Past  work  on  many  pro- 
teins has  indicated  that  it  is  often  well  to  use  more  than  one  method  in 
order  to  increase  the  reliability. 

The  most  commonly  used  method  at  present  is  the  spectrophotometric 
titration  with  p-MB  developed  by  Boyer  (1954),  since  it  is  convenient  and 
appears  to  be  generally  accurate.  Furthermore,  the  sensitivity  is  as  high 
as  with  the  amperometric  methods,  namely,  around  0.01-0.1  ulM  SH. 
Reaction  of  p-MB  with  SH  groups  leads  to  an  increase  in  the  absorbancy 
at  250  m//  at  pH  7  (Fig.  7-7);  at  pH  4.6  the  maximal  increment  occurs  at 
255  m//,  but  it  is  usually  preferable  to  titrate  enzymes  at  the  more  physi- 
ological pH  of  7  to  determine  the  number  of  reactive  SH  groups,  unless 
the  mercaptide  formation  occurs  too  slowly.  The  increase  in  absorbancy  is 
a  linear  function  of  the  SH  groups  reacted,  for  both  simple  thiols  and  pro- 
teins (Fig.  7-8),  but  the  absorbancy  change  is  somewhat  different  for  dif- 
ferent SH  groups,  a  fact  of  no  importance  in  titrations.  The  p-MB  may  be 
titrated  with  protein  (as  in  Fig.  7-8),  the  end-point  being  the  sharp  break 
between  the  two  linear  segments,  or  the  protein  may  be  titrated  with  p-MB; 
in  both  cases  one  determines  when  all  the  reactive  SH  groups  are  transform- 
ed into  mercaptides.  Although  the  details  of  the  method  may  be  found  in 
the  original  paper  of  Boyer  (1954)  and  the  excellent  review  of  R.  Benesch 
and  R.  E.  Benesch  (1962),  it  may  be  useful  in  interpreting  such  titrations 
applied  to  enzymes  to  note  briefly  certain  precautions  and  difficulties. 

(1 )  The  only  mercurial  which  can  be  used  is  jj-MB,  and  since  it  is  a  rather 
large  molecule  with  a  negative  charge,  steric  or  electrostatic  factors  may 
reduce  its  reaction  with  certain  SH  groups.  PM  and  7)-MPS  exhibit  ab- 
sorbancy shifts  but  at  lower  wavelengths  where  protein  absorbs  much  more 
strongly  than  at  250-255  m//. 

(2)  The  addition  of  excess  p-MB  occasionally  results  in  a  further  small 
increment  in  the  absorbance,  so  that  the  flat  portion  of  the  curve  may  not 
be  exactly  horizontal,  and  this  may  indicate  reaction  with  non-SH  groups 


764 


7.  MERCURIALS 


or  less  readily  reacting  SH  groups.  Such  behavior  seems  to  be  rare  and  does 
not  seriously  interfere  with  the  determination  of  the  end-point. 

(3)  The  time  relations  must  be  considered.  It  is  common  practice  to  in- 
cubate the  protein  and  p-MB  for  10-15  min  to  allow  reaction,  but  a  decision 
as  to  this  depends  on  one's  definition  of  a  reactive  SH  groups.  In  some 
proteins,  additional  SH  groups  are  reacted  when  the  incubation  is  prolonged; 
is  such  cases  one  is  not  certain  if  these  groups  were  initially  exposed,  or 
if  they  arise  during  a  progressive  denaturation  of  the  protein. 


Fig.  7-7.  The  spectral  absorption  curves  for 

p-MB   and   its   mercaptide  with  cysteine,  at 

pH   7   in   60   raM   phosphate   buifer.    (From 

Boyer,   1954.) 


(4)  Both  proteins  and  p-MB  absorb  significantly  at  250-255  m/^  and  the 
appropriate  controls  must  be  run.  For  example,  when  protein  is  titrated 
with  p-MB,  equivalent  increments  of  the  mercurial  are  added  to  the  blank 
cell. 

(5)  Protein  or  enzyme  solutions  should  be  as  pure  as  possible,  since  even 
small  amounts  of  certain  impurities  may  cause  large  errors,  and  the  solu- 
tions should  be  clear  so  that  light  scattering  is  reduced. 

(6)  Special  consideration  should  be  given  to  the  pH  since  it  has  been 
shown  that  both  the  rate  and  extent  of  reaction  are  markedly  affected,  as 
in  Boyer's  experiments  with  ovalbumin.  Although  reaction  may  be  more 


REACTIONS   WITH   PROTEINS 


765 


rapid  at  pH  4.6  than  at  7,  it  is  preferable,  as  mentioned  above,  to  use  as 
physiological  a  pH  as  possible  if  the  normal  state  of  the  protein  or  enzyme 
is  to  be  established. 

(7)  Salt  effects  on  the  reaction  of  p-MB  with  proteins  are  also  often  of 
some  magnitude,  so  that  attention  must  be  given  to  the  ionic  composition 
of  the  medium  and  the  buffers  used.  It  is  preferable  in  most  cases  to  use  as 
low  concentrations  of  salt  and  buffer  as  possible,  if  maximal  reactivity  of 
the  SH  groups  is  desired,  but  occasionally  it  is  useful  to  add  some  salt, 
such  as  KCl,  to  reduce  the  p-MB  reactivity  in  order  to  eliminate  non-SH 
group  effects. 


0.5 


0-  LACTOGLOBULIN 


'0   5 

MilO 


Fig.  7-8.  Titration  of  cysteine  and  proteins  with  p-MB  at 

pH  4.6  in  330  mM  acetate  medium.  The  reaction  times 

are:  cysteine   and  ovalbumin   15   min,  and  lactoglobulin 

20  hr.  (From  Boyer,  1954.) 


(8)  It  is  worth  noting  that  Boyer  found  EDTA  to  interfere,  presumably 
due  to  a  complex  with  p-MB,  so  it  is  advisable  to  omit  this  substance. 

(9)  Masking  of  the  reactive  SH  groups,  e.g.  with  alkylating  agents,  abol- 
ishes the  absorbancy  changes  on  adding  p-MB,  this  providing  evidence 
that  it  is  indeed  the  SH  groups  which  are  responsible  for  the  changes.  Ti- 
tration of  proteins  or  enzymes  treated  with  various  agents  can  thus  provide 
information  on  the  disappearance  of  SH  groups, 

(10)  The  p-MB  must  be  pure,  should  be  analyzed  iodometricaUy  or  spec- 
trophotometrically,  should  be  standardized  against  glutathione  (details  are 


766 


7.  MERCURIALS 


given  by  Benesch  and  Benesch),  and  should  be  used  in  freshly  made  so- 
lutions. 

(11)  The  presence  of  two  or  more  SH  groups  close  together  on  the  protein 
may  prevent  the  reaction  of  each  with  jj-MB,  as  is  the  case  with  hemoglobin. 
This  will  lead  to  low  values  for  the  total  number  of  SH  groups  in  proteins 
or  enzymes. 

A  typical  titration  of  an  enzyme  is  shown  in  Fig.  7-9.  The  titration  of 
3-phosphoglyceraldehyde  dehydrogenase  at  pH  4.6  presents  a  clear  end- 
point  indicating  a  rapid  reaction  of  the  SH  groups.  This  yields  10.3  SH 
groups  per  molecule  of  enzyme  (assumed  molecular  weight  of  118,000). 


0.3 


4  5 

/i MOLES/ML  iio' 

Fig.  7-9.  Titration  of  5  times  recrystal- 
lized  3-phosphoglyceraldehyde  dehydro- 
genase with  p-MB  at  pH  4.6  and  0.03 
/<mole/ml.  (From  Boyer  and  Segal,  1954.) 

The  reaction  of  the  SH  groups  occurs  more  slowly  at  pH  7  and  a  sharp 
end-point  was  not  obtained  by  incubations  up  to  15  min;  however,  longer 
incubations  would  probably  have  given  a  sharp  break  in  the  curve.  Here 
the  end-point  yields  8.3  SH  groups  per  enzyme  molecule,  suggesting  that 
2  SH  groups  become  much  more  reactive  when  the  pH  is  lowered.  A  value 
of  10.7  half-cystines  per  molecule  for  this  enzyme  has  been  reported  (Velick 
and  Ronzoni,  1948),  so  it  is  evident  that  most  of  these  SH  groups  are  free 
and  reactive. 


Colored  Mercurials  and  Histochemical  Determination  of  Protein  SH  Groups 

Various  colored  mercurials,  usually  azobenzene  derivatives,  have  been 
known  for  many  years  but  were  not  applied  to  biological  material  until 
Bennett  (1948  a)  studied  the  reaction  of  p-mercuriphenylazo-/5-naphthol 
with  tissue  thiols.  Direct  visualization  of  thiol  distribution  in  the  tissues  is 
possible,  but  the  dye  has  a  very  low  solubility  in  water  and  at  the  usual 
pH's  so  low  a  molecular  extinction  coefficient  that  its  use  is  limited.  How- 


REACTIONS    WITH    PROTEINS  767 

ever,  Flesch  and  Kun  (1950)  found  that  the  addition  of  strong  acid  intensifies 
the  color  markedly.  It  has  been  claimed  that  this  mercurial  is  as  specific  as 
PM  for  SH  groups,  but  one  wonders  if  this  complex  molecule  does  not 


N=N^    y-Hg 


/)-Mercuriphenylazo-/:i-naphthol 

through  other  groups  occasionally  react  with  various  tissue  components 
(/5-naphthol  derivatives  being  fairly  potent  enzyme  inhibitors),  although 
previous  treatment  of  the  tissue  with  Hg++  or  iodoacetamide  is  said  to 
prevent  staining.  Aqueous  solutions  of  thiols,  proteins,  or  tissue  homogenates 
are  shaken  with  an  amyl  acetate  solution  of  the  mercurial  dye,  and  a  red 
precipitate  slowly  forms  in  the  aqueous  phase  as  the  reaction  proceeds;  the 
amount  of  precipitate  is  proportional  to  the  number  of  SH  groups  and  can 
be  determined  colorimetrically  after  centrifuging  and  redissolving  in  acid 
solution.  Fragments  of  dehydrated  tissues  may  also  be  placed  in  butanol 
or  propanol  solutions  of  the  mercurial  for  several  hours,  and  the  staining 
demonstrated  histologically  (Bennett,  1951).  Bennett  ran  controls  with  phe- 
nylazo-/5-naphthol  to  determine  if  this  portion  of  the  molecule  contributed 
to  the  binding,  and  generally  found  little  or  no  staining.  This  mercurial 
has  been  used  to  investigate  thiol  distribution  in  muscle  (Bennett,  1948  b), 
skin  (Mescon  and  Flesch,  1952),  and  a  variety  of  other  tissues  (Bennett, 
1951). 

Another  colored  mercurial,  4-mercuri-4'-dimethylaminoazobenzene,  has 
been  used  by  Horowitz  and  Klotz  (1956)  to  determine  protein  SH  groups. 
The  solubility  in  water  is  so  low  that  colorimetric  determinations  cannot 

4-Mercuri-4'-dimethylaminoazobenzene 

be  made,  but  it  dissoves  sufficiently  in  100  raM  glycine  (due  to  the  forma- 
tion of  a  glycinate  complex)  that  reactions  with  SH  groups  in  aqueous 
medium  can  be  carried  out.  However,  it  is  also  possible  to  determine  the 
amount  of  the  mercurial  removed  from  heptanol  when  shaken  with  an 
aqueous  solution  containing  the  protein,  although  equilibrium  usually  re- 
quires several  hours.  The  specificity  of  reaction  appears  to  be  satisfactory, 


768  7.  MERCURIALS 

since  the  amount  bound  to  bovine  seralbumin  increases  with  the  dye  con- 
centration until  the  molar  ratio  of  dye  to  protein  is  0.66,  following  which 
no  more  is  bound  although  the  dye  concentration  is  increased  50-fold.  This 
ratio  corresponds  quite  closely  to  the  known  SH  content  of  the  protein, 
lodination  of  the  seralbumin  prevents  the  reaction  with  the  mercurial. 
Ovalbumin  reacts  readily  with  two  of  its  SH  groups,  slowly  with  a  third, 
and  more  slowly  with  the  fourth,  the  dye  perhaps  differentiating  the  rel- 
ative reactivities  more  closely  than  does  p-MB.  This  method  has  a  high 
sensitivity  and  can  be  used  for  very  low  concentrations  of  protein. 

A  more  recently  examined  mercurial  dye,  4-(p-mercuriphenylazo)-l-naph- 
thylamine-7-sulfonate,  must  also  be  dissolved  in  glyine  buffer  (Nosoh,  1961). 
Absorption  at  470  m//  is  determined  and  the  titration  of  glutathione  and 
proteins  appears  to  be  quite  satisfactory. 

INHIBITION    OF  ENZYMES 

The  early  concept  of  the  mercurials  as  nonspecific  denaturing  and  coagu- 
lating agents  for  enzymes  has  gradually  been  abandoned  in  favor  of  a  pic- 
ture in  which  definite  and  often  isolatable  mercurial  complexes  are  formed 
under  the  proper  experimental  conditions.  A  selective  reaction  with  SH 
groups  on  enzymes  is  now  generally  assumed  and  the  mercurials  are  exten- 
sively used  for  the  detection  of  these  groups.  The  possibility  of  reaction  with 
other  than  SH  groups  has  been  discussed  (pages  737  and  753)  and  should 
never  be  ignored.  We  shall  note  instances  in  which  a  selective  action  on  SH 
groups  is  well  established,  and  a  few  examples  of  inhibition  not  involving 
SH  group.  We  shall  also  see  that  mercurial  inhibition  does  not  necessarily 
imply  an  SH  group  within  the  active  center  or  the  participation  of  an  SH 
group  in  the  catalysis.  In  this  connection,  it  is  well  to  bear  in  mind  the  dif- 
ferent groups  which  are  introduced  on  the  surfaces  of  enzymes  when  the 
different  mercurials  are  used  (Fig.  7-10),  inasmuch  as  the  steric  and  elec- 
trostatic effects  of  these  side  chains  may  be  critical  in  producing  inhibition. 

Crystalline   Mercuri-enzymes 

The  crystallization  of  the  mercuric  derivative  of  mercaptalbumin  was 
not  the  first  instance  of  such  a  procedure.  Warburg  and  Christian  (1941, 
1942)  introduced  this  technique  for  the  isolation  of  fermentation  enzymes 
and  obtained  the  crystalline  Hg-enolase  complex  from  yeast,  whereas  the 
normally  active  Mg-enolase  could  not  be  crystallized.  Kubowitz  and  Ott 
(1941)  in  Warburg's  laboratory  also  crystallized  the  Hg++  complexes  of 
lactate  dehydrogenases  from  Jensen  sarcomata  and  rat  muscle.  The  Hg++ 
complexes  in  all  cases  are  enzymically  inactive,  but  dialysis  against  cyanide 
solution  removes  the  Hg++  and  restores  the  activity.  There  is  no  better 
evidence  for  the  homogeneous,  stoichiometric,  and  reversible  Hg++  deriva- 


INHIBITION  OF  ENZYMES 


769 


tives  of  enzymes  than  such  complexes,  which  is  the  reason  they  are  discussed 
briefly  at  this  point.  Warburg  and  Christian  suggested  that  the  isolation  of 
mercuri-enzymes  might  be  generally  useful,  but  this  technique  either  was 
not  used  or  was  unsuccessful  until  Kimmel  and  Smith  (1954)  reported  the 
crystallization  of  mercuri-papain.  Krebs  (1930)  had  shown  that  papain  is 
very  sensitive  to  Hg++,  50%  inhibition  requiring  only  0.005  milf ,  and  this 


-H,hQ-C 


S-H,hQ 


S-H,^^S 


p-MPS 

Fig.  7-10.  The  side  chains  introduced 
onto  proteins  by  various  mercurials. 
The  S — Hg — R  bonds  are  not  actu- 
ally linear  but  are  shown  in  this  way 
for  convenience. 


indicated  that  a  tight  complex  is  formed  and  might  be  susceptible  to  crystal- 
lization. Twice  recrystallized  papain  (1.5-2%)  was  reacted  with  1  mM  Hg++ 
in  70%  ethanol  in  the  cold;  within  24  hr  a  precipitate  formed  and  in  3-4 
days  90%  of  the  activity  was  in  crystalline  form.  These  crystals  are  long 
rectangular  plates,  often  large  enough  to  be  visible  to  the  eye,  and  are 
soluble  in  water.  The  properties  of  mercuri-papain  have  been  reviewed  by 
Kimmel  and  Smith  (1957)  and  we  shall  discuss  only  those  aspects  relevant 
to  enzyme  inhibition. 

Mercuri-papain  contains  0.49%  Hg  and  has  a  minimal  molecular  weight 
of  41,400;  this  corresponds  to  1  Hg  atom  per  molecule  of  mercuri-papain. 
Since  the  molecular  weight  of  reduced  papain  is  around  20,500,  mercuri- 
papain  must  be  a  1  :  2  complex  or  dimer  to  be  represented  by  E — S — Hg — 


770  7.  MEKCURIALS 

S — E.  However,  the  situation  is  more  complex,  the  pH  being  an  important 
factor  in  determining  the  type  of  complex  occurring,  and  it  is  likely  that 
the  crystalline  mercuri-papain  is  the  least  soluble  form  of  several  possible 
derivatives  (Smith  et  al.,  1954  b).  Sedimentation  studies  at  pH  4  indicate 
a  monomer  or  1  :  1  complex,  probably  to  be  designated  by  HS — E — S — Hg+, 
while  at  pH  8  there  is  a  heavy  component  corresponding  to  a  hexamer, 
possibly  cyclic  with  alternating  — S — S —  and  — S — Hg — S —  bonds.  It  is 
interesting  that  there  are  two  electrophoretic  peaks  at  pH  4,  one  of  unit  + 
charge  greater  than  the  other;  since  dissociation  of  the  dimer  must  result 
in  equal  proportions  of  HS — E — SH  and  HS — E — S — Hg+,  this  would  tend 
to  confirm  the  dimeric  structure.  Oxidized  papain  is  a  mixture  of 

e(^\         and         E— S— S— E 

and  does  not  react  with  Hg++;  thus  it  is  very  important  in  studying  the 
combining  ratios  to  be  certain  that  the  papain  is  fully  reduced.  Mercuri- 
papain  is  actually  purer  than  papain,  as  indicated  by  electrophoretic  stud- 
ies, has  fewer  N-terminal  residues  detected  by  the  fluorodinitrobenzene 
technique  (Thompson,  1954),  and  has  some  10%  greater  activity  following 
removal  of  the  Hg++  with  cysteine  and  EDTA,  and  it  is  also  more  stable. 
The  proteolytic  enzyme,  pinguinain,  also  forms  stable  complexes  with  Hg++ 
which  are  stable  for  much  longer  times  than  the  pure  enzyme  (Messing, 
1961).  Other  enzymes  to  be  crystallized  as  the  mercury  complexes  are  a 
lysozyme  from  papaya  latex  (Smith  et  al.,  1955)  and  3-phosphoglyceralde- 
hyde  dehydrogenase  from  yeast  (Velick,  1953),  the  latter  after  reaction  with 
p-MB.  There  is  some  evidence  that  a  mercuric  dimer  of  ficin  occurs  (Liener, 
1961)  while  carboxy peptidase  forms  very  stable  Hg++  complexes  which  still 
possess  esteratic  activity,  although  they  no  longer  function  as  peptidases 
(Vallee  et  al.,  1961;  Coleman  and  Vallee,  1961).  There  is  thus  sufficient 
evidence  that  many  enzymes  form  well-characterized  mercurial  complexes 
and  are  quite  stable  in  this  state;  we  shall  note  other  examples  in  the  dis- 
cussion of  SH  titrations  of  enzymes. 

These  complexes  of  enzymes  with  Hg++  offer  strong  support  to  the  con- 
cept that  completely  selective  reaction  with  SH  groups  can  occur.  How- 
ever, if  Hg++  is  added  in  excess  of  that  required  for  mercaptide  formation, 
it  is  quite  possible  that  other  enzyme  groups  may  be  attacked.  It  is  likely 
that  other  enzymes  under  the  appropriate  conditions  can  form  dimers,  or 
other  polymers,  with  Hg++,  in  which  case  the  active  centers  may  be  made 
inaccessible  even  though  the  SH  group  is  not  within  the  confines  of  the  cen- 
ter. The  appearance  of  polymers  will  presumably  depend  strongly  on  the  pH 
since,  at  pH's  progressively  removed  from  the  isoelectric  point,  one  might 
expect  polymerization  to  be  more  and  more  reduced,  due  to  the  increasing 
charge  on  the  enzymes. 


INHIBITION  OF  ENZYMES  771 

Types  of  Inhibition  Observed  with  the  Mercurials 

The  concentration-inhibition  curves  for  mercurials  are  generally  sigmoid 
and  rather  steep,  as  would  be  expected  of  inhibitors  combining  tightly  with 
enzyme  groups.  Indeed,  when  such  curves  are  fairly  flat,  encompassing  sev- 
eral pi  units,  one  has  the  right  to  question  if  the  inhibition  is  related  to 
mercaptide  formation,  although  it  may  well  be.  It  should  be  emphasized 
that  adequate  kinetic  studies  can  be  made  only  in  preparations  of  pure 
enzymes.  The  presence  of  impurities  may  distort  the  entire  picture  and  the 
kinetics  of  inhibition. 

One  may  classify  the  inhibitions  classically  into  competitive,  noncompet- 
itive, uncompetitive,  and  mixed  types,  but  the  proper  plotting  procedures 
have  seldom  been  used  so  that  in  the  majority  of  cases  we  have  little  or  no 
information.  However,  sufficient  has  been  done  to  show  that  all  these  types 
of  inhibition  occur  (Table  7-8).  Competitive  behavior  has  been  observed 
in  a  surprisingly  large  number  of  instances.  This  is  surprising  at  first  if  one 
assumes  reaction  with  SH  groups  to  be  the  primary  mechanism  of  inhibi- 
tion, because  the  tightness  of  the  binding  might  be  considered  to  prevent 
the  exhibition  of  competition.  Actually,  most  inhibitions  by  mercurials  are 
probably  competitive  —  either  with  substrate,  coenzyme,  or  cofactor  —  in 
the  fundamental  sense  of  the  word,  but  it  is  often  difficult  to  demonstrate 
this  by  the  usual  analytical  techniques  which  assume  equilibrium  conditions. 
It  is  easier  to  show  that  the  presence  of  the  substrate,  coenzyme,  or  cofactor 
slows  the  development  of  the  inhibition,  although  the  equilibrium  inhibition 
may  not  be  detectably  different  (page  778).  Formally  competitive  behavior 
might  be  expected  to  occur  in  the  following  circumstances.  (1)  The  inhibitor 
acts  by  a  non-SH  reaction;  the  organic  mercurials  particularly  possess  group- 
ings capable  of  interacting  with  active  sites  independently  of  the  Hg  atom, 
and  such  might  be  involved,  for  example,  in  the  inhibition  of  D-amino  acid 
oxidase  by  p-MB,  the  benzoate  structure  being  of  primary  importance.  (2) 
The  binding  of  the  mercurial  to  the  SH  groups  may  for  some  reason  be 
weaker  than  usual  and  of  a  comparable  magnitude  to  the  affinity  for  the 
substrate.  (3)  The  mercurial  is  bound  much  more  tightly  than  the  substrate 
but  measurements  are  made  before  equilibrium  is  reached,  as  in  the  exper- 
iments showing  protection  of  the  enzyme  by  the  substrate;  when  the  inhi- 
bitions are  determined  soon  after  adding  the  mercurial  in  the  presence  of 
variable  concentrations  of  the  substrate,  the  data  may  provide  formally 
competitive  plots.  One  would  expect  this  third  explanation  in  certain  exam- 
ples given  in  Table  7-8,  e.g.,  carbonic  anhydrase,  where  KJK,,^  =  3.87  X  10^' 
for  ??-MB  (Chiba  et  al.,  1954  b).  In  the  case  of  homogentisate  oxidase,  p-MB 
and  MM  inhibit  competitively  with  respect  to  Fe++  but  noncompetitively 
with  respect  to  homogentisate,  the  mercurials  being  bound  roughly  40-100 
times  as  tightly  as  the  Fe++  (Flamm  and  Crandall,  1963).  Here,  and  in 
other  instances  where  metal  ion  cof actors  are  involved,  both  cofactor  and 


772 


7.  MERCURIALS 


P5 


^  ^  o>   S 


-^  ^H  ^ —  -^ 

s  _  —  ^ 

-2  n3  -=5    >=* 

Dh  W  M  o 


j  Cm  o  <: 


a  -2 

S  o 

=5     S  .^  ® 

c3     0)  V  ;: 

t-       (O  >  S 

«     ea  Qj  •« 

to     c  «  -^ 

==   S  fe  -^ 

O     S  O  cS 

§  ffi  W  tf 


>. 

^ 

X 

c 

e3 

o 

4) 

S 

-Q 

S3 

? 

u 

0 

M 

t 

-s 

2 

V 

J3 

73 

'O 

>> 

H) 

a> 

-C 

-d 

Oi 

c 

c 

10 

cS 

eS 

Oi 

w 

S 

-ri 

rS 

c8 

,^  "S 

h4 

J3    ^^  _cS 


cS    05    F^ 


N     00 

O   !5 


2    o 


05     B 


1  s 


tf 


M    O     _ 
>H    O    O 


2    -^    — ' 


w 


fe 


-S 

J 

83 

eS     © 

A 

-C    -S 

A 

ft  -2 

<n 

1 

ft  .5 

0 

JS 

P. 

-t^ 

^^   0   „ 

>i 

f^  0    ® 

s 

c  ■?  "P 

4:> 

V 

S    c    >> 

■tS 

^ 

-£3    ^  JS 

ce 

p. 

P<   ^    iJ       ^^ 

c 

0 

0    ^    T5     2 

0 

u 

iH     01    "TS     C 

A 

J2 

J,^ 

i  ^  1  "5 

Q 

0 

a  fe;  <:  <1 

2 

05 

0 

« 


a  t3  S 

-^  •r  'S  !s  iJ 

I  ^  ^  ft  S 

o    o  ft  J=r  o 

TT      tn  o  O  :3 

<t3    a  (^ 


CO  -H 


h5         2Q  Ph 


> 

S 

;s 

^ 

M 

M 

s 

S 

cj    .™   .Of 


lis 

-      4)      o 

HI 

Oi  W  CQ 


P5 


73     « 


s  B, 


O   yA% 


M 

% 

^ 

§ 

m 

0 

cd 

<a 

V 

« 

'« 

-C 

00 

00 

CO 

c« 

eS 

cS 

0 

ft 

3 

75, 

"V 

«J 

0 

^ 

'S 

'x 

s 

•§ 

J3 
ft 

V 

ft 

0 

0 

0 

00 

« 

S 
*?■» 

0 

j3 

0; 

T3 

0 

C 

S. 

_C 

•^ 

ft 

>> 

!>. 

^ 

g 

^ 

^ 

-C 

J3 

'g 

0 

c3 

^ 

^ 

4) 

OJ 

>v 

8 

^ 

'0 

-0 

-0 

<! 

% 

ti 

< 

0 

<: 

^ 

< 

Q 

TT       r-       Q^       '-' 

sl-s  § 


3 
3 


0 

»c 

n 

OS 

CD 

Oi 

■^^ 

1—4 

^— ' 

a 

^ 

--^ 

ci 

"d 

CO 

CD 

a 

-c 

05 

^ 

c 

0 

c3 
t4 

^^ 

M 

0 

^ 

a 

t3 

Ti 

C 

fl 

^ 

c8 

cS 

s 

s 

1 

0 

a 

s 

-2 

■5 

c3 

c 

>> 
X 
0 

0 

-S 

i 

+ 

>> 

0 

t4 

+ 

W 

0 

fe 

«c 

GO 

> 

b 
« 
_> 

-^ 

Cm 

<«4 

03 

Is 

4> 

<» 

0 

e 

CO 

'x 

0 

tH 

0^ 

9i 

d 

ci 

^ 

'x 

'S 

0 

2 

^ 

xi 
-1-2 

eS 

c 

_tE 

c8 

'■3 

>. 

c 

X 

aj 

0.1 

0 

tc 

0 

-S 

o3 

s 

0 

>. 

w 

> 

M 

eo 

C 
1— 1 

-a 

X 

X 

a 

"ft 

-*j 

a 

c8 

0 

j3 

« 

fS 

4) 

ft    b£ 


o  *^ 


M 


CH    I— I 


INHIBITION  OF  ENZYMES  773 

mercurial  are  bound  to  the  same  SH  group.  In  most  cases,  increase  in  the 
substrate  concentration  does  not  reduce  the  inhibition  once  established,  but 
Robert  et  al.  (1952)  claim  that  acetylcholine  is  able  to  displace  Hg++  from 
horse  serum  cholinesterase,  the  affinity  of  the  enzyme  for  the  acetylcholine 
actually  being  greater  than  for  the  Hg++;  it  is  not  known  if  mercaptide 
formation  is  involved.  One  suspects  that  the  inhibition  may  sometimes  ap- 
pear to  be  competitive  where  actually  the  mercurial  is  reacting  with  the 
substrate,  either  exclusively  or  in  addition  to  the  enzyme  (compare  curves 
in  Figs.  1-5-1  and  1-5-14),  and  such  might  be  the  case  in  the  inhibition  of 
/^-amylase  by  p-MB  and  PM  (Ghosh,  1958),  although  the  extent  of  reaction 
of  mercurials  with  starch  is  not  known. 

Noncompetitive  inhibition  may  be  observed  when  the  mercurial  reacts 
with  groups,  SH  or  other,  adjacent  to  the  active  center,  and  thus  suppresses 
the  rate  of  breakdown  of  the  ES  complex,  and  when  the  affinity  for  the 
enzyme  is  not  so  high  that  mutual  depletion  kinetics  hold.  The  interference 
with  ES  breakdown  may  be  steric  through  the  side  chains  introduced  or 
secondarily  by  alteration  of  the  protein  structure.  It  must  be  remembered 
that  mutual  depletion  systems  usually  indicate  formally  noncompetitive 
behavior  if  the  common  plotting  procedures  are  used  (compare  Figs.  1-5-3 
and  1-5-24),  despite  the  fact  that  the  inhibition  may  be  fundamentally 
competitive.  No  pure  instances  of  uncompetitive  or  coupling  inhibition  have 
been  reported,  but  it  is  not  unlikely  that  preferred  reaction  with  the  ES 
complex  occurs.  The  inhibition  of  alkaline  phosphatase  is  actually  mixed 
(noncompetitive  and  uncompetitive),  but  p-MB  reacts  more  readily  with 
the  ES  complex  {K,'  =  0.163  mM)  than  with  E  {K,  =  2.5  milf)  (Lazdunski 
and  Ouellet,  1962).  There  are  some  examples  in  which  such  reaction  with 
the  ES  complex  is  possible,  e.g.,  the  inhibition  of  urease  by  Hg++  (Evert, 
1952),  of  acid  phosphatase  by  p-MB  (Newmark  and  Wenger,  1960),  and  of 
succinate  dehydrogenase  by  p-MB  (Warringa  and  Giuditta,  1958).  The  in- 
hibition of  lactate  dehydrogenase  from  Propionibacterium  pentosaceum  by 
p-MB  is  greater  in  the  presence  of  lactate  than  when  no  substrate  is  present 
during  incubation,  and  this  was  postulated  to  be  due  to  the  greater  number 
of  free  SH  groups,  presumably  arising  through  reduction  by  lactate  (Moli- 
nari  and  Lara,  1960),  and  a  similar  situation  may  occur  with  glutathione 
reductase  and  NADPH,  the  presence  of  the  reduced  coenzyme  increasing 
the  inhibition  markedly  (Mapson  and  Isherwood,  1963). 

Another  approach  to  the  classification  of  mercurial  inhibitions,  and  per- 
haps the  primary  one,  is  the  determination  of  the  component  —  enzyme, 
substrate,  coenzyme,  or  cofactor  —  with  which  the  mercurial  reacts.  Reac- 
tion with  the  apoenzyme  has  generally  been  assumed  above  and  with  respect 
to  the  molecular  mechanism  might  be  divided  into  three  types:  (1)  binding 
to  an  SH  group  at  the  active  center,  preventing  complexing  of  the  apo- 
enzyme with  any  of  the  other  components,  (2)  binding  with  an  SH  group 


774  7.  MERCURIALS 

vicinal  to  the  active  center  and  interference  with  the  catalysis  sterically  or 
electrostatically,  and  (3)  secondary  altering  of  the  protein  structure  to  dis- 
rupt the  normal  configuration  of  the  active  center.  In  the  last  case,  which 
is  probably  fairly  common  (see  page  787),  the  inhibition  may  be  formally 
competitive  (if  the  substrate  stabilizes  the  enzyme  structure),  noncompeti- 
tive, or  quite  complex.  Reaction  imth  the  substrate  must  often  occur,  espe- 
cially when  the  substrate  is  protein,  nucleic  acid,  nucleotide,  or  thiol,  but 
in  most  cases  this  possibility  seems  to  have  been  ignored.  It  is  obvious  for 
glutathione  reductase  and  this  complicates  the  analysis  of  the  inhibition 
(Mapson  and  Isherwood,  1963),  but  it  may  also  be  an  important  mechanism 
when  thioesters  are  involved,  e.g.,  acetoacetyl-CoA  in  fatty  acid  synthesis 
(Stern,  1956)  or  malonyl  semialdehyde  pantetheine  in  propionate  metabol- 
ism (Vagelos  and  Earl,  1959).  The  inhibition  of  NADPH:  methemoglobin 
oxidoreductase  by  p-MB  occurs  when  either  the  enzyme  or  the  methemo- 
globin is  incubated  with  the  mercurial  (Bide  and  Collier,  1964).  Sometimes 
one  finds  indirect  evidence  for  reaction  with  the  substrate,  as  with  5'-aden- 
ylate  deaminase  (Lee,  1957).  Here  the  inhibition  by  p-MPS  is  much  greater 
when  it  is  preincubated  with  adenylate  and  the  reaction  started  by  adding 
the  enzyme  than  when  preincubation  is  with  the  enzyme  and  reaction  start- 
ed by  adding  the  substrate.  Reaction  with  coenzymes  is  evident  when  lipoate 
or  coenzyme  A  is  involved,  but  may  be  more  general  than  is  usually  sup- 
posed. A  reaction  of  -p-MB  with  NAD  was  detected  spectrophotometrically 
by  Palmer  and  Massey  (1962)  and  this  was  considered  to  be  significant  in 
titrations  of  certain  dehydrogenases.  Hill  (1956)  had  previously  established 
a  1  :  1  complex  of  Hg++  with  NADH,  but  had  found  no  complex  with  p-MB. 
Onrust  et  al.  (1954)  considered  the  possibility  that  at  least  part  of  the  inhi- 
bition of  pyruvate  oxidase  by  p-MB  might  be  due  to  reaction  with  the  sul- 
fur of  thiamine-diP,  but  excluded  this  when  they  found  that  thiamine-diP 
does  not  reverse  the  inhibition.  However,  Pershin  and  Shcherbakova  (1958) 
observed  that  thiamine  is  able  to  reduce  the  bacteriostatic  action  of  Hg++, 
although  this  could  be  by  a  mechanism  other  than  reaction  of  the  thiamine 
with  Hg++.  Kuratomi  (1959),  on  the  basis  of  preincubation  experiments 
with  components  of  the  pyruvate  oxidase  system,  postulated  that  j'-MB 
can  react  with  thiamine-diP.  This  problem  remains  to  be  settled  and  possibly 
is  an  important  one.  It  would  be  interesting  to  know  if  mercurials  can  open 
the  thiazole  ring  under  physiological  conditions  (which  is  not  likely)  or 
react  with  the  SH  groups  after  ring  opening,  in  which  case  the  state  of  the 
thiamine-diP  in  the  preparation  would  be  important.  Another  possibility 
is  that  a  complex  is  formed  with  groups  other  than  the  sulfur  since  oppor- 
tunities for  chelation  exist. 

It  may  be  suggested  that  in  all  studies  of  enzyme  inihibition,  in  which 
substrates  or  coenzymes  capable  of  reacting  with  mercurials  are  involved, 
the  appropriate  preincubations  with  the  inhibitor  be  carried  out,  as  pre- 


INHIBITION  OF  ENZYMES  775 

viously  described  (page  1-569),  since  this  technique  will  often  provide  in- 
formation on  complexes  formed  with  components  other  than  the  apoenzyme. 
Whatever  the  mechanism  or  formal  type  of  inhibition  by  mercurials,  it 
is  certain  that  many  systems  must  be  represented  by  mutual  depletion 
kinetics.  This  is  clearly  seen  in  many  of  the  enzyme  titrations  (page  804), 
inhibition  being  produced  by  mercurials  at  roughly  equimolar  concentra- 
tions relative  to  the  enzymes,  but  at  this  point  the  problem  will  be  treated 
in  a  more  general  manner.  Mutual  depletion  behavior  implies  that  the  inhi- 
bition will  depend  on  the  concentration  of  the  enzyme.  This  is  seen  with 
yeast  pyruvate  decarboxylase  in  the  work  of  Stoppani  et  al.  (1953)  (see 
accompanying  tabulation),  and  even  more  markedly  with  pig  heart  suc- 


Pyruvate  %  Inhibition  by: 


aecarooxyiase 

(/'g/ml) 

Hg++ 

p-MB 

7.8 

85.0 

95.0 

15.7 

— 

75.0 

30.5 

43.0 

33.0 

01.0 

15.9 

7.0 

cinate  oxidase,  which  is  inhibited  89%  by  0.01  mM  p-MB  when  the  enzyme 
concentration  is  0.15  mg/ml  but  only  59%  by  0.76  mM  p-MB  when  the 
enzyme  concentration  is  30  mg/ml  (Stoppani  and  Brignone,  1957).  Another 
example  is  muscle  p>Tuvate  oxidase  (see  accompanying  tabulation)  (Onrust 


Enzyme  extract         „^  inhibition  by  p-MB  0.11  mM 
(ml) 


0.4  82 

0.8  56 

1.2  34 

1.5  33 


et  al.,  1954).  These  few  examples  well  illustrate  the  importance  of  this  factor 
and  very  clearly  demonstrate  the  quantitative  meaninglessness  of  most 
reported  inhibitions  if  the  relative  enzyme  concentration  is  not  known  or 
stated.  Impurities  also  may  contribute  to  the  depletion  of  the  mercurial. 
The  crude  bacterial  enzyme  for  converting  histidinol  to  histidine  is  not 
inhibited  by  0.02  mM  p-MB,  but  the  partially  purified  enzyme  is  inhibited 
50%  (Adams,  1954),  and  it  is  likely  that  the  pure  enzyme  would  be  inhib- 


776  7.  MERCURIALS 

ited  even  more  strongly;  such  work  points  out  the  importance  of  enzyme 
purity  for  accurate  studies  of  mercurial  inhibition.  The  elevation  of  the 
plgo  from  0.0002  mM  to  0.014  uiM  by  serum  for  the  inhibition  of  3-phos- 
phoglyceraldehyde  dehydrogenase  by  p-MB  is  a  further  example  (Weitzel 
and  Schaeg,  1959). 

When  an  enzyme  is  reported  to  be  inhibited  to  a  specified  degree,  say 
50%,  by  a  certain  concentration  of  mercurial,  exactly  how  is  this  to  be 
interpreted?  Is  50%  of  the  enzyme  combined  with  the  mercurial  in  a  com- 
pletely inactive  EI  complex,  or  is  all  the  enzyme  combined  with  the  mer- 
curial and  the  EI  complex  possesses  50%  of  the  original  activity?  If  the 
ordinary  equilibrium  formulation  is  followed  and  it  is  assumed  that  the 
fractional  activity  of  the  EI  complex  is  r,  noncompetitive  inhibition  will 
be  given  by 

.^    (l-r)(I) 
(I)  +  Ki 
and 


(7-4) 


1  Ki 

+  -r. -^;vr  (7-5) 


i  (1-r)  (l-r)(I) 

so  that  a  plot  of  \ji  against  1/(1)  will  give  a  straight  line  intersecting  the 
1/i  axis  at  1/(1  —  r),  or  \jimax-  If  mutual  depletion  occurs  (zone  C),  a  sim- 
ilar result  is  obtained,  although  the  slope  will  be  different.  A  simple  plot 
of  this  type  may  help  to  decide  between  the  two  possibilities  above.  If  the 
plot  is  not  linear  near  the  \ji  axis,  one  might  suspect  that  another  type  of 
inhibition  is  occurring  at  higher  inhibitor  concentrations,  or  that  secondary 
inactivation  of  the  enzyme  is  a  factor. 

One  example  of  the  deviations  from  classic  inhibition  kinetics  that  may 
be  seen  with  the  mercurials  is  the  inhibition  of  human  plasma  cholinesterase 
by  Hg++  as  analyzed  by  Goldstein  and  Doherty  (1951 ).  This  slowly  develop- 
ing, pH-  and  temperature-dependent  inhibition  presents  some  interesting 
but  often  uninterpretable  results.  The  l/v-l/(S)  plots  exhibit  two  sorts  of 
deviation  (Fig.  7-11).  The  results  from  long  incubation  with  low  concentra- 
tions of  Hg++  fall  on  reasonably  straight  lines  (A  and  B),  but  the  slopes  are 
a  good  deal  greater  than  expected  for  pure  noncompetitive  inhibition,  as 
for  mixed  inhibition  (Fig.  I-5-6A)  the  interaction  constant  a  being  some 
finite  value  >  1.  Of  course,  it  may  not  actually  be  true  mixed  inhibition, 
the  deviation  being  due  to  some  other  factor.  The  results  from  short  in- 
cubations with  high  concentrations  of  Hg++  differ  so  much  from  any  sort 
of  classic  behavior  that  it  is  impossible  to  interpret  them  (C  and  D).  It  was 
suggested  that  low  and  high  concentrations  of  Hg++  inhibit  by  different 
mechanisms,  possibly  with  different  SH  groups,  the  former  with  groups 
outside  the  active  center  causing  secondary  irreversible  inactivation  and 
the  latter  directly  with  groups  in  the  active  center.  This  would  to  some 


INHIBITION  OF  ENZYMES 


777 


extent  explain  why  the  inhibition  is  more  competitive  at  high  Hg++  and 
more  noncompetitive  at  low  Hg++,  but  it  does  not  explain  the  deviations 
discussed  above.  Curves  C  and  D  presumably  do  not  represent  equilibrium 
inhibitions  and  are  more  illustrative  of  protection  of  the  enzyme  by  the 
substrate;  it  would  seem  that  acetylcholine  above  50  mM  protects  the  en- 
zyme almost  completely  against  very  short  exposures  to  Hg++,  which  is 
not  too  unreasonable  considering  the  relatively  high  affinity  of  the  enzyme 
for  acetylcholine.  Although  the  kinetics  of  protection  and  the  application 


2000- 


1500- 


1000- 


500 


. 

A 

0  0126  mM    FOR    4  5  HOURS 

/ 

B 

O0Z73mM    FOR    4    HOURS 

/ 

C 

2  28  mM    FOR    20  MIN 

^X          / 

0 

4  S4  mM    FOR    3  MIN 

/        / 

/c 

^/      /      _J— — 

CONTROL 

20 


60 


Fig.  7-11.  Double  reciprocal  plots  for  the  inhi- 
bition of  human  plasma  cholinesterase  by  Hg++, 
showing  deviations  from  linearity  at  high  Hg++ 
concentrations.  (Modified  from  Goldstein  and 
Doherty,   1951.) 


to  plotting  procedures  have  never  been  worked  out  as  far  as  I  know  —  and 
it  would  be  difficult  to  treat  the  phenomenon  rigorously  one  might  predict 
that  curves  with  rather  steep  slopes  in  the  l/v-l/(S)  plot  would  be  found,  and 
that  such  curves  would  occasionally  intersect  the  control  curve  to  the  right 
of  the  \\v  axis,  i.e.,  the  longer  the  incubation,  the  closer  to  equilibrium 
would  the  inhibition  come,  and  the  less  competition  or  protection  would 
be  exerted  by  the  substrate.  One  also  wonders  if  the  increased  tilt  of  curves 
A  and  B  might  be  due  to  the  fact  that  these  relatively  high  substrate  con- 
centrations protect  the  active  center  against  structural  changes  brought 
about  by  reaction  of  the  Hg++  at  vicinal  sites,  since  there  are  many  examples 


778  7.  MERCURIALS 

in  which  the  substrate  can  slow  down  spontaneous  or  induced  enzyme  de- 
naturation.  However,  neostigmine,  which  can  protect  chohnesterase  against 
thermal  denaturation,  does  not  protect  at  all  against  Hg++.  There  is  ac- 
tually some  doubt  as  to  whether  the  inhibition  is  related  to  SH  groups, 
since  p-MB  and  MM  up  to  1  raM  do  not  inhibit  even  after  2  hr  at  37^,  or  it 
might  mean  that  the  reacting  SH  groups  are  not  at  the  active  center  and 
the  inhibition  by  Hg++  is  due  to  a  dimerization  or  polymerization.  If  all 
enzymes  subjected  to  mercurials  were  studied  in  as  much  detail  as  in  this 
work,  there  would  probably  be  many  more  interesting  examples  of  devia- 
tions from  classic  theory;  as  long  as  one  tests  an  enzyme  under  standard 
conditions  with  one  concentration  of  a  mercurial,  as  is  done  in  most  reports, 
interpretation  presents  no  problems. 

Protection  of  Enzymes  against  Mercurials 

Enzymes  may  be  protected  against  mercurials  by  (1)  substrates,  (2)  co- 
enzymes, (3)  metal  ion  cofactors,  (4)  reversible  inhibitors,  and  (5)  thiols  or 
other  mercurial  complexers.  Various  conclusions  have  been  drawn  from 
such  experiments,  mainly  regarding  the  relation  of  SH  groups  to  the  bind- 
ing of  the  protector,  but  there  are  many  pitfalls;  the  discussion  of  protection 
with  respect  to  iodoacetate  (page  47  and  Fig.  1-5  in  Volume  III)  applies 
equally  well  to  the  mercurials.  Protection  may  occur  by  two  general  mech- 
anisms: reaction  of  the  protector  with  the  enzyme  to  block  off  the  mer- 
curial, or  reaction  of  the  protect&r  with  the  mercurial.  The  latter  mechanism 
applies  to  the  thiols  such  as  cysteine  or  glutathione,  which  have  been  widely 
used  for  this  purpose,  but,  as  has  been  pointed  out  several  times,  such  pro- 
tection does  not  provide  much  useful  information,  since  in  reality  all  one 
does  is  to  reduce  the  effective  mercurial  concentration.  It  also  applies  to 
other  complexers  and  perhaps  is  involved  in  the  following:  the  protection 
of  fumarate  hydratase  (Mello  Ayres  and  Lara,  1962)  and  fumarase  (Fave- 
lukes  and  Stoppani,  1958)  by  phosphate,  of  ascorbate  oxidase  by  amino 
acids  and  RNA  (Frieden  and  Maggiolo,  1957),  of  acid  phosphatase  by  EDTA 
(Macdonald,  1961),  of  urease  by  ascorbate  (Mapson,  1946),  and  of  thyroxine 
delahogenase  by  FMN  (Tata,  1960).  However,  in  these  cases  it  is  often 
difficult  to  interpret  the  mechanism  of  the  protection.  We  shall  not  be  con- 
cerned with  this  type  of  protection,  but  only  with  those  protectors  presum- 
ably reacting  with  the  enzyme. 

Some  examples  of  protection  are  summarized  in  Table  7-9  along  with 
instances  in  which  protection  does  not  occur  (or  at  least  is  not  observed 
under  the  conditions  used).  The  +  sign  does  not  indicate  that  complete 
protection  can  be  achieved;  indeed,  in  most  cases  only  partial  protection 
has  been  reported,  and  this  is  what  we  would  expect.  The  degree  of  pro- 
tection may  depend  on  the  concentrations  of  mercurial  and  protector;  e.g., 
protection  may  be  complete  with  low  mercurial  concentrations,  whereas 


INHIBITION  OF  ENZYMES  779 

the  protector  may  be  relatively  ineffective  against  high  concentrations, 
as  in  the  effects  of  arginine  on  the  inhibition  of  its  oxidative  decarboxylation 
(see  accompanying  tabulation)  (Van  Thoai  and  Olomucki,  1962).  In  most 


%  Inhibition 


p-MB 


(mM) 

0.033 

57 

0.05 

66 

0.067 

89 

0.083 

98 

p-MB  alone  p-M.B  +  arginine  10  mM 


0 

0 

43 

61 


reports  it  is  difficult  to  decide  if  the  protection  is  simply  due  to  a  slowing 
of  the  rate  of  inhibition  or  to  a  true  effect  on  the  final  equilibrium  inhibi- 
tion, since  measurements  are  often  made  over  arbitrary  time  intervals.  It 
is  evident  that  it  is  easier  to  slow  down  an  inhibition  than  to  modify  its 
final  level;  enough  substrate,  coenzyme,  or  cofactor  to  saturate  the  enzyme 
substantially  will  quite  markedly  slow  the  reaction  of  the  enzyme  with 
the  inhibitor,  but  the  final  inhibition  need  not  be  significantly  changed, 
particularly  with  the  mercurials  which  are  usually  bound  tightly,  if  slowly. 
Most  investigators  have  noted  that  although  the  protectors  in  Table  7-9 
are  effective  when  present  during  the  development  of  the  inhibition,  they 
do  not  reverse  the  inhibition  at  all  once  it  has  reached  a  steady  level,  this 
apparently  indicating  that  most  of  the  protection  results  are  fundamentally 
due  to  a  slowing  of  the  rate  of  inhibition. 

Occasionally  two  components  of  the  enzyme  reaction,  forming  a  ternary 
complex  with  the  enzyme,  protect  more  than  each  component  alone.  This 
is  the  situation  with  malate  oxidative  decarboxylase,  the  protections  by 
malate  and  Mn++,  or  malate  and  NADP,  being  additive;  the  protections 
by  Mn++  and  NADP  are  not  (Rutter  and  Lardy,  1958).  It  may  also  be  the 
case  with  liver  alcohol  dehalogenase,  ethanol  and  NAD  protecting  more 
than  either  one  alone  (Yonetani  and  Theorell,  1962).  In  one  situation,  aspar- 
tate carbamyltransf erase,  neither  substrate  alone  protects,  but  together  they 
do  so  quite  effectively  (Reichard  and  Hanshoff,  1956).  An  example  of  pro- 
tection by  a  reversible  inhibitor  is  the  reduction  in  the  inhibition  of  succinate 
dehydrogenase  by  p-MB  or  Hg++  in  the  presence  of  oxalacetate  (Stoppani 
and  Brignone,  1957).  Actually,  an  effective  competitive  inhibitor  might  be 
expected  to  protect  better  than  the  substrate. 

The  information  derived  from  protection  experiments  is  frequently  not 
as  reliable  as  commonly  assumed,  for  reasons  to  be  discussed  in  Chapter  1, 
Volume  III.  The  fact  that  the  action  of  a  mercurial  is  reduced  by  a  sub- 


780 


7.  MERCURIALS 


C^ 


'-'■'        aj 


it       oi 


C 

'2 

& 
ft 


C5        5 


. — . 

2 

CD 

© 

lO 

o 

'*-' 

05 

Oi 

^ 

^H 

^ 

^-^ 

s 

te 

i 

o 

o 

CO 

O 

O 

1 

-d 

Ti 

w 

C3 

o3 

^ 

Lh 

,^ 

eS 

0) 

eS 

>> 

O 

V 

lU 

^ 

ti 

s 

H 

^ 

P3 


+ 


+    +  I  + 


+  I     +  + 


+ 


fu 


2  o 


fL, 


n      '^  P^ 


«  Q  Q  Q 
<  <  <  < 


fe 


JS        cS 


m     m 


pq 


PS 


pq 


m 

s 
i. 


P3  m     m 


p;     PQ 


a 


Ah 

Q 

'o 

s 

<'1 

ft 

_> 

^ 

CO 

to 

+3 

5 

(4H 

o 

o3 
4J 

S 

^ 

w 

>^ 

P5 

ei5 


s 

d 

<u 

M 

O 

-S 

>> 

jS 

« 

-d 

V 

-o 

<U 

43 

CO 

-2 

CI 

"o 

T) 

73 

-s        c    ?? 


fe)         GQ         fe) 


s 


eS 

CO 

73 

C 

1 

1 

>. 

X 
O 

>> 

a 

^ 

cS 

^ 

INHIBITION  OF  ENZYMES 


781 


lO  '— 


« 


o: 

-2 

C 

c3 

"e 

"3 

-3 

-a 

C 

"S 

s 

<s 

(Z3 

cS 

^H 

-a 
S 

>. 
^ 

fe 

M 

-2 

in 

o 

c 

01 

6 

cS 

§ 

^ 

^ 

CO      ^       _, 


>. 
«< 


m 


o 

C 

«  g 


+     +++++     +     +     +++     + 


+    + 


+ 


+ 


73 


S 


-8 


2  Q 

w  12; 


o 


C     tS 


O 


C 

o 

^  H 
O  <1 


W  M         pq     pq     pq     pq         pq 

§  §         g     S     S     S         g 

?S.  SI,  ?4,         a         ?>.         ?l.  ?^ 


t    m    P3 


pq 


pq 


pq 


pq 


pq 


an 

o 

Si 

Si 

o 

-Q 

C 

s 

cS 

v 

ft 

^ 

"S 

_o 

» 

S 

2q 

o 

S     K-i   fe;    >^ 


^ 

fi 

a> 

o 

V, 

a 

« 

3 

c3 

s 

M 

s 

pq 


a:       .s       ^ 


S 

fi 

O 

s 

M 

o 

o 

a 

>% 

>> 

p 

tH 

^ 

S-i 

w 


2  S 

i 

It 

3 

ce  -c 

O  ^ 

o 


^1 


782 


7.  MERCURIALS 


P5 


(^      C^ 


^        6 


O 


^     2 


o 


§      05 


c6 

03 

t3 

K 

^-^ 

t3 

"e 

e« 

"^ 

"^ 

fl 

a> 

Xi 

« 

-« 

<D 

^ 

cS 

M 

Q 

H 

Cm 


+ 


+  ++      +      +++        + 


+     +  +     + 


+ 


C^ 


X 

o 

o  O 

>><1 

CM 

O  Z 

O       N       W  fq 


Wl 


2  i^ 

^'     ^« 

^  o 

W  -g  <t; 

U-^ 

oi.  ^  |2; 

W  !^ 

+ 


m     pq     fQ 


m     M    w 


M 


pq 


m 


=«    s 


^    § 


m 


m 


fM 


w     M 


s    ^ 


o     o    ^ 


cS        c 


Ph    w    w 


>> 

y. 

o 

« 

t-i 

M 

T3 

c3 

>. 

^ 

a 

"S 

O 

3 

eS 

T3 

C 

© 

« 

(H 

M 

S 

4) 
> 

1 

a 

3 

-C 

iH 

ii 

OJ 

f>> 

T3 

ft 

-2 

1 

1 

4> 

>i 

'o 

W 

CO 

1— 1 

J3 


INHIBITION  OF  ENZYMES  783 


05         rt  -^ 


CO 


CI 

O    — ' 

W 

2  "e 

c 

-^ 

eS 

5 

O 

-2 

c  ^ 

^J 

^^ 

3 

3 
P3 

^     ^ 


. — .  O  C5 

K       [p        c;- 


K     2 


ffit-ji:^  (25  mmWWSwm  SQ 


t3 

e 

■» 

d 

cS 

^ 

c 

X 

ft 

iS 

-ii 

ft 

D5 

w. 

O 
02 

g 
O 

^ 

c 
o 

c« 

«D 

.Sf 

P 

C5 

PQ 

.» 

fO 

T5 

' — ^ 

lO 

t*.wa 

c 

CO 

Oi 

e 

c3 

Tt^ 

3i 

>» 

C 

*_^ 

u 

cS 

Eh 

% 
■A 

M 

ft 
ft 
O 

w. 

+  +  ++   +  +  +   +    I    ++      +    I    +   +    I+  + 


o 

to 


c    2      s 
"3    c      s^ 


>      »     c3 


fi 

73 

cS 

>> 

(H 

j3 

-4^ 

cS 

4> 

^ 

o 

GO 

t-( 

^ 

a 

c 

o 

"m 

® 

o 

M 

c 

■ft 

O 

O 

1- 

^  ^ 


-c 

"tt 

cS 

"S 

(-1 

>> 

CO 

2  oi  ?^-1         11 


pamw  p^  pqpqm  pq  pqpqpQpqpspqpq  CQM 


u      •->     t; 


42        42        *f  S     QJ)        cS         fl 


784  7.   MERCURIALS 

strate,  for  example,  does  not  necessarily  imply  that  the  substrate  reacts 
with  an  SH  group  nor  that  the  SH  group  is  involved  in  the  catalysis,  al- 
though these  may  well  be  the  case.  A  positive  result  is  more  valuable  than 
a  negative  one.  The  failure  to  achieve  protection  may  be  due  to  an  inade- 
quate concentration  of  the  protector,  too  low  a  relative  affinity  of  the  en- 
zyme for  the  protector,  or  a  long  incubation  wherein  equilibrium  is  reached, 
and  yet  the  substance  examined  may  participate  in  the  reaction  and  inter- 
act with  the  enzyme  in  the  same  way  as  effective  protectors.  Definite  pro- 
tection allows  one  to  make  the  reasonable  assumption  that  the  mercurial 
binds  somewhere  in  the  region  occupied  by  the  protector.  Potter  and  Du- 
Bois  (1943)  postulated  an  SH  group  to  be  located  between  the  two  cationic 
groups  binding  succinate  to  the  dehydrogenase;  protection  against  mercuri- 
als by  succinate  simply  implies  that  succinate  is  able  to  shield  this  SH 
group  from  the  mercurial,  and  not  that  the  SH  group  is  involved  in  the 
succinate  binding  or  participates  in  the  oxidation-reduction  reaction.  For 
steric  reasons,  the  smaller  the  molecular  sizes  of  the  protector  and  the  mer- 
curial, the  more  certain  can  one  be  that  a  common  SH  group  is  involved 
in  the  binding  of  both. 

Displacement  of  Coenzymes  and  Cofactors  from  Enzymes 

Closely  allied  to  protection  experiments  are  those  in  which  a  mercurial 
is  shown  to  dissociate  an  enzyme-coenzyme  or  enzyme-cofactor  complex. 
It  is  now  believed  that  several  coenzymes  and  metal  ion  activators  may  be 
bound  to  apoenzymes  through  SH  groups  in  part  (Shifrin  and  Kaplan, 
1960),  and  if  this  is  so  one  would  expect  tightly  bound  SH  reagents,  such 
as  the  mercurials,  to  displace  the  coenzymes  or  activators.  Certain  coen- 
zymes or  cofactors,  such  as  NAD,  have  been  shown  to  react  with  thiols 
(van  Eys  and  Kaplan,  1957  b),  but  in  most  cases  the  evidence  for  binding 
to  enzyme  SH  groups  is  circumstantial.  Certainly  such  displacement  of 
necessary  components  of  the  enzyme  reaction  would  be  an  important 
mechanism  in  the  inhibition  produced  by  mercurials,  especially  in  vivo  where 
the  total  binary  or  ternary  complexes  usually  occur.  One  molecule  of  crys- 
talline horse  liver  alcohol  dehydrogenase  binds  2  molecules  of  NADH  at 
physiological  pH  and  this  is  accompanied  by  a  shift  in  the  absorption 
spectrum  of  the  NADH.  Addition  of  p-MB  was  found  by  Theorell  and 
Bonnichsen  (1951),  to  reverse  the  spectral  shift,  and  it  was  concluded 
that  the  bond  between  an  enzyme  SH  group  and  the  NADH  pyridine 
ring  is  broken  by  the  mercurial.  However,  some  doubts  have  recently  been 
cast  on  this  simple  interpretation.  The  liver  alcohol  dehydrogenase  mol- 
ecule has  28  SH  groups  as  determined  by  p-MB  titration;  the  presence 
of  NADH  does  not  reduce  this  number,  although  NADH  protects  the  en- 
zyme moderately  (Witter,  1960).  On  a  rather  tenuous  basis,  Witter  postulat- 
ed that  the  function  of  the  SH  groups  is  to  maintain  the  stable  enzyme 


INHIBITION  OF  ENZYMES  785 

structure  rather  than  bind  the  NADH;  disintegration  of  the  structure 
brought  about  by  p-MB  would  secondarily  lead  to  release  of  the  coenzyme. 
As  the  NADH  is  split  from  the  apoenzyme  by  p-MB,  rotatory  dispersion 
titration  indicates  changes  in  optical  rotation  associated  with  denaturation, 
so  that  Li  et  al.  (1962)  likewise  inclined  to  a  theory  involving  structural 
changes  as  a  basis  for  the  displacement,  since  it  is  known  that  denaturation 
by  heat  or  other  agents  releases  the  NADH.  Yonetani  and  Theorell  (1962) 
have  used  the  very  sensitive  spectrofluorometric  method  for  the  measure- 
ment of  NADH  binding  and  dissociation.*  They  demonstrated  that  the  en- 
zyme configuration  is  stabilized  by  the  NADH  and  additionally  by  the  iso- 
butyramide,  although  these  can  be  associated  directly  with  only  a  small 
fraction  of  the  total  number  of  SH  groups,  and  suggested  that  denaturation 
may  be  Initiated  by  local  changes  at  the  active  centers  and  from  there 
spread  throughout  the  molecule.  The  SH  groups  may  form  a  network  of 
hydrogen  bonds  contributing  to  the  stability  of  the  tertiary  structure,  so 
that  mercurials  could  create  instability  either  locally  or  generally.  All  of 
this  recent  work  shows  that  mercurials  probably  induce  configurational 
changes  in  the  enzyme,  these  being  irreversible  by  the  usual  means,  and 
they  provide  an  alternative  explanation  for  NADH  release,  but  do  not  dis- 
prove the  original  hypothesis  that  a  direct  binding  between  NADH  and 
SH  groups  occurs.  The  Zn++-dependent  alcohol  dehydrogenase  of  yeast  is 
inhibited  by  mercurials,  and  this  was  attributed  to  displacement  of  the 
Zn++  from  SH  groups  (Wallenfels  and  Sund,  1957  a)  on  the  basis  that  res- 
toration of  activity  requires  both  glutathione  and  Zn++.  However,  an  inves- 
tigation of  the  time  course  of  the  inhibition  showed  that  glutathione  alone 
is  sufficient  to  reactivate  if  it  is  added  soon  after  the  mercurial,  but  the  inhi- 
bition progressively  becomes  irreversible,  at  which  time  no  Zn++  has  been 
released  (Snodgrass  and  Hoch,  1959).  Zn++  is  displaced  progressively  over 
a  period  of  several  hours,  but  the  inhibition  does  not  appear  to  be  mediat- 
ed through  this  displacement.  Such  a  slow  release,  without  correlation  with 
the  mercurial  reaction  or  the  inhibition,  is  probably  due  to  structural 
changes  in  the  enzyme. 

The  early  work  on  NADH  splitting  from  alcohol  dehydrogenase  was  soon 
confirmed  for  3-phosphoglyceraldehyde  dehydrogenase  (3-PGDH)  by  Velick 
(1953).  The  yeast  enzyme  requires  2  SH  groups  for  full  activity  and  inhi- 
bition by  p-MPS  increases  until  2  equivalents  of  the  mercurial  are  added. 
In  contrast  to  the  alcohol  dehydrogenase,  this  inhibition  is  readily  revers- 

*  Liver  alcohol  dehydrogenase  forms  a  very  stable  ternary  complex  with  NADH 
and  t«obutyramide,  and  this  complex  is  strongly  fluorescent.  The  enzyme  may  be 
titrated  in  the  presence  of  100  mM  isobutyramide  with  NADH,  measuring  the  fluo- 
rescence increase  at  410  m/<,  and  then  back-titrated  with  p-MB  or  p-MPS  as  the  NADH 
is  released  from  the  apoenzyme.  This  will  probably  be  a  very  valuable  technique  for 
the  study  of  coenzyme  binding. 


786 


7.  MERCURIALS 


ible  with  cysteine.  Muscle  3-PGDH  behaves  similarly  but,  after  charcoal 
treatment,  reacts  with  3  equivalents  of  p-MPS.  If  the  enzyme  is  ultracen- 
trifuged  it  carries  down  most  of  the  NAD  but  the  inhibited  enzyme  does 
not,  most  of  the  NAD  being  free  in  the  medium.  The  correlations  between 
these  various  events  are  shown  in  Fig.  7-12.  The  change  in  spectral  absorp- 
tion at  340  ji/m  associated  with  NAD  binding  disappears  progressively  with 
added  p-MPS,  again  indicating  coenzyme  release.  Velick  was  careful  not  to 
assume  that  these  results  necessarily  point  to  a  binding  of  the  NAD  by 
the  SH  groups,  but  stated  that  the  p-MPS  can  sterically  or  electrostatically 


1.5 


p-MPS  BOUND 


0.5     , 


"0  I  2 

EOUIVS    p-MPS    ADDED    


Fig.  7-12.  Dissociation  of  NAD  from  3-phosphogly- 
ceraldehyde  dehydrogenase  by .  p-MPS,  and  the  si- 
multaneous   loss    of  activity.    (From   Velick,    1953.) 


interfere  with  the  coenzyme  binding  if  the  SH  groups  are  close  enough  to 
the  active  center.  Fluorometric  titration  of  the  3-PGDH- (NADH)3  complex 
from  rabbit  muscle  with  p-MB  demonstrates  NADH  dissociation,  but  the 
kinetics  indicate  that  each  p-MB  bound  weakens  the  coenzyme  binding  at 
the  other  sites,  so  that  perhaps  structural  changes  in  the  protein  occur 
(Velick,  1958).  The  same  number  of  equivalents  of  p-MB  is  required  to 
release  the  NADH  from  3-PGDH- (NADH )i,  3-PGDH- (NADH)2,  and  3- 
PGDH-(NADH)3.  If  the  displacement  were  due  to  direct  competition  with 
the  NADH  for  SH  groups,  one  would  expect  the  p-MB  to  attack  the  unoc- 
cupied SH  groups  first,  with  no  release  of  NADH,  but  this  is  not  the  case, 
the  release  beginning  immediately,  as  if  the  reaction  of  any  SH  group  al- 
tered the  enzyme  structure  throughout.  These  structural  changes,  if  they 
occur,  must  be  readily  reversible. 

Results  with  a  few  other  enzymes  will  be  discussed  briefly.  Reports  on 
coenzyme  displacement  from  lactate  dehydrogenase  have  not  been  entirely 
consistent.  Apparently  the  state  of  the  enzyme  and  particularly  the  source 
are  important  factors.  Kaplan  and  Ciotti  (1954)  found  that  p-MB  releases 
NAD  from  the  liver  enzyme,  this  being  associated  with  a  fall  in  absorption 
at  300  m//,  but  Chance  (1954)  could  detect  no  release  from  heart  lactate 


INHIBITION  OF  ENZYMES  787 

dehydrogenase  at  5°,  an  observation  confirmed  by  Velick  (1958).  However, 
Winer  et  al.  (1959)  find  a  slow  dissociation  of  NADH  from  heart  lactate 
dehydrogenase,  complete  release  occurring  after  1  hr  at  26°  and  pH  7  with 
0.135  vciM  2?-MB.  The  L(+)-lactate  dehydrogenase  of  yeast  (cytochrome 
bj)  possesses  a  flavin  prosthetic  group  and  this  is  readily  dissociated  by 
p-MPS  (Armstrong  et  al,  1960,  1963).  The  binding  of  NADH  and  NADPH 
to  cytochrome  65  aporeductase  is  blocked  by  p-MB  (Strittmatter,  1961  b) 
and  there  is  some  evidence  that  pyridoxal-P  may  be  split  from  L-threonine 
deliydrase  by  the  same  mercurial  (Nishimura  and  Greenberg,  1961).  The 
evidence  for  the  displacement  of  Fe++  from  homogentisate  oxidase  by  p-MB 
has  already  been  discussed,  and  the  reactions  of  inhibition  and  reactivation 
(Crandall,  1955)  may  be  written: 

Inhibition:        E— S— Fe+  +  R— Hg— X  ->  E— S— Hg— R  +  Fe++  +  X- 

E— S— Hg— R  +  GSH         ->  E— SH  +  GS— Hg— R 
Reactivation:  „     „     „ 

E— SH  +  Fe++  ->  E— S— Fe+  +  H+ 

The  nonheme  Fe  of  succinate  dehydrogenase  is  lost  more  rapidly  by  dialysis 
after  treatment  with  p-MPS,  and  this  may  be  related  to  the  marked  spectral 
changes  observed  upon  reaction  with  the  mercurial  (Massey,  1958).  The 
iron  of  the  photosynthetic  pyridine  nucleotide  reductase  is  released  as  Fe+++ 
by  p-MB  with  proportional  loss  of  activity  (Katoh  and  Takamiya,  1963). 
The  inhibition  of  aminopeptidase  by  EDTA  is  made  irreversible  by  simul- 
taneous treatment  with  p-MB  and  it  was  concluded  that  the  Mn++  is  bound 
to  an  SH  group  (Bryce  and  Rabin,  1964).  A  final  type  of  experiment  will 
be  mentioned.  Mn++  activates  the  hydroxylamine  reductase  of  P.  aeruginosa 
and  this  activation  is  prevented  by  p-MB,  "suggesting  that  SH  groups  may 
be  involved  in  binding  the  metal  to  the  enzyme"  (Walker  and  Nicholas, 
1961).  It  seems  to  me  that  such  conclusions  are  unjustified,  inasmuch  as 
any  mechanism  of  inhibition  would  presumably  abolish  activation  by  Mn++, 
whether  it  affected  the  binding  or  not. 

The  results  on  coenzyme  displacement  may  be  summarized  by  stating 
that  the  same  difficulties  are  encountered  as  in  protection  experiments. 
There  are  three  general  mechanisms  by  which  a  mercurial  could  dissociate 
an  enzyme-coenzyme  complex:  (1)  compete  with  the  coenzyme  for  the  SH 
group,  (2)  sterically  or  electrostatically  interfere  with  coenzyme  binding  by 
reacting  at  an  adjacent  site,  and  (3)  alter  the  enzyme  configuration  in  such 
a  way  as  to  disrupt  secondarily  the  coenzyme  binding.  In  no  case  have  these 
mechanisms  been  distinguished. 

Changes  in  Enzyme  Structure  Brought  About  by  Mercurials 

Evidence  has  accumulated  during  the  past  several  years  that  mercurials 
occasionally  initiate  configurational  changes  in  enzymes;  certain  aspects  of 


788  7.  MERCURIALS 

this  have  been  discussed  in  the  previous  section  and  we  shall  now  inquire 
what  further  evidence  on  this  important  problem  has  come  to  light.  One 
of  the  more  obvious  reasons  for  suspecting  denaturation  is  the  progressive 
development  of  irreversibility  during  contact  with  the  mercurial,  such  as 
has  been  reported  for  cholinesterase  (Goldstein  and  Doherty,  1951),  pros- 
tatic phosphomonoesterase  (Tsuboi  and  Hudson,  1955  a),  muscle  aldolase 
(Swenson  and  Boyer,  1957).  and  muscle  3-PGDH  (Elodi,  1960)  —  to  men- 
tion only  a  few  instances,  in  most  cases  reversal  being  attempted  with  glu- 
tathione or  dimercaprol.  Of  course,  one  might  attribute  failure  to  reverse 
to  very  tight  binding  to  the  enzyme,  but  the  progressive  increase  in  the 
irreversibility  points  more  to  structural  changes.  The  question  often  remains 
as  to  whether  these  changes  are  responsible  for  the  inhibition  or  are  super- 
imposed upon  it,  i.e.,  inhibition  followed  by  inactivation. 

Elodi  (1960)  investigated  the  changes  in  several  properties  of  pig  muscle 
3-PGDH  treated  with  p-MB,  and  found  significant  deviations  in  the  op- 
tical rotation  and  the  intrinsic  viscosity,  the  latter  increasing  linearly  with 
the  equivalents  of  mercurial  added.  The  following  phases  were  postulated: 
(1)  an  initial  reversible  binding  and  inhibition,  (2)  a  progressive  disintegra- 
tion of  the  secondary  structure  of  the  enzyme  as  a  result  of  the  blocking  of 
SH  groups,  this  probably  involving  an  unfolding  of  the  polypeptide  helices, 
and  (3)  polymerization  and  precipitation  consequent  to  the  freeing  of 
groups  which  form  intermolecular  bridges.  The  simultaneous  changes  in  the 
activity,  NADH  binding,  and  rotatory  dispersion  of  yeast  alcohol  dehydro- 
genase treated  with  p-MB  led  Wallenfels  and  Miiller-Hill  (1964)  to  postulate 
that  modifications  of  the  secondary  and  tertiary  protein  structure  occur 
when  the  SH  groups  are  blocked.  Reaction  of  10  SH  groups  on  muscle 
aldolase  with  p-MB  does  not  reduce  the  activity  but  the  susceptibility  to 
tryptic  digestion  is  increased  (Szabolcsi  and  Biszku,  1961).  Untreated  al- 
dolase or  treated  enzyme  in  the  presence  of  fructose-diP  is  not  digested  by 
trypsin;  thus  the  substrate  apparently  protects  the  active  center,  and  per- 
haps the  entire  molecule,  from  hydrolysis.  It  is  thought  that  reaction  of  the 
first  7  free  SH  groups  labilizes  the  tertiary  structure  of  the  enzyme,  and 
from  then  on  a  progressive  denaturation  occurs.  Addition  of  substrate  may 
restore  to  some  degree  the  normal  structure.  The  inhibition  that  occurs 
later  or  with  excess  p-MB  does  not  seem  to  be  directly  related  to  mercaptide 
formation  but  dependent  on  the  structural  changes  when  they  have  pro- 
ceeded past  a  certain  point.  Another  interesting  approach  was  made  by 
Massey  (1958)  in  showing  that  the  chelation  of  the  nonheme  iron  of  suc- 
cinate dehydrogenase  by  o-phenanthroline  is  accelerated  by  treatment  with 
p-MPS,  this  being  interpreted  as  a  structural  change  exposing  the  iron. 
The  SH  groups  of  yeast  hexokinase  can  be  titrated  with  p-MB  in  the  pre- 
sence of  glucose  without  loss  of  activity,  but  spontaneous  denaturation 
quickly  follows  (Fasella  and  Hammes,  1963).  Glucose-6-P  does  not  prevent 


INHIBITION  OF  ENZYMES  789 

loss  of  activity  during  the  titration.  These  results  indicate  that  the  SH 
groups  are  not  directly  involved  in  the  catalysis,  but  function  to  stabilize 
the  enzyme  in  the  active  configuration. 

Another  type  of  structural  change  is  depolymerization  of  the  enzynae 
into  subunits  following  mercurialization.  Muscle  phosphorylase  is  progres- 
sively inhibited  by  p-MB  until  around  18  equivalents  of  the  mercurial  are 
combined,  and  this  is  accompanied  by  the  appearance  of  a  new  molecular 
species  in  the  ultracentrifuge,  the  sedimentation  constant  being  lower  than 
that  for  either  phosphorylase  a  or  b  (Madsen  and  Cori,  1955): 

Phosphorylase  a:  S  =  13.2 

Phosphorylase  b:  S  =    8.2 

Inactive  enzyme:  S  =    5.6 

Both  phosphorylase  a  and  b  form  this  new  species  with  p-MB  and  it  was 
suggested  that  the  former  is  split  into  4  subunits,  the  latter  into  2  subunits. 
Light  scattering  studies  are  consistent  with  this  interpretation  (Madsen, 
1956).  The  inhibition  develops  more  rapidly  than  the  depolymerization, 
however,  so  the  relationship  between  them  is  not  clear.  Removal  of  the 
p-MB  with  cysteine  restores  both  activity  and  the  normal  dimer  or  tetra- 
mer  (Madsen  and  Cori,  1956).  The  extent  of  the  conversion  of  the  phos- 
phorylase tetramer  to  the  monomer  is  proportional  to  the  number  of  SH 
groups  reacted  and  an  all-or-none  dissociation  of  the  units  is  likely  (Madsen 
and  Gurd,  1956).  The  sedimentation  constant  of  yeast  alcohol  dehydroge- 
nase is  reduced  from  7.2  to  3.3  by  p-MB,  this  being  secondary  to  the  inhi- 
bition of  the  enzyme,  so  that  here  dissociation  into  subunits  apparently 
occurs  (Snodgrass  et  al.,  1960).  Reaction  of  myosin  ATPase  with  MM  also 
causes  the  appearance  of  a  small  subunit,  but  this  is  not  related  to  the 
binding  to  the  SH  groups  responsible  for  the  activity  (Kominz,  1961).  In 
addition  there  is  some  aggregation  to  a  faster  sedimenting  species  and  this 
is  perhaps  correlated  with  reaction  of  SH  groups  at  the  active  center.  The 
inhibition  of  rabbit  muscle  enolase  by  p-MB  was  considered  to  be  secondary 
to  denaturation  and  not  directly  due  to  SH  group  reaction,  on  the  basis  of 
the  variation  of  activity  with  the  equivalents  of  mercurial  present  and  the 
appearance  of  turbidity  (Malmstrom,  1962).  The  sedimentation  constant  of 
liver  glutamate  dehydrogenase  is  reduced  by  MM  and  PM  and  again  the 
most  likely  explanation  is  a  splitting  into  subunits  (Rogers  et  al.,  1962, 
1963;  Greville  and  Mildvan,  1962).  The  relationship  of  the  disaggregation 
to  the  unique  changes  in  enzyme  activity  is  not  clear. 

There  is  no  doubt  that  mercurials  can  induce  structural  changes  in  cer- 
tain enzymes,  and  cause  aggregation  or  fractionation  into  subunits  in  others, 
but  the  significance  for  primary  inhibition  has  not  been  clarified.  Is  the 
inhibition  due  to  the  blocking  of  functional  SH  groups  or  secondarily  to 
the  structural  changes?  Does  the  denaturation  result  from  general  SH  group 


790  7.  MERCURIALS 

reaction  or  can  it  originate  solely  by  mercaptide  formation  at  the  active 
center?  The  results  taken  all  in  all  tend  to  signify  that  inhibition  usually 
occurs  upon  the  initial  reaction  of  the  SH  groups  at  or  near  the  active 
center,  this  being  reversible,  and  that  slower  structural  alterations  proceed 
as  a  result  of  either  the  mercurial  already  combined  or  the  continued  reac- 
tion with  more  mercurial  (perhaps  with  the  less  available  SH  groups),  these 
changes  becoming  more  and  more  irreversible,  a  progressive  inactivation 
being  superimposed  on  the  primary  inhibition.  Lability  may  also  come 
about  by  a  displacement  of  coenzyme  or  cofactors,  since  these  undoubtedly 
help  maintain  configurational  integrity,  especially  in  the  abnormal  state  in 
which  isolated  enzymes  find  themselves.  This  does  not  imply  that  all  en- 
zymes behave  in  this  fashion;  it  is  quite  possible  that  in  some  the  structural 
changes  may  be  primary  and  the  sole  cause  of  the  inhibition.  Certain  en- 
zymes suffer  only  the  primary  inhibition  and  the  stability  is  not  reduced 
by  the  mercurial,  and  indeed  stability  may  be  increased,  as  we  have  ob- 
served with  papain  and  pinguinain.  The  requirement  to  solve  the  problems 
of  the  relation  between  inhibition  and  inactivation,  and  between  both  proc- 
esses and  the  types  of  SH  group  reacted,  is  for  more  detailed  studies  cor- 
relating the  time  courses  of  as  many  of  these  changes  as  possible  as  they 
occur  after  introduction  of  the  mercurial.  Another  approach  might  be  to 
do  occasional  experiments  at  low  temperatures,  where  inactivation  or  de- 
naturation  would  occur  very  slowly,  in  this  way  possibly  separating  the 
primary  inhibition  from  these  other  changes.  Finally,  it  might  be  suggested 
that  every  effort  to  create  conditions  favoring  stability  of  the  enzymes  be 
made.  One  gets  the  impression  that  often  so  little  attention  is  paid  to  the 
proper  pH,  ion  concentrations,  buffers,  and  other  factors,  that  the  enzyme 
as  studied  is  in  a  relatively  unstable  state  and  hypersusceptible  to  any  in- 
hibitors subjecting  the  normal  protein  configuration  to  even  minor  stress. 

Effects  of  pH,  Ions,  and  Buffers  on  Mercurial  Inhibition 

The  effects  of  pH  on  OH""  complexes  with  mercurials  (page  736),  on 
mercaptide  formation  (page  749),  and  on  reactions  of  proteins  with  mer- 
curials (page  760)  have  been  discussed.  The  results  and  the  factors  which 
may  be  involved  can  be  summarized  as  follows:  (1)  pH  affects  the  ioniza- 
tion of  the  SH  groups  or  the  competition  between  mercurial  and  H+  for 
the  S~  group,  (2)  pH  alters  the  concentration  of  0H~  and  hence  the  amount 
of  mercurial  complexed  with  this  anion,  (3)  pH  influences  the  protein 
charge  possibly  attracting  or  repelling  charged  mercurials,  (4)  pH  deter- 
mines the  rate  of  secondary  inactivation  or  denaturation,  (5)  pH  affects 
the  aggregation  state  of  protein-mercurial  complexes  (e.g.,  the  degree  of 
dimerization  of  mercaptalbumin  complexes  with  Hg++),  (6)  pH  affects  both 
the  rate  and  the  number  of  SH  groups  reacted,  and  (7)  generally  there  is 
an  increased  rate  of  protein  reaction  with  mercurials  as  the  pH  is  reduced. 


INHIBITION  OF  ENZYMES 


791 


The  effects  of  pH  on  enzyme  inhibition  by  the  mercurials  are  even  more 
complex  and  one  would  not  anticipate  consistent  behavior,  a  prediction 
that  is  borne  out  in  the  following  discussion. 

Some  pH  effects  on  mercurial  inhibition  are  shown  in  Table  7-10.  In  7 
cases  the  inhibition  is  greatest  at  low  pH  and  in  6  cases  at  high  pH,  and 
this  certainly  indicates  that  more  than  one  factor  must  be  involved.  Indeed, 
many  of  the  inhibition-pH  curves  are  complex  (Figs.  7-13  and  7-14)  and 


80 


60 


20 


Fig.  7-13.  Effects  of  pH  on   the  inhibitions   of  /J-fruc- 

tofuranosidase    by    HgClj   and    Hg(N03)2.    (Data    from 

Myrback,   1926.) 


often  biphasic  (/^-glucuronidase,  /J-fructofuranosidase,  ATPase,  and  ribo- 
nuclease).  In  some  cases  marked  stimulation  is  found  within  a  certain  pH 
range,  inhibition  occurring  outside  this  range.  This  implies  that  the  activity- 
pH  curves  and  the  pH^p^  are  shifted  by  the  mercurials,  usually  to  lower 
pH's,  as  for  yeast  proteinase  (Lenney,  1956)  and  ascites  cell  ribonuclease 
(EUem  and  Colter,  1961;  Colter  et  al.,  1961).  It  is  difficult  to  interpret  such 
shifts  in  pH^p^,  but  if  the  pH^p^  is  related  to  the  ionization  of  two  or  more 
groups  on  enzyme  and  substrate  (see  page  1-660),  a  shift  implies  some  mod- 
ification of  the  enzyme  groups  in  or  near  the  active  center  with  a  resultant 
alteration  of  the  interaction  of  the  substrate  with  the  enzyme.  Dixon's 
method  of  plotting  K,„  and  K^  against  pH  (page  1-683)  was  applied  to  ^- 
glucuronidase  by  Fernley  (1962),  and  the  curves  are  shown  in  Fig.  7-15. 
The  pK^  curve  is  suggestive  of  two  ionizing  groups  on  the  enzyme  with 
pi^L^'s  around  4.4  and  6.3  if  it  could  be  simply  interpreted,  or  possibly  the 


792 


7.  MERCURIALS 


P3 


O    (O    'O 

eo  CO  •* 


+  + 


ec  o 


■>*    CD 


00    tJ<    <© 

(M  eo  Tj< 


O 


^ 

a 

■^ 

?= 

y 

ft 

■^ 

M 

o 

S 

o 

M 


CO  00  :d  lo 

^  -H  o  e<i  -^ 

o  o  o  o  o 

o  o  o  o  o 

o  o  o  o  o 

<z>  c:>  <6  <£  S 


m 


« 


pq 


a 


w 


pL, 


eS 

ft 


a 

c 


M 


INHIBITION  OF  ENZYMES  793 


O 


fe 


t^OO        COi— lOi^OSOSCJOOOM         Co         OO         OOOO        05CC        OiOt^TjtO 
iO<£>        -^tM-^OOOOiCTtH-^iO        J310         Ouo         100»00        00  lO^GOO 


+ 


ICIO        »00>00»COiOOO        OiiO         Tt<o         O  •^  OO        fOOOO 

V 


"2   2  Tt<       ^ 

S,  Ocq         OOOO 

^l'  do      dodo      ^ 
d 


bli  g  S  oj  1^  bo  ^ 

W  Ph  PL,  S  a  ffi  a 


P5  M  Ph  ^  O 


g  ^  ?    ffi  o    s  .s        .s 

^  ^  ?    o  :2    .Si  5-2 


794 


7.  MERCURIALS 


A 


ft 


o 

^ 

■* 

1 

e 

OS 

1— 1 

"S 

^ 

OQ 

'o 

cS 

O 

QQ 

Q 

e 

_o 

'-P 

VD 

<M 

IC 

O 

cq 

cn 

05    CD    lO    t^ 

^ 

n 

CO 

M 

lO 

Ti< 

t^ 

-H   ■<*   i>   00 

(N 

>o 

2 

+  +  +  + 

o 


S 


o 


o  o  o  o  >o  o 

ic  CO  eo  -H 

CO 

C2 

so 
o  o 

00 

>0     CO     t-     GO     00    05 

■<#  lo  CO  t^ 

Tt* 

lO 

t^    00 

05 

m 


Oh 


w 

»o 

ft 

K 

ft 

d 

a 

-1-5 

o3 

'S 

<A 

M 

+ 

o 

s 

>> 

^ 

CO 

o3 

-t^ 

CO 

cS 

'^ 

o 

X 

-2 

'S 

ft 

cS 

fl 

-O 

*^ 

-1-2 

s 

P5 


INHIBITION  OF  ENZYMES 


795 


group  with  p^a  near  4.4  could  be  the  benzoate  group  of  p-MB,  particularly 
as  no  inflection  is  shown  here  in  the  K„,  curve.  However,  as  Fernley  clearly 
pointed  out,  one  must  proceed  with  caution  because  of  the  many  complexes 
possible  with  the  mercurial,  and  particularly  the  buffer  effects.  It  seems 
unlikely  that  the  decrease  in  K,  above  pH  5  is  due  to  complexing  with  OR- 
ions  entirely,  but  it  might  be  the  result  of  reactions  with  the  buffer;  the  sim- 
ilar inflection  for  K„^  makes  it  reasonable  that  an  enzyme  group  is  involved. 
The  results  with  PM  are  quite  different  than  with  Hg++  or  p-MB  (Fig.  7-14); 

100 


80 


60 


40 


20 


-20 


Hg=0.05  mM 

Ag^O  1  mM 

Hg** 

Cui=  1  mM 

PM<0  4  mM 

/           /*'* 

K 

^  PM 

r" 

■ 

y^ 

Fig.  7-14.  Effects  of  pH  on  the  inhibitions  of  ^-glu- 
curonidase by  heavy  metal  ions.  (From  Fernley,  1962.) 


PM  is  generally  less  inhibitory  than  p-MB,  does  not  increase  Kf„  whereas 
p-MB  does,  facilitates  the  formation  of  the  ESg  complex  (favoring  substrate 
inhibition)  whereas  p-MB  does  not,  and  increases  the  pH^pt  whereas  Hg++ 
and  p-MB  decrease  it.  The  marked  effect  of  PM  on  substrate  inhibition  is 
shown  in  Fig.  7-16  but  no  explanation  is  available.  There  is  certainly  a  need 
for  more  accurate  comparisons  of  the  different  mercurials,  not  only  with 
respect  to  pH  effects  but  generally.  The  different  curves  obtained  with  Hg++ 
and  PM  acting  on  pancreatic  amylase  (Fig.  7-17)  are  intriguing  and  one 
feels  that  an  explanation  of  such  phenomena  might  well  aid  in  our  under- 


796 


7.  MERCURIALS 


3- 


■ 

■— -~^' 

^^^ 

•-^iljm 

' 

^~~^^.,^WG  V^                                  N. 

Fig.  7-15.  Effects  of  pH  on  the  ^„  and   V„  for  ^-glucuronidase, 
and  the  K,  for  the  inhibition  by  p-MB.  (From  Fernley,  1962.) 

standing  of  the  mechanisms  of  enzyme  inhibition.  The  pH  changes  may  af- 
fect the  reactivities  of  different  SH  groups  on  an  enzyme  to  different  de- 
grees. Xanthine  oxidase  has  two  types  of  SH  group,  one  reacting  relatively 


100 


80 


40 


20 


CU     (0  63mM| 
PM(025mM) 


Fig.  7-16.  Effects  of  substrate  concentration 

on    the    inhibition    of    /3-glucuronidase    by 

p-MB,    PM,   and   Cu++   at   pH    5.9.    (From 

Fernley,   1962.) 


INHIBITION  OF  ENZYMES 


797 


rapidly  and  unrelated  to  the  enzyme  activity  and  the  other  reacting  slowly 
to  produce  inactivation;  altering  the  pH  from  5.2  to  7.2  modifies  the  rates 
of  reaction  of  these  groups  with  p-MB  in  quite  different  ways  (Gilbert, 
1963). 

Increase  of  any  ligand  capable  of  complexing  the  mercurials  should  re- 
duce the  inhibition,  but  this  has  been  studied  very  little.  Fernley  (1962) 
noted  that  raising  the  CI"  concentration  suppresses  the  inhibition  of  /?- 
glucuronidase  by  Hg++.  It  is  very  difficult  in  such  cases  to  separate  a  direct 
complexing  action  from  an  ionic  strength  effect.  Green  and  Neurath  (1953) 


-6  -5  -4 

LOG    CONCENTRATION    (mM) 


Fig.  7-17.  Concentration-inhibition  curves  for  the  actions  of  Hg++ 
and  PM  on  pancreatic  amylase.  (From   Owens,  1953  b.) 


varied  the  ionic  strength  with  NaCl,  SrClg,  and  (NH4)2S04  and  found  the 
inhibition  of  trypsin  by  Hg++  to  be  suppressed  at  high  ionic  strengths 
(Fig.  7-18).  Since  all  the  salts  had  essentially  the  same  effect,  they  assumed 
that  this  is  not  due  to  specific  ions,  but  certainly  part  of  the  reduction 
must  be  due  to  increasing  formation  of  the  Hg++  complexes  with  CI",  NH4+, 
and  864=.  The  nature  of  the  inhibition  of  trypsin  is  not  clear  since  free  SH 
groups  are  generally  not  considered  to  play  a  role  in  the  active  center. 
The  inhibition  of  pancreatic  amylase  by  Hg++  and  PM  was  postulated  by 
Owens  (1953  a)  to  involve  phosphate  ion,  and  the  unusual  configuration  of 
the  PM  inhibition  curve  was  attributed,  at  least  in  part,  to  the  formation 
of  phosphate  complexes,  although  it  is  strange  that  Hg++  is  not  similarly 
affected.  It  is  also  not  clear  to  me  why  the  curve  should  assume  this  shape, 
particularly  why  increase  in  PM  from  0.1  to  1  vaM  should  bring  about  a 


798 


7.  MERCURIALS 


lessening  of  the  inhibition.  The  inhibition  of  NADH  dehydrogenase  by  p- 
MPS  is  weaker  in  phosphate  than  in  THAM  buffer,  and  this  could  again  be 
due  to  complexes  with  phosphate  (Minakami  et  al.,  1963).  Despite  the  pau- 
city of  experimental  information,  the  effects  of  ions  and  buffers  should  be 
more  frequently  taken  into  account  in  the  use  of  the  mercurials.  In  addition 
to  the  ionic  effects  just  described,  certain  specific  actions  have  been  noted, 


100 


50 


25 


X 
ACT 


rONIC    STRENGTH   ON 
HgClj    INHIBITION 
(HgClj).0  05mM 


'0  0.01 

IONIC     STRENGTH 


0.02 


0.03 


0.04 


005 


Fig.  7-18.  Inhibition  of  trypsin  by  Hg++  and  the  ef- 
fect of  ionic  strength.  (From  Green  and  Neurath,  1953.) 

especially  with  ATPase  (Novikoff  et  al.,  1952;  Lardy  and  Wellman,  1953; 
Sacktor  et  al,  1953),  where  the  concentrations  of  Mg++  and  Ca++  determine 
to  some  extent  the  susceptibility  of  the  mitochondrial  enzyme  to  mercurials, 
generally  Mg++  lessening  the  inhibition,  although  here  one  may  be  dealing 
with  more  than  one  enzyme. 


Titration  of  Enzyme  SH  Groups 

The  primary  purpose  of  a  titration  with  a  mercurial  is  to  determine  the 
number  of  reactive  SH  groups  which  an  enzyme  possesses,  either  normally 
or  after  being  treated  in  various  ways  (e.g.,  denatured  with  urea,  guanidine, 


INHIBITION  OF  ENZYMES  799 

or  high  temperature),  and  in  this  respect  the  problem  is  no  different  from 
that  of  protein  titration  (page  762).  We  are  interested  here  not  so  much 
in  the  number  of  SH  groups  on  an  enzyme,  but  how  these  SH  groups  relate 
to  the  catalytic  activity  and  the  mechanisms  of  mercurial  inhibition,  and 
thus  in  following  simultaneously  the  loss  of  SH  groups  and  the  develop- 
ment of  the  inhibition  as  more  and  more  mercurial  is  added.  Let  us  assume 
that  we  have  a  solution  of  a  pure  enzyme  and  we  add  to  this  a  solution 
of  p-MB,  or  other  mercurial,  so  that  the  molar  ratio  of  mercurial  to  enzjnne 
is  slowly  increased,  and  further  assume  that  we  allow  time  for  the  reaction 
to  come  to  equilibrium.  We  measure  the  number  of  SH  groups  reacted 
(spectrophotometrically,  polarographically,  argentimetrically,  or  otherwise) 
and  the  enzyme  activity.  There  are  only  a  few  fundamental  relationships 
between  mercaptidization  and  inhibition  that  could  emerge,  and  these  are 
illustrated  in  Fig.  7-19.  These  are  extreme  situations,  of  course,  and  inter- 
mediate behavior  would  more  often  be  expected.  Let  us  consider  what  each 
result  may  mean  and  how  valid  certain  interpretations  may  be. 

Case  A:  The  inhibition  runs  parallel  to  the  SH  groups  reacted. 

(1 )  The  only  SH  groups  that  react  are  at  the  active  center  and  mercaptide 
formation  abolishes  the  enzyme  activity;  titration  to  100%  inhibition  will 
give  the  number  of  SH  groups  at  the  active  center. 

(2)  There  are  n  equireactive  SH  groups  on  the  enzyme,  but  only  a  cer- 
tain fraction  of  these  is  at  the  active  center  or  involved  in  the  catalysis; 
the  titration  will  not  provide  the  actual  number  at  the  active  center. 

(3)  The  SH  group  or  groups  are  not  at  the  active  center,  but  reaction 
of  them  leads  to  inactivation  of  the  enzyme  by  some  means;  titration  will 
not  provide  useful  information. 

It  is  impossible  to  distinguish  between  these  possibilities  by  simple  ti- 
tration nor  can  one  determine  accurately  the  number  of  SH  groups  at  the 
active  center.  Let  us  assume  that  an  enzyme  has  10  reactive  SH  groups 
totally  but  that  complete  inhibition  occurs  when  only  3  are  reacted.  It  has 
sometimes  been  concluded  that  3  SH  groups  are  necessary  for  the  enzyme 
activity.  This  is  not  a  valid  conclusion.  If  1  of  the  3  groups  were  related  in 
some  way  to  the  activity,  one  would  obtain  the  same  data.  In  other  words, 
the  number  of  equivalents  of  mercurial  added,  or  the  number  of  SH  groups 
reacted,  to  achieve  100%  inhibition  does  not  provide  directly  the  number 
of  SH  groups  involved  in  the  catalysis. 

Case  B:  SH  groups  are  reacted  but  inhibition  does  not  occur. 

The  conclusion  here  is  obvious:  the  reactive  SH  groups  are  not  involved, 
directly  or  indirectly,  in  the  enzyme  activity.  It  is,  of  course,  possible  that 
there  are  SH  groups  at  the  active  center,  perhaps  even  functional,  but  that 
they  do  not  react  with  the  mercurial  under  the  experimental  conditions. 


800 


7.  MERCURIALS 


Case  C:  Inhibition  develops  only  after  so  many  SH  groups  have  reacted. 

(1)  The  most  reactive  SH  groups  are  not  related  to  the  activity  of  the 
enzyme,  but  the  less  readily  available  ones  are. 

(2)  None  of  the  SH  groups  is  directly  related  to  the  activity,  but  when 
a  sufficient  number  are  reacted  the  enzyme  is  structurally  altered  so  that 
inhibition  appears. 

These  situations  may  be  distinguished  sometimes  by  determining  the 
reversibility  of  the  inhibition,  but  occasionally  denaturation  is  reversible. 

Case  D:  Inhibition  develops  completely  before  all  the  SH  groups  react. 
(!)  The  most  reactive  SH  groups  are  at  the  active  center  and  when  they 


P-MB    m^ 

D 

1 

/ 

/         / 

/     / 

/    / 

/  / 

/  / 

i' 

F 

/                            / 

/                              / 

'                              / 

/ 

/ 

/ 

/ 

/ 

/ 

/ 

/ 

Fig.  7-19.  Illustrations  of  the  various  relation- 
ships between  the  reaction  of  enzyme  SH 
groups  with  mercurials  and  the  inhibition  of 
enzyme  activity.  These  are  all  extreme  cases 
and  intermediate  situations  can  also  occur. 
p-MB  has  been  designated  but  any  mercurial 
may  be  used.  The  solid  lines  show  the  inhi- 
bition and  the  dashed  lines  the  reaction  with 
SH  groups;  i  is  the  fractional  inhibition  and  r 
is  the  fraction  of  the  total  SH  groups  reacted. 


INHIBITION  OF  ENZYMES  801 

are  all  combined  with  the  mercurial  the  enzyme  is  completely  inhibited, 
the  less  readily  reacting  SH  groups  being  now  available. 

(2)  Only  a  small  number  of  SH  groups  are  normally  reactive,  but  follow- 
ing mercaptide  formation  the  enzyme  unfolds  and  exposes  other  SH  groups, 
the  groups  reacted  after  total  inhibition  being  secondarily  released. 

As  discussed  under  Case  A,  the  titration  to  100%  inhibition  does  not 
necessarily  provide  the  number  of  SH  groups  involved  in  the  catalysis. 
Reversal  experiments  run  in  parallel  to  the  titration  may  give  some  infor- 
mation as  to  which  mechanism  is  involved. 

Case  E:  Inhibition  parallels  the  SH  group  reaction  but  levels  off. 

One  must  assume  that  complete  combination  of  the  enzyme  with  the 
mercurial  does  not  produce  complete  inhibition,  which  could  easily  be  the 
case  if  the  SH  groups  reacted  are  sufficiently  far  from  the  active  center  so 
that  the  introduced  side  chain  would  modify  only  the  catalysis.  A  control 
without  the  enzyme  is,  of  course,  mandatory  to  eliminate  nonenzymic  com- 
ponents to  the  rate. 

Case  F:  Inhibition  develops  without  the  reaction  of  SH  groups. 

(1)  The  mercurial,  which  was  assumed  to  be  specific  for  SH  groups,  is 
not,  and  is  inhibiting  by  another  mechanism. 

(2)  Reaction  of  SH  groups  leads  to  structural  changes  by  which  more 
SH  groups  appear;  lack  of  reaction  might  be  concluded  from  titrations  in 
which  residual  SH  groups  are  determined,  not  by  direct  spectrophotometric 
methods. 

(3)  The  method  for  the  estimation  of  SH  group  disappearance  may  be 
faulty,  i.e.,  those  groups  reacted  by  the  mercurial  may  be  resistant  to  the 
titrating  agent. 

There  are  some  instances  in  which  non-SH  enzymes  are  inhibited  by 
mercurials,  but  in  general  a  result  of  this  type  should  suggest  a  re-examina- 
tion of  the  methods  used. 

Let  us  now  examine  a  simple  system  in  greater  detail.  An  enzyme  having 
2  SH  groups  per  molecule,  one  of  these  groups  involved  in  the  catalysis  and 
one  not,  is  treated  with  increasing  amounts  of  a  mercurial;  how  will  the 
relative  reactivities  of  the  SH  groups  affect  the  results?  Representing  the 
mercurial  by  M,  the  two  equilibria  may  be  written: 


802  7.  MERCURIALS 

If  we  assume  that  the  mercurial  is  so  tightly  bound  that  fiee  mercurial  in 
the  medium  may  be  neglected,  the  conservation  equations  take  the  form: 

(M,)     =   (MSJ   +  (MS,)  (7-8) 

(Sr).  =  (Si-)     +  (MSJ  (7-9) 

(Sr)e  =  (S,-)     +  (MS,)  (7-10) 

The  fractions  of  each  S~  group  combined  with  mercurial  will  be  represented 
by  fy  and  f^,  and  it  is  easy  to  show  that  these  are  related  by: 


l-/i         1-/2 

The  ionization  of  the  SH  groups  has  purposefully  been  assumed  to  be  com- 
plete in  order  to  simplify  the  expressions.  If  S^"  is  assumed  to  be  necessary 
for  enzyme  activity,  i  =  fi  =  (MSi)/Si-)^,  while  the  reaction  of  Sg"  is  with- 
out effect  on  the  activity.  Finally,  we  shall  designate  by  r  the  fraction  of 
the  total  SH  groups  reacted  with  the  mercurial,  this  being  determined  by 
titration.  In  a  specific  case  where  (S^-)^  =  (82")^  =  10"^  M,  if  we  vary  the 
ratio  K2IK1  —  i.e.,  the  relative  affinities  of  the  S~  groups  for  the  mercurial  — 
the  curves  in  Fig.  7-20  are  obtained,  r  always  being  linear  to  complete  reac- 
tion while  i  can  follow  any  of  the  curves  between  KJE^  =  0  and  00.  When 
K2IK1  <  1,  the  situation  corresponds  to  case  C  in  Fig.  7-19,  and  when 
KJKy  >  1,  it  corresponds  to  case  D.  One  thing  we  immediately  note  is 
that  the  affinities  of  the  mercurial  for  the  different  S"  groups  must  be  quite 
different  if  the  i  curves  are  to  deviate  from  the  r  line  appreciably;  i.e., 
unless  K2IK1  is  much  greater  or  much  less  than  1,  it  will  be  difficult  to 
demonstrate  that  only  one  of  the  S~  groups  is  necessary  for  the  enzyme 
activity.  This  treatment  can  be  readily  extended  to  enzymes  with  more 
than  2  S~  groups,  in  which  case: 

fiKi  fiKz  f„K„  (7-12) 


1-/1         1-A        ■■■        l-/n 

and  to  situations  in  which  more  than  one  S"  group  are  involved  in  the 
catalysis.  One  can  also  plot  i  against  r  to  obtain  curves  characteristic  of 
the  various  situations  described  above. 

A  few  examples  of  the  different  types  of  behavior  are  summarized  here, 
as  far  as  it  is  possible  to  evaluate  the  data  published. 

Type  A 

Lactate  dehydrogenase  —  beef  heart  (Millar  and  Schwert,  1963) 

Malate  dehydrogenase  —  pig  heart  (Wolfe  and  Neilands,  1956;  Pfleiderer  et  al.,  1962) 

3-Phoshoglyceraldehyde  dehydrogenase  —  yeast  (Velick,   1953) 

Pyrophosphatase  —  pig  brain  (Seal  and  Binkley,  1957) 

Succinate  dehydrogenase  —  rat  liver  (Hirade  and  Hayaishi,  1953) 


INHIBITION  OF  ENZYMES 


803 


0  8- 


0.6 


0.2- 


FiG.  7-20.  Theoretical  curves  showing  the  relationship  between 
reaction  of  enzyme  SH  groups  and  inhibition  of  the  enzyme  ac- 
tivity for  an  enzyme  containing  two  SH  groups,  only  one  of  which 
is  involved  in  the  catalytic  activity.  (SHi)^  =  (SHa)^  =  0.01  mM 
and  K2/K1  varied  as  indicated  by  the  numbers  on  the  curves;  r  is 
the  fraction  of  SH  groups  reacted  and  is  represented  by  the 
straight  line.  It  is  assumed  that  free  mercurial  is  zero. 


Type  B 

Catalase  —  beef  liver   (Schiitte  and  Niirnberger,    1959) 
Enolase  —  rabbit  muscle  (Holt  and  Wold,  1961) 
Hexokinase  —  yeast   (Fasella  and  Hammes,   1963) 

Type  C 

ATPase  —  moysin  (Singer  and  Barron,  1944;  Barany,  1959;  Gilmour  and  Gellert, 

1961) 
Alcohol  dehydrogenase  —  yeast  (Barron  and  Levine,  1952) 

Aldolase  —  rabbit  muscle  (Swenson  and  Boyer,  1957;  Szabolcsi  and  Biszku,  1961) 
/?- Amylase  —  barley   (Rowe  and  Weill,    1962) 
Malate  dehydrogenase  —  pig  heart   (Thome  and  Kaplan,   1963) 
Phosphorylase  —  rabbit  muscle  (Madsen  and  Cori,  1955,  1956) 
Rhodanese  —  beef  liver  (Sorbo,  1963) 
Urease  —  jack  bean  (Hellerman  et  ah,  1943) 
Xanthine  oxidase  —  milk  (Gilbert,  1963) 

Type  D 

Aldehyde  oxidase  —  rabbit  liver  (Rajagopalan  et  al.,  1962) 


804  7.  MERCURIALS 

Carbamyl-P  synthetase  —  frog  liver  (Marshall  et  ah,   1961) 

Cytochrome  c  reductase  —  calf  liver  (P.  Strittmatter,   1959) 

Malate  dehydrogenase  —  beef  heart  (mitochondria)   (Siegel  and  Englard,   1962). 

Type  E 

Lactate  dehydrogenase  —  pig  muscle  (Jecsai  and  Elodi,  1963) 
Phosphoglucomutase  —  rabbit  muscle  (Milstein,  1961) 
Phosphorylase  —  potato  (Lee,  1960  b). 

No  clear-cut  example  of  type  F  has  been  reported,  but  presumably  the 
inhibition  of  D-amino  acid  oxidase  by  p-MB,  which  is  competitive  with 
respect  to  the  benzoate  portion  of  the  inhibitor,  would  fall  into  this  category. 
Certain  enzymes  have  been  shown  to  have  a  single  SH  group  at  the  active 
center  and  necessary  for  activity,  namely,  papain  (Kimmel  and  Smith,  1957; 
Finkle  and  Smith,  1958;  Sanner  and  Pihl,  1963),  ficin  (Liener,  1961),  and 
glycerol-P  dehydrogenase  (van  Eys  et  al.,  1959).  They  seem  generally  to 
belong  to  type  A. 

A  few  selected  titration  curves  may  further  illustrate  the  relations  be- 
tween SH  reaction  and  inhibition.  The  titration  of  3-phosphoglyceraldehyde 
dehydrogenase  with  p-MPS  has  been  discussed  (Fig.  7-12)  and  is  seen  to 
follow  type  A  behavior  (deviating  toward  type  C),  although  the  release  of 
NAD  is  not  exactly  parallel  to  the  disappearance  of  SH  groups.  Yeast  alcohol 
dehydrogenase  contains  10-12  SH  groups  per  molecule  but  some  inhibition 
occurs  when  only  one  is  reacted,  although  the  curve  (Fig.  7-21)  shows  the 
inhibition  at  first  to  lag  behind;  it  is  difficult  to  know  if  this  is  type  C  or  D. 
ATPase  presents  a  more  complex  situation  (Fig.  7-22)  since  reaction  of  the 
first  4  SH  groups  seems  to  produce  only  some  stimulation  of  the  activity, 
reaction  of  the  next  2  SH  groups  causing  complete  inhibition.  Other  prop- 
erties of  myosin,  e.g.,  the  ability  to  complex  with  actin  and  the  viscosity 
response  of  actomyosin  to  ATP,  are  more  directly  dependent  on  the  reac- 
tion of  the  first  SH  groups.  ITPase  activity  conforms  more  to  type  A  be- 
havior. The  relationship  of  ATPase  activity  to  SH  reaction  depends  on  the 
state  of  the  enzyme  (Fig.  7-23),  no  initial  stimulation  being  observed  when 
EDTA  is  the  activator  instead  of  Ca++.  The  titration  of  muscle  phosphor- 
ylase a  gives  partial  type  C  behavior,  but  over  most  of  the  range  there  is  a 
linear  relationship  between  SH  reaction  and  inhibition  (Fig.  7-24).  Since 
the  inhibition  may  be  completely  reversed  by  cysteine,  it  is  unlikely  that 
a  secondary  inactivation  is  involved.  Microsomal  cytochrome  c  reductase 
demonstrates  typical  type  D  behavior,  one  SH  group  being  closely  related 
to  the  enzyme  activity,  as  shown  by  the  extrapolation  of  the  inhibition 
curve  to  complete  inhibition  (Fig.  7-25),  from  which  it  may  be  estimated 
that  K2IK1  is  around  25-50. 

Accurate,  reliable,  and  directly  interpretable  titrations  of  enzymes  are 
not  easy  to  perform  in  some  cases.  Some  of  the  possible  difficulties  which 
may  arise  will  be  summarized.  (1)  There  is  a  failure  to  reach  equilibrium, 


INHIBITION  OF  ENZYMES 


805 


i.e.,  reaction  of  the  mercurial  is  not  complete  at  the  time  chosen  for  the 
readings.  Some  enzyme  SH  groups  react  almost  instantaneously  with  mer- 
curials and  others  require  30-60  min  at  least  (see  page  809).  Kinetic  studies 
should  always  accompany  any  titration;  to  decide  arbitrarily  that  n  min- 
utes of  incubation  with  the  mercurial  at  each  concentration  is  adequate  is 
not  a  satisfactory  procedure.  (2)  The  enzyme  may  be  altered  structurally 
by  the  mercaptide  formation  so  that  new  SH  groups  are  progressively  ex- 


FiG.  7-21.    Titration   of  yeast    alcohol   clehydrogenase 
with    77-]\IB.    sliowiny    the    nonlinearity    at    low    con- 
centrations   of    the     mercurial.     (From     Barron    and 
Levine,   19.-i2.) 


posed,  in  which  case  there  is  no  clear  end-point  and  the  results  do  not 
correspond  to  the  original  native  enzyme.  Sometimes  the  stability  of  the 
enzyme  can  be  increased  by  creating  a  more  physiological  environment. 
(3)  The  enzyme  SH  groups  may  be  oxidized  during  the  titration,  reducing 
the  number  of  titratable  SH  groups.  Use  of  oxygen-free  solutions  and  a 
nitrogen  atmosphere  often  eliminates  this  problem.  (4)  The  mercurial  may 
react  with  other  groups  or  other  components  of  the  enzyme  system,  e.g., 
in  the  spectrophotometric  titration  with  p-MB  or  jj-MPS,  causing  absorp- 
tion changes  unrelated  to  SH  groups.  (5)  The  mercurial  may  do  something 
that  secondarily  alters  the  ultraviolet  absorption,  e.g.,  split  off  a  coenzyme, 
as  demonstrated  for  NADH:  lipoamide  oxidoreductase  (Palmer  and  Massey, 
1962).  (6)  The  presence  of  substances,  especially  buffers,  reacting  with  the 


806 


7.  MERCURIALS 


mercurial  may  alter  the  rate  and  extent  of  reaction  with  the  SH  groups. 
The  titrations  of  yeast  alcohol  dehydrogenase  by  p-MB  in  phosphate  and 
in  THAM  buffers  at  pH  7.5  are  quite  different  (Hoch  and  Vallee,  1960). 
It  is  probably  advisable  to  reduce  the  buffer  concentration  as  far  as  possible. 
(7)  The  number  of  reactive  SH  groups  on  an  enzyme  and  the  titration  of 
these  groups  vary  with  several  experimental  conditions,  such  as  pH,  temper- 
ature, and  absence  or  presence  of  substrate,  and  the  question  often  arises 


100 


80 


60 


40 


S  20 


-20 


SENSITIVrTY    OF 

ACTOMYOSIN    TO 

ATP 


8  ,     6 

SH  (M/IO    G       MYOSIN) 


Fig.  7-22.  Titration  of  myosin  with  mersalyl,  showing 
the  effects  on  ATPase  activity,  the  ability  to  form 
actomyosin,  and  the  sensitivity  of  the  actomyosin 
to  ATP  (measured  by  viscosity  changes).  (From 
Barany,  1959.) 


as  to  what  conditions  are  optimal.  Titrations  are  often  done  at  unphysi- 
ological  pH's  because  reaction  is  faster  or  more  complete,  but  it  must  be 
remembered  that  the  results  do  not  necessarily  apply  to  the  enzyme  under 
normal  conditions.  Boyer  and  Segal  (1954)  showed  definite  difference  in 
the  titration  of  3-phosphoglyceraldehyde  dehydrogenase  spectrophotome- 
trically  at  pH  4.6  and  7,  and  this  is  probably  a  general  phenomenon.  The 
effect  of  temperature  is  weU  illustrated  by  the  study  of  yeast  hexokinase, 


INHIBITION  OF  ENZYMES 


807 


the  SH  groups  at  the  active  center  becoming  unavailable  for  reaction  below 
30°  (Barnard  and  Ramel,  1962).  The  presence  of  substrate  may  either  fa- 
cilitate reaction  of  SH  groups  —  as  with  xanthine  oxidase  (Fridovich  and 
Handler,  1958)  and  myosin  ATPase  (Gilmour  and  Grellert,  1961)  —  or 
protect  certain  SH  groups.  The  question  as  to  which  pH,  temperature,  and 
medium  should  be  used,  or  whether  substrate  or  coenzyme  should  be  pres- 
ent during  the  incubation,  can  only  be  answered  generally  by  stating  that 


REACTION    WITH 
SH     GROUPS 


0.25 


020 


0.05 


16  20 

MMOLESiIO^/MG 


Fig.  7-23.  Titration  of  myosin  ATPase  with  p-MB,  showing 

the  different  responses  of  the  EDTA-treated  and  Ca++-acti- 

vated  activity.   (From  Kiellye  and  Bradley,  1956.) 


whenever  possible  one  should  strive  for  physiological  conditions.  It  is  nec- 
essary, of  course,  to  vary  these  factors  in  many  instances  in  order  to  study 
the  behavior  of  the  SH  groups,  but  the  variation  should  be  from  a  standard 
set  of  conditions  designed  to  provide  information  relevant  to  the  enzyme 
in  a  normal  state. 

It  is  frequently  difl&cult  to  determine  with  certainty  the  total  number  of 
free  SH  groups  in  a  native  enzyme  under  standard  conditions  and  especially 
to  relate  certain  SH  groups  to  the  catalytic  activity.  Thorne  and  Kaplan 
(1963)  titrated  pig  heart  malate  dehydrogenase  with  p-MB,  allowing  1  hr 


808 


7.  MERCURIALS 


for  reaction  at  25°,  and  could  obtain  no  reliable  end-point.  As  the  molar 
ratio  of  I  :  E  is  increased  there  is  no  marked  effect  on  the  activity,  except 
for  a  slight  stimulation,  until  after  a  value  of  5  is  exceeded,  and  then  there 
is  a  progressive  loss  of  activity  as  the  ratio  is  elevated,  nearly  complete  inhi- 
bition occurring  at  a  value  of  21.6.  It  is  impossible  to  interpret  these  data  in 
terms  of  relating  SH  groups  to  activity.  Indeed,  it  is  likely  that  the  enzyme 
is  structurally  altered  so  that  SH  groups  normally  not  accessible  are  second- 
arily unmasked,  since  good  titrations  can  be  determined  with  the  urea- 


100 


2         4         6         8         10 

MOLES    p-MB/MOLES     ENZYME 


Fig.  7-24.  Titration  of  muscle  phosphorylase  a  with 
p-MB.   (From  Madsen  and  Cori,   1955.) 


denatured  enzyme.  The  aspartate:  or-ketoglutarate  transaminase  from  pig 
heart  contains  a  total  of  7  SH  groups;  when  2  moles  of  p-MB  per  mole  of 
enzyme  are  added,  this  reacting  with  1-2  SH  groups,  the  activity  is  reduced 
by  50%  (Turano  et  al.,  1963).  Such  data  again  are  uninterpretable  and  it  is 
impossible  to  conclude  that  SH  groups  are  related  in  any  way  to  the  cat- 
alysis. Di  Sabato  and  Kaplan  (1963)  titrated  the  lactate  dehydrogenases 
from  a  variety  of  sources  with  both  Hg++  and  p-MB.  The  total  number  of 
SH  groups  per  mole  of  enzyme  varied  from  17  to  27  but  generally  inactiva- 
tion  occurred  when  4  moles  of  mercurial  were  bound  for  each  mole  of  en- 
zyme. It  is  likely  that  no  major  configurational  changes  occur  because  no 
alterations  of  fluorescence,  sedimentation  constant,  or  rotatory  dispersion 
and  no  immunological  changes  could  be  detected,  and  furthermore  cysteine 
could  essentially  completely  reverse  the  inhibition.  They  felt  that  certain 
SH  groups  are  part  of  the  active  site  rather  than  being  vicinal,  since  statis- 


INHIBITION  OF  ENZYMES 


809 


tically  it  is  unlikely  that  in  all  the  dehydrogenases  such  a  distribution  would 
occur.  On  the  other  hand,  Jecsai  and  Elodi  (1963)  claimed  that  pig  muscle 
lactate  dehydrogenase  in  the  native  state  does  not  react  with  p-MB,  but 
that  at  pH  10  the  blocking  of  20  SH  groups  leads  to  50%  inactivation.  They 
concluded  that  in  this  particular  enzyme  the  SH  groups  are  not  at  all  in- 
volved in  the  catalysis.  These  examples  only  illustrate  some  of  the  problems 
which  arise  in  enzyme  titrations  and  emphasize  that  a  program  for  relating 
SH  groups  to  enzyme  activity  cannot  be  undertaken  lightly. 


100 


^      20 


Fig.  7-25.  Titration  of  microsomal  cytochrome  c 
reductase  with  p-MB.  The  sohd  curve  gives  the 
development  of  the  inhibition  as  equivalents  of 
mercurial  are  added.  The  dashed  line  continuing 
from  the  initial  linear  portion  of  the  curve  shows 
that  one  SH  group  is  required  for  activity;  the 
other  dashed  curve  shows  the  assumed  reaction 
with  SH  groups  (it  was  not  experimentally  de- 
termined). It  may  be  estimated  that  K^IK^  is 
near  25-50  if  there  are  two  SH  groups.  (From  P. 
Strittmatter,   1959.) 


Kinetics  of  Mercaptide  Formation  and  Development  of  Inhibition 

Apparently  SH  groups  range  in  reactivity  all  the  way  from  those  which 
combine  with  mercurials  so  rapidly  that  the  rates  are  difficult  to  measure, 
to  those  which  are  completely  blocked  and  do  not  react  at  all.  It  is  thus 


810 


7.  MERCUKIALS 


not  surprising  that  one  finds  a  great  deal  of  variation  in  the  rates  at  which 
enzyme  SH  groups  react  and  at  which  inhibition  occurs.  In  some  cases  the 
inhibition  has  been  said  to  appear  instantaneously,  or  to  reach  full  magni- 
tude within  1-2  min;  such  is  the  inhibition  of  succinate  dehydrogenase 
(Fig.  1-12-12)  (Slater,  1949),  bromelain  (Murachi  and  Neurath,  1960),  pyru- 
vate decarboxylase  (Stoppani  et  al.,  1953),  and  leucine  decarboxylase  (Sut- 
ton and  King,  1962).  Then  there  are  enzymes  which  require  about  5  min 
for  maximal  inhibition  to  develop;  examples  are  catalase  (Cook  et  ah,  1946), 
transaminases  (Grein  and  Pfleiderer,  1958;  Segal  et  al.,  1962)  and  phospho- 
glucomutase  (Milstein,  1961),  although  in  the  last  instance  only  2  of  the 
3  SH  groups  react  so  rapidly.  It  is  interesting  to  note  that  Nygaard  (1955) 
has  reported  marked  differences  in  rates  between  the  mercurials,  lactate 
dehydrogenase  being  very  rapidly  inhibited  by  Hg++  but  only  slowly  by 
p-MB.  The  next  group  of  enzymes  seems  to  require  about  15-20  min  for 
complete  inhibition:  3-phosphoglyceraldehyde  dehydrogenase  (Boyer  and 
Segal,  1954)  and  enolase  (Malmstrom,  1962)  may  be  cited.  These  are,  of 
course,  arbitrary  categories  and  if  one  knew  the  rates  of  reaction  for  many 


100 
80 
60 
40 
20 
0 


. 

^______ 

/35° 

^^                   "^ 

/      / 

/zo" 

(/ 

ALCOHOL 

DEHYDROGENASE 

0  20 

TIME      » 


40 


60 


80 

MIN 


Fig.  7-26.  Rates  of  inhibition  of  liver  alcohol 
dehydrogenase  by  p-MB,  at  pH  7.6  and  two 
different  temperatures.  ADH  =  1.78x  IQ-s  M, 
p-MB  =  10-«  M,  and  NAD  =  3x10"*  M. 
(From  WaUenfels  et  al,   1959.) 


enzymes,  there  would  be  a  continuous  distribution,  and  furthermore  the 
rate  in  any  particular  case  will  depend  on  a  number  of  factors,  so  that  the 
values  given  above  and  below  must  be  taken  as  applying  only  to  the  ex- 
perimental conditions  imposed  on  each  enzyme. 

More  interesting  are  those  enzymes  which  react  slowly  enough  with  mer- 
curials for  the  kinetics  to  be  investigated.  Some  typical  curves  for  p-MB  are 
given  in  Figs.  7-26  and  7-27,  and  similar  rate  curves  have  been  previously 
presented  for  cholinesterase  (Fig.  1-12-8)  and  lactate  dehydrogenase  (Fig. 
1-12-11).  The  results  in  Figs.  7-26  and  7-27  have  been  exponentially  plotted 


INHIBITION  OF  ENZYMES 


811 


in  Fig.  7-28  to  indicate  more  clearly  the  relative  rates  of  inhibition  (see 
Eq.  1-12-14  and  Figs.  1-12-3  and  1-12-9),  the  slopes  being  proportional  to 
the  bimolecular  rate  constants.  One  notes  that  most  of  the  curves  deviate 
from  linearity,  frequently  at  high  inhibitions;  this  is  probably  an  expression 
of  the  different  relative  reactivities  of  the  SH  groups  on.  a  single  enzyme, 
some  reacting  initially  at  a  rapid  rate  and  others  reacting  more  slowly. 
The  rate  constants  have  been  calculated  for  some  enzymes,  e.g.,  18.8  liters/ 
mole/sec  for  glutamate  decarboxylase  (Shukuya  and  Schwert,  1960),  51  li- 
ters/mole/sec for  muscle  phosphorylase  (Madsen  and  Cori,  1956),  and  61.4 
liters/mole/sec  for  heart  lactate  dehydrogenase  (Takenaka  and  Schwert, 
1956),  in  all  cases  p-MB  being  the  inhibitor.  The  effects  of  temperature  and 
mercurial  concentration  on  the  rates  of  inhibition  are  well  illustrated  for 
alcohol  dehydrogenase  (Fig.  7-26)  and  /?-fructofuranosidase  (Fig.  7-27),  re- 
spectively. Glutamate  decarboxylase  presents  an  interesting  phenomenon, 
in  that  exposure  of  the  enzyme  to  low  temperatures  appears  to  liberate 
additional  SH  groups  (Fig.  7-27). 


)9  -  FRUCTOFURANOSIDASE 


GLUTAMATE   DECARBOXYLASE 


80  "  0   60   120   180  240  300  360 

MIN  TIME      »-  MIN 


Fig.  7-27.  Rate  titrations  of  various  enzymes  with  p-MB.  The  reaction 
of  SH  groups  was  determined  by  absorption  changes  at  250  m/<.  Phos- 
phorylase a  from  rabbit  muscle:  p-MB  =  0.04  mM,  pH  6.7,  and  21° 
(Madsen  and  Cori,  1956.)  Lactate  dehydrogenase  from  heart:  p-MB 
=  0.00448  Mm,  pH  6.8,  and  25°  (Takenaka  and  Schwert,  1956.) 
^-Fructofuranosidase  from  Neurospora:  p-MB  concentrations  given  in 
the  graph,  pH  6.8,  and  0°  (Metzenberg,  1963).  Glutamate  decarboxylase 
from  E.  coli:  preincubation  with  p-MB  for  4  hr  at  either  0°  or  25°, 
and  reaction    run  at  25°   and    pH   6.5    (Shukuya   and  Schwert,    1960). 


812 


7.  MERCURIALS 


Progressive  inhibition  or  inactivation  of  enzymes  by  mercurials  is  very 
common  and  takes  a  variety  of  forms.  Epididymal  a-mannosidase  is  inhi- 
bited 62%  by  0.01  mM  Hg++  without  preincubation  with  the  inhibitor; 
the  inhibition  is  67%  at  30  min,  77%  at  60  min,  and  88%  at  120  min 
(Conchie  and  Hay,  1959).  This  is  one  of  the  numerous  examples  in  which 
an  enzyme  is  rapidly  inhibited  to  a  certain  level,  further  increase  in  inhibi- 


FiG.  7-28.  Logarithmic  plots  of  if—  i  for  the  en- 
zymes in  Figs.  7-26  and  7-27,  either  loss  of  ac- 
tivity or  reaction  of  SH  groups  being  used  as  a 
measure  of  the  reaction  with  p-MB.  These  curves 
have  been  estimated  from  the  published  curves 
and  hence  are  not  strictly  accurate,  but  indicate 
the  relative  rates  for  the  more  slowly  reacting 
SH  groups.  1,  Alcohol  dehydrogenase  (35°); 
2,  Alcohol  dehydrogenase  (20");  3,  phosphorylase; 
4,  lactate  dehydrogenase;  5,  glutamate  decar- 
boxylase (25°);  6,  glutamate  decarboxylase  (0"); 
7,  ^-fructofuranosidase  (0.1  mM);  8,  ^-fructo- 
furanosidase  (0.02  raM);  9,  j3-fructofuranosidase 
(0.01  mM);     10,  /3-fructofuranosidase  (0.04  mM.) 


tion  being  slow.  Such  behavior  is  not  surprising  when  one  measures  SH 
reaction,  since  one  assumes  generally  the  occurrence  of  SH  groups  of  dif- 
ferent reactivities;  thus  when  3-phosphoglyceraldehyde  dehydrogenase  is 
titrated  with  p-MB,  11  SH  groups  react  immediately,  but  3  more  require 
at  least  40  min  (Koeppe  et  al.,  1956),  and  aldolase  behaves  very  similarly, 
7  SH  groups  being  blocked  rapidly  and  3-4  more  groups  taking  40  min  for 
reaction  (Szabolcsi  and  Biszku,  1961).  But  the  interpretation  of  inhibition 
following  such  a  time  course  is  not  so  clear.  If  a  certain  level  of  inhibition 
is  reached  rapidly  and  then  the  rate  falls  off  markedly,  one  must  assume 
that  the  enzyme  is  not  completely  mhibited  when  its  rapidly  reacting  SH 


INHIBITION  OF  ENZYM.ES  813 

groups  are  combined  with  mercurial  (assuming  that  there  is  sufficient  mer- 
curial to  react  with  all  these  groups).  The  more  slowly  developing  inhibition 
could  be  due  to  reaction  of  less  readily  available  SH  groups  or  to  a  secondary 
inactivation  following  the  initial  mercaptide  formation.  Quite  different  re- 
sults are  obtained  with  amylase,  mersalyl  at  1  mM  not  inhibiting  at  all 
during  the  first  hour,  but  slowly  inhibiting  until  there  is  40%  depression 
after  48  hr  (Muus  et  ah,  1956),  or  with  bromelain,  jJ-MB  inhibiting  only 
25%  after  4  hr  and  80%  after  20  hr  at  0.1  mM  (Ota  et  al,  1961).  Many 
different  time  courses  of  inhibition  are  observed  and  it  is  likely  that  the 
major  factors  involved  are  (1)  the  relative  reactivities  of  the  SH  groups, 

(2)  the  relationship  between  the  SH  groups  and  the  catalytic  activity,  and 

(3)  the  tendency  for  structural  changes  leading  to  inactivation  to  occur. 
However,  it  is  quite  clear  that  many  enzymes  react  quite  slowly  with  mer- 
curials and  require  2-4  hr  (and  occasionally  more)  to  complete  the  process. 
Such  enzymes  are  difficult  to  titrate,  since  one  does  not  know  how  many 
of  the  SH  groups  finally  reacted  were  originally  present,  and,  when  one  is 
adding  increasing  amounts  of  mercurial  to  correlate  mercaptide  formation 
and  inhibition,  it  is  not  easy  to  decide  on  the  optimal  preincubation  in- 
terval. 

In  the  previous  section  the  correlation  between  inhibition  and  SH  reac- 
tion by  mercurials  was  considered  in  terms  of  variable  quantities  of  mer- 
curial. Another  approach  to  relate  these  phenomena  is  to  determine  their 
changes  with  time  at  a  particular  mercurial  concentration.  If  the  SH  groups 
which  are  combined  initially  are  necessary  for  enzyme  activity,  one  would 
expect  inhibition  to  parallel  blocking  of  these  groups;  if  the  most  readily 
reacting  SH  groups  are  not  related  to  activity,  or  the  enzyme  undergoes 
progressive  inactivation,  the  inhibition  may  lag  behind  mercaptide  forma- 
tion. Madsen  and  Cori  (1956)  observed  that  inhibition  of  phosphorylase  by 
p-MB  developed  more  slowly  than  the  change  in  absorbance  at  250  mjit 
(Fig.  7-27),  so  that  when  50%  of  the  reactive  SH  groups  had  been  blocked 
the  inhibition  was  only  14%.  If  the  SH  blocking  itself  is  not  responsible 
directly  for  the  inhibition,  but  initiates  an  unfolding  of  the  enzyme,  the 
rate  of  inhibition  may  be  more  dependent  on  the  rate  of  configurational 
change.  In  the  case  of  phosphorylase,  we  have  seen  that  splitting  into  sub- 
units  occurs  during  reaction  with  p-MB,  so  the  rate  at  which  this  occurs 
may  have  something  to  do  with  the  inhibition  rate.  Inasmuch  as  cysteine 
reverses  the  inhibition  completely,  marked  structural  alterations  would  not 
be  very  likely. 

Another  phenomenon  which  must  be  taken  into  account  in  kinetic  studies 
is  the  spontaneous  recovery  of  enzyme  activity  in  the  presence  of  the  mer- 
curial, first  observed,  I  believe,  by  von  Euler  and  Svanberg  (1920)  in  studies 
of  the  inhibition  of  yeast  /5-fructofuranosidase  by  Hg++.  Reisberg  (1954) 
reported  that  the  inhibition  of  choline  acetylase  by  p-MB  is  less  at  30  min 


814 


7.  MERCURIALS 


than  at  10  min.  Other  more  recently  observed  examples  of  this  include 
epididymal  /5-galactosidase  with  low  concentrations  (0.0002  milf )  of  Hg++ 
(Conchie  and  Hay,  1959),  xanthine  oxidase  with  0.44  mM  p-MB,  which 
inhibits  NADH  oxidation  50%  initially  but  less  and  less  as  the  reaction 
proceeds  (Westerfeld  et  at.,  1959),  and  leucine  decarboxylase  with  0.005  mM 
p-MB  (Sutton  and  King,  1962).  The  most  marked  spontaneous  recovery  is 
seen  with  pig  heart  lactate  dehydrogenase,  the  rate  and  degree  of  reactiva- 
tion being  dependent  on  the  molar  ratio  of  p-MB  to  enzyme  (Fig.  7-29) 


50 

HOURS 


Fig.  7-29.  Effects  of  p-MB  on  pig  heart 
lactate  dehydrogenase,  showing  the  ini- 
tial inhibition  and  the  spontaneous  reac- 
tivation. The  numbers  on  the  curves 
are  the  molar  ratios  of  p-MB  to  LDH. 
(From  Gruber  et  al.,  1962.) 


(Gruber  et  al.,  1962).  The  most  common  explanation  for  such  recovery  is 
a  slow  migration  of  the  mercurial  from  those  groups  initially  attacked  to 
other  groups  not  involved  in  the  enzyme  activity.  The  rates  at  which  var- 
ious SH  groups  react  with  a  mercurial  are  not  necessarily  related  to  the 
affinities  of  the  groups  for  the  mercurial.  Groups  which  bind  the  mercurial 
very  tightly  may  be  masked  and  react  very  slowly,  as  fairly  conclusively 
demonstrated  for  myosin  ATPase  by  Gilmour  and  Gellert  (1961).  Another 
factor  which  may  be  of  importance  when  the  inhibition  decreases  during 
the  period  when  the  enzyme  activity  is  measured,  as  was  the  case  with 
leucine  decarboxylase,  is  the  displacement  of  the  mercurial  by  the  substrate. 
Leucine  was  shown  to  protect  the  enzyme  against  2)-MB  and  it  could  even- 
tually overcome  the  inhibition  somewhat,  especially  since  its  concentration 
was  some  1000  times  greater  than  the  mercurial.  A  substrate  might  also  be 
able  to  restore  toward  a  normal  configuration  a  slightly  luxated  active  cen- 
ter, substrates  being  known  to  stabilize  the  active  forms  of  certain  enzymes. 


INHIBITION  OF  ENZYMES  815 

This  phenomenon  of  spontaneous  recovery  must  be  even  more  common  in 
celkilar  preparations  than  with  pure  enzymes,  because  there  is  much  greater 
opportunity  for  redistribution  of  the  mercurial. 

Stimulation  of  Enzymes   by  Mercurials 

Mercurials  is  common  with  other  heavy  metals  and  SH  reagents  frequent- 
ly increase  enzyme  activity,  especially  at  low  concentration,  the  action- 
concentration  curves  being  biphasic.  Polis  and  Meyerhof  (1947)  first  ob- 
served the  stimulation  of  Ca++-activated  myosin  ATPase  by  PM,  a  30-40% 
elevation  of  the  rate  occurring  with  concentrations  between  0.005  and  0.12 
mM,  and  this  has  been  confirmed  in  several  more  recent  reports,  the  degree 
of  stimulation,  however,  varying  greatly  with  the  experimental  conditions. 
Many  different  types  of  enzyme  exhibit  this  phenomenon  (Table  7-11)  but 
the  mechanisms  involved  have  only  rarely  been  clarified.  Let  us  briefly 
consider  some  possible  mechanisms  and  what  relevant  evidence  is  available. 

(A)  The  mercurial  inactivates  a  naturally  occurring  inhibitor.  If  an  in- 
hibitor is  isolated  w^ith  the  enzyme  and  is  suppressing  the  activity,  and  if 
this  inhibitor  is  an  SH  protein  (as  many  natural  inhibitors  seem  to  be),  a 
mercurial  by  reacting  preferentially  with  the  inhibitor  may  release  the  en- 
zyme from  its  inhibition.  J.  S.  Roth  (1953  a,  1956,  1958)  has  been  a  pro- 
ponent of  this  theory  with  respect  to  the  activation  of  rat  liver  homogenate 
ribonuclease  by  p-MB  or  PM,  and  has  found  a  natural  inhibitor  with  which 
the  mercurials  react  at  concentrations  having  no  direct  effect  on  ribonu- 
clease. The  degree  of  stimulation  varies  with  the  tissue  from  which  the  ri- 
bonuclease is  obtained  —  all  the  way  from  0%  with  pancreas,  57%  with 
brain,  104%  with  muscle,  284%  with  liver,  to  1500%  with  ascites  carci- 
noma —  and  this  may  be  due  to  the  different  amounts  of  inhibitor  present 
(Ellem  and  Colter,  1961).  Indirect  evidence  often  points  to  such  a  mechan- 
ism for  other  enzymes.  PhiUips  and  Langdon  (1962)  found  that  p-MB  stim- 
ulates microsomal  NADPHxytochrome  c  reductase  but  only  inhibits  the 
purified  enzyme.  Of  course,  the  activation  could  also  be  due  to  some  effect 
of  the  mercurial  on  the  microsomal  structure.  It  has  also  been  noted  occa- 
sionally that  stimulation  occurs,  and  is  relatively  constant,  over  a  wide 
range  of  mercurial  concentration,  inhibition  appearing  rather  suddenly  when 
this  range  has  been  exceeded,  and  this  indicates  some  component  with  which 
the  mercurial  reacts  readily  and  completely. 

(B)  The  mercurial  reacts  with  the  substrate  to  labilize  it.  Ledoux  (1953) 
initially  attempted  to  explain  the  stimulation  of  ribonuclease  by  p-MB  as 
due  to  a  reaction  with  RNA,  this  favoring  in  some  manner  the  enzymic 
hydrolysis,  and  detected  spectral  changes  upon  mixing  RNA  and  p-MB 
(see  page  741).  Although  mercurials  do  complex  with  nucleic  acids,  it  is 
doubtful  if  this  is  a  major  factor  in  the  activation.  The  proper  preincubation 


816 


7.  MERCURIALS 


& 
o 
<A 

n 
o 

_  H 

^  "^ 

la 


P2 


,C5      CS 

^3 


in  CO 
lO   »o 

05     05 


ri  >>  ^ 


-C     =0     C     eS 


^^  IC    >c 

c8  C5    OS 

*  Co 

o  ^ 
c    S 


-^  «^     -in 


6  i5 


c  „ 
o  S 
o  -5 


O      C5 


S      =3 


pq  2       3 
c 


Ci     CO 


cs    £ 


=3      S 


Zh^omwwp^k      Stf 


OiCOtJHOiOICcQCC     »0     O 

-H      rt     tq  IC 


o   o 
o   o 

rt     CO 


<N    (M    O    O    O 
CO    (M    lO    O    O 


M    be    M 

-t  ^^ 

^    ^  ^   t^ 

>c 

"o  "3  "o  g 

>o 

o 

o 

-^    (N 

£      £      E    O 

o 

o 

m 

c^ 

d  -! 

l-O      TjH      IC      ® 

o  o  o 

d 

d  ^ 

d 

d 

CO     GO 

o  o 

-^    lO    o    o 


o   o   o   o   o   o  o 


60 

fl 

hn 

~ar 

S 

0) 

0; 

s 

o 

O 

S 

u 

or 

5. 

"^ 

CO 

05 

lO 

o 

^ 

o 

o 

o 

M 


m 


M   pq  pq   pq 


pq 


ss,  fi^    ?4. 


S       + 

J3  O 


>  ^  c 

:^  ys  'm 

^  o  o 

P5  S  S 


K 


55,      ?!,      a      ?».    Ph       ?5, 


pq   + 

^      M    §    §    g 


fi=^  p^ 


>-  > 


"^   ■"   'S 


•  —     i'     cj 


A 


KCoPqfe5PLioP-iW 


e    > 


o  pq 


c 

tn 

s 

>> 

O 

.& 

B 

'C 

0) 

C 

CO 

O 

c 

.S      03 


G      >>     >>     & 


9    :5    -^   ^   ^ 


<:<5i<;<3oPHOO 


IXHIBITION  OF  ENZYMES 


817 


>o 

m 

Ul 

, — . 

^^ 

C5 

05 

05 

n 

>— ^ 

M 

CS 

■M 

<n 

^— ' 

^-^ 

^— ' 

w5 

■^ 

05 

M 

^.^ 

M 

t£ 

:3^ 

C5 

^ 

_C 

^..^ 

(*. 

c 

_C 

^_^ 

'C 

W 

CO 

'C 

'C 

J2 

c 

^-^ 

(0 

«o 

05 

« 

® 

CK 

c 

o 

M 

Xi 

Oi 

x> 

Xi 

s 

T! 

CO 

<D 

•»— " 

a> 

<0 

S 

tc 

C5 

s 

^-^ 

s 

3 

eS 

S3 

c 

J 

^ 

i 

H-; 

hJ 

t^ 

^ 

a 

s 

« 

o 

T3 

S 

c 

c3 

-t^ 

■t^ 

-u 

•t^ 

■t^ 

a 

bi 

b4 

o 

ti 

b. 

cj 

CO 

>> 

o 

o 

O 

o 

^ 

s 

CO 

c 

o 

o 

^ 

& 

o 

g 

_a 

o 

.5     a 

a  tf  ei  a 


tf  s 


fl^     <!l 


t^  g  s 
^       ^  2  — 


ri         O       -— 


;?;  a  c» 


K    ^    -  ri 


T3 

C 

w 

C 

C5 

t-^ 

'3 

cS 

s^ 

^H 

te 

-'-^ 

^ 

0 

C 

02 

^■^ 

V 

-<t 

O 

cS 

f^ 

S 

0 

CO 

e 

c3 

T! 

C5 

C 

,^ 

"5 

— 

C 

^^ 

sa 

c3 

ce 

'-3 

03 

3 

'3 

C 

0) 

-a 

3 

"3 

"o 

cS 

"S 

'o 

0 

o: 

< 

K 

Q 

2; 

S 

t3£j 

,^ 

Lt 

lO 

^ 

O 

^ 

c 

CO 

(M 

M 

5^ 

c 

-M 

3 
O 


>  T3 

a  •- 

a>  > 
3 

':3  c 


0000000 


03  +  +    +  +  +  CQ 

a  +  +   +  +  +  ^ 

^  M    M    M  6JD  H  ^ 

9,  M  HH     ffi  ffi  W  S. 


CQ  op    pq 
S  §    S 

?s,    ?a.     a 


o 

O     -H     !tL 


pq  m  pa 
§  §  § 

a       i       ?s 


8  S 


d    d 


10    o 

d    d 


+     Cu    2    P3  P3    P3 

M     S      ^      S      g      ^      S 

K    ?!.    a    i,  Ph    a    a 


S       «! 

3 

03       O 

O    .15 

5  i 

>  <:! 


CO     tiJD 


p 

0 

<» 

,  r; 

0 

0 

m 

>, 

3 

3 

® 

0) 

C 

> 

-t^ 

-k^ 

*i 

'    ' 

,0 

-0 

-0 

-Q 

-0 

Xi 

es 

ca 

c! 

pp 

tf 

P5 

« 

£  -5 

Q>        CO 


cS       ce       » 

«    P5    !>^ 


03  -^ 


0^ 

£ 

00 

> 

4^ 

s 

s 

^ 

'~* 

-Q 

■*^ 

-0 

?!, 

CS 

cS 

s 

Ch 

tf 

K 

►^ 

-C 

1^ 

0 

ft 

P- 

m 

0 

.c 

0 

0 

n 

(J 

0 

4) 

-a 

ft 

0 
c 

c3 

>> 

0 

c3 
-4^ 

(M 

CO 

3 

0 
3 

0 

3 

<0 

<D 

U' 

T3 

-tJ 

nS 

1^. 

w 

a; 

W 

t 

»H 

b 

0 

Ch 

0 

0 

0 

0 

-a 

Q 

'O 

>. 

0 

>> 

0 

< 

0 

0 

a        :2 


f^      flH 


o     ©    4-    c 


P     -3     ^ 


o    a    -s 


T 

.!:> 

m 

M 

a 

PL^ 

>^ 

H    H    O    Ph 


3     -     j5     j3 
^     *"     Pm     Ph 


ft 

•c 

<r<i 

0) 

0) 

C 

>-. 

H 

^. 

C3 

ft 

Q1 

9> 

(i> 

0 

tl) 

,    , 

j3 
ft 

0 

L4 

0 

0 

X 

0 

0 

3 

0 
3 

0 

t^, 

ft 

C 

c 

0 

0 

Ph 

CO 

-0 

€ 

Pm 

X 

g:      ^     --I 


3       c3 


.-^     V^       3 


(A^  ^  0  ^  W 


02     H 


o    g 
-3     eg 

Is- 


ft 

4> 


CD 


O 

3     » 

-"  x: 
c  -^ 

•— I     a; 
1^ 


818  7.   MERCURIALS 

procedures  might  easily  solve  this  problem.  There  is  no  other  instance  where 
this  mechanism  has  been  postulated,  but  it  is  a  possibility  that  must  be 
borne  in  mind. 

(C)  The  mercurial  reacts  with  a  SH  group  to  create  a  more  favorable  electric 
field  at  the  active  center.  Since  the  binding  of  the  substrate  and  the  break- 
down of  the  ES  complex  are  influenced  by  the  electric  field  arising  from 
charged  groups  vicinal  to  the  active  center,  it  is  possible  that  the  binding 
of  a  charged  mercurial  could  so  alter  this  field  as  to  facilitate  the  catalysis. 
There  is  no  evidence  for  this  and  it  would  be  difficult  to  obtain. 

(D)  The  mercurial  increases  the  active  form  of  the  enzyme.  Certain  enzymes 
within  the  cell  or  as  extracted  may  be  in  an  inactive  form,  possibly  with 
the  active  center  not  in  the  proper  configuration.  Reaction  with  a  mercurial 
might  so  alter  the  protein  structure  as  to  release  the  activity,  much  as 
some  enzymes  can  be  activated  by  heat.  This  would  probably  apply  only 
to  true  activations,  i.e.,  where  the  initial  state  of  the  enzyme  is  inactive, 
as  in  the  work  of  Hilz  and  Klempien  (1959)  on  ascites  tumor  ribonuclease, 
Hg++  at  0.005  and  0.05  uvM  increasing  the  enzyme  rate  from  0  to  12  and 
24  ^moles  TCA-soluble  phosphate/hr,  respectively,  but  here  it  may  be  only 
that  the  enzyme  is  completely  inactivated  by  a  natural  inhibitor. 

(E)  The  Hg^+  ion  may  replace  a  normal  metal  ion  activator.  Carboxypepti- 
dase  which  is  normally  activated  by  Zn++  can  be  activated  with  respect  to 
the  esterase  activity  with  Hg++  (Coleman  and  Vallee,  1961).  This  situation 
is  probably  very  uncommon.  Rapoport  et  al.  (1955)  reported  that  Hg++ 
stimulates  glycerate-2,3-diphosphatase  only  in  the  presence  of  certain  or- 
ganic nitrogen  substances,  and  the  activation  was  assumed  to  be  due  to 
the  complex  formed.  Substances  which  are  necessary  for  the  Hg++  effect 
include  nonenzyme  proteins  and  certain  amino  acids,  of  w^hich  histidine  is 
the  most  effective  (Sauer  and  Rapoport,  1959).  It  was  concluded  that  in 
addition  to  an  SH  group,  activity  requires  the  presence  of  some  metal  ion 
and  some  complexer  bound  in  a  cylic  resonance  system: 


COO 

CHo — CH 

Resonance 
system 

NH, 

Enzyme 

Me  '^ 

^             cf 

with  the  enzyme.  In  this  case,  Hg++  would  simply  function  as  a  metal  ion 
for  linking  the  resonating  system  to  the  enzyme. 

(F)  The  mercurial  may  disrupt  water  structure.  The  structure  of  water 
around  the  active  center  may  be  such  as  to  retard  somewhat  the  access  of 


INHIBITION  OF  ENZYMES  819 

the  substrate.  Reaction  of  a  mercurial  with  a  vicinal  SH  group  could  by 
introducing  a  new  side  chain  break  down  this  water  structure. 

(G)  The  mercurial  reduces  the  binding  of  an  inhibitory  'product.  L-Gluta- 
mate  dehydrogenase  is  stimulated  by  PM  at  pH  8.5  and  this  is  reduced  by 
substrate,  NAD,  and  glutarate,  a  competitive  inhibitor  (Greville  and  Mild- 
van,  1962).  Thus  PM  must  combine  at  or  near  the  active  center.  The  prod- 
uct of  the  reaction,  a-ketoglutarate,  is  inhibitory.  PM  increases  the  K^  for 
a-ketoglutarate  8-fold  and  for  glutarate  more  than  20-fold.  Part  of  the  stim- 
ulation by  PM  can  be  due  to  reduction  of  the  effects  of  a-ketoglutarate. 
It  would  seem  that  such  behavior  would  be  reflected  in  the  rate  curves, 
little  stimulation  being  expected  initially. 

(H)  The  mercurial  dissociates  the  e^izyme  into  active  subunits.  Some  of  the 
active  centers  may  be  more  accessible  when  the  enzyme  is  disaggregated, 
and  it  is  known  that  mercurials  can  sometimes  split  enzymes  into  subunits 
(page  788).  GreviUe  and  Mildvan  (1962)  observed  that  PM  dissociates  glu- 
tamate  dehydrogenase,  and  Rogers  et  al.  (1962)  also  noted  effects  on  the 
sedimentation  properties.  The  possibility  of  such  dissociation  playing  a  role 
in  the  mercurial  activation  was  studied  by  Rogers  et  al.  (1963),  who  found 
no  change  in  molecular  weight  upon  treatment  with  MM  when  the  enzyme 
is  in  high  concentration.  However,  when  low  enzyme  concentrations  were 
used,  a  disaggregation  sensitive  to  the  mercurial  was  detected,  but  it  is  not 
certain  if  this  is  related  directly  to  the  stimulation. 

(I)  The  mercurial  reacts  primarily  with  an  inhibitory  SH  groups.  The  stim- 
ulation of  ATPase  by  mercurials  has  generally  been  explained  since  the 
report  of  Kielley  and  Bradley  (1956)  in  terms  of  differently  located  SH 
groups  around  the  active  center.  An  SH  group,  for  example,  might  bind  a 
group  on  ATP  and  interfere  with  the  optimal  orientation  on  the  enzyme. 
This  SH  group  has  been  postulated  to  react  with  the  6-amino  group  of  ATP; 
when  mercaptide  formation  occurs,  this  discouraging  action  on  ATP  is  abol- 
ished (Gilmour,  1960;  Greville  and  Tapley,  1960).  In  essence,  the  mercurial 
prevents  the  excessive  occupancy  of  the  active  center  by  disoriented  ATP, 
allowing  ATP  to  proceed  directly  to  hydrolysis.  A  somewhat  different  view 
has  been  voiced  by  Blum  (1960):  ATP  induces  a  configurational  change  in 
the  active  center,  this  involving  the  SH  groups,  and  mercurials  at  low  con- 
centrations tend  to  prevent  this  change.  ITP  does  not  so  alter  the  structure 
and  its  hydrolysis  is  inhibited  only  by  mercurials.  Mercurials  would  thus 
maintain  the  active  center  in  the  configuration  binding  ITP,  a  state  con- 
ducive to  rapid  hydrolysis  of  ATP. 

(J)  The  mercurial  inhibits  a  second  enzyme  which  suppresses  the  reaction 
rate.  A  number  of  possibilities  for  stimulation  were  discussed  in  Chapter 
1-7.  In  a  monolinear  chain: 

E,  Es 


820  7.  MERCURIALS 

inhibition  of  the  Eg  will  increase  the  steady-state  level  of  B  or  its  rate  of 
formation.  If  the  formation  of  B  is  being  measured  and  an  enzyme  destroy- 
ing it  is  present,  inhibition  of  this  enzyme  will  appear  to  stimulate  E^. 
Likewise  in  a  divergent  chain,  inhibition  of  one  branch  may  increase  the 
rate  of  the  other  branch;  whether  stimulation  will  be  observed  will  depend 
on  what  is  measured.  In  the  incorporation  of  nucleotides  into  amino  acid 
transfer  UNA,  the  presence  of  any  enzyme  attacking  the  nucleotides  will 
reduce  the  incorporation,  and  the  inhibition  of  such  an  enzyme  will  ap- 
pear to  stimulate  the  incorporation.  Stimulation  was  indeed  observed  with 
p-MB  by  Starr  and  Goldthwait  (1963),  who  thought  that  this  might  be 
due  to  a  contaminating  phosphodiesterase,  but  Anthony  et  al.  (1963)  ob- 
tained evidence  that  such  an  enzyme  did  not  occur  at  significant  levels  in 
their  preparation. 

The  stimulation  of  many  enzymes  by  mercurials  is  strongly  pH-depend- 
ent.  The  activating  effect  of  Hg++  on  glycerate-2,3-diphosphatase  becomes 
progressively  less  as  the  pH  is  increased  beyond  7  (Rapoport  et  al.,  1955), 
and  this  is  true  also  for  ascites  cell  ribonuclease,  although  here  the  activa- 
tion disappears  as  the  pH  is  decreased  to  5  (Colter  et  al.,  1961).  In  the  latter 
case,  a  50%  stimulation  changes  to  a  42%  inhibition  as  the  pH  rises  from  8 
to  8.5.  Liver  mitochondrial  ATPase  stimulation  by  p-MB  is  optimal  around 
a  pH  of  9  and  falls  off  rapidly  on  both  sides  (Myers  and  Slater,  1957  b), 
and  the  optimal  pH  for  activation  of  myosin  ATPase  is  7.8,  little  stimula- 
tion being  observed  at  pH  5.7  or  10  (Tonomura  and  Furuya,  1960).  Such  pH 
effects  may  be  important  in  working  out  the  mechanisms  of  the  stimulation 
but  have  not  so  far  been  studied  in  enough  detail  to  contribute  evidence. 
Temperature  can  apparently  also  play  a  role,  since  p-MB  stimulates  myosin 
ATPase  at  25°  but  only  inhibits  at  0^  (Fig.  7-30)  (Gilmour  and  Griffiths, 
1957).  It  is  also  evident  in  this  figure  that  the  DNP-activated  ATPase  is 
not  further  stimulated  by  p-MB.  The  EDTA-activated  ATPase  is  likewise 
not  stimulated  by  p-MB,  whereas  in  the  presence  of  Ca++  the  stimulation 
is  marked  (Fig.  7-23).  Finally,  the  effect  of  mercurial  concentration  is  oc- 
casionally very  striking,  as  in  the  case  of  myosin  ATPase  (Fig.  7-23),  max- 
imal stimulation  by  p-MB  being  observed  when  about  half  the  SH  groups 
are  reacted.  Another  example  of  variation  of  stimulation  with  mercurial 
concentration  is  shown  in  Fig.  III-l-l  for  ribonuclease,  although  here  the 
form  of  the  curve  may  depend  primarily  on  the  nature  and  amount  of 
the  natural  inhibitor,  as  well  as  the  susceptibility  of  the  enzyme  itself. 

The  problem  of  enzyme  stimulation  by  mercurials,  or  other  inhibitors, 
is  a  very  interesting  one,  and  probably  quite  important  in  understanding 
not  only  the  mechanisms  whereby  the  mercurials  can  affect  enzymes  but 
some  of  the  many  instances  of  stimulation  of  metabolism,  which  will  be 
mentioned  later  in  this  chapter. 


INHIBITION  OF  ENZYMES 


821 


Reversal  of  Mercurial  Inhibition 

The  general  treatment  of  the  reversal  of  enzyme  inhibition  by  substances 
binding  the  inhibitor  (page  1-615)  wiU  now  be  extended  to  those  situations 
often  encountered  in  studies  of  the  mercurials.  The  most  common  reversors 
are  thiols,  e.g.,  cysteine,  glutathione,  mercaptoethanol,  mercaptoacetate, 
and  dimercaprol  (BAL),  and  the  reduction  in  the  inhibition  can  be  consid- 
ered as  a  transfer  of  the  mercurial  from  enzyme  SH  groups  to  the  reversor 
(R)  SH  groups.  In  many  reactivation  experiments  the  concentration  of  free 
mercurial  is  very  small  and  the  reversal  reaction  can  be  written  as: 


EI        +        R        ±5        E 

(E,)  -  X       (Rt)  -  X  X 


+ 


RI 

X 


(7-13) 


If  the  enzyme  is  initially  completely  inhibited,  (EI)  =  (EJ,  and  the  final 
equilibrium  concentrations  are  as  written  under  the  equation.  The  equilib- 
rium is  characterized  by  the  following  constants: 


(EI)(R)  [(E,)  -x]  [(Rt)-x] 


(E)  (RI) 


(E)  (I) 
(EI) 


K, 


(R)  (I) 
(RI) 


K 


=   Kr 


(7-14) 


Fig.  7-30.  Effects  of  p-MB  on  myosin  ATPase 
at  0°  and  25°,  and  in  the  presence  of  2,4-di- 
nitrophenol.  The  ATPase  activity  is  given 
as  /^moles  P,/mg/min.  (From  Gilmour  and 
Griffiths,   1957.) 


822  7.  MERCURIALS 

The  value  of  x,  the  concentration  of  active  enzyme  regenerated,  can  be 
calculated  from  the  quadratic  equation: 

x^{\  -  K)-x  [(R,)  +  (E,)]    +   (E,)  (R,)  =  0  (7-15) 

if  the  total  concentrations  of  enzyme  and  reactor  are  known.  The  value  of 
the  final  inhibition  is  given  simply  by: 

If  both  enzyme  and  reversor  SH  groups  ionize  similarly,  we  need  not  in- 
clude the  effect  of  (H+),  and  since  the  mercurial  is  assumed  to  be  bound 
to  either  the  enzyme  or  the  reversor,  we  need  not  consider  (X~),  the  con- 
centration of  mercurial-complexing  ligands  in  the  medium.  The  situation 
described  here  is  that  in  which  an  enzyme  is  titrated  to  complete  inhibition 
and  the  reversor  then  added.  If  any  irreversible  inactivation  has  occurred, 
the  experimentally  measured  inhibition  will  be  greater  than  in  Eq.  7-16. 
In  most  reactivation  experiments,  (R^)  is  much  greater  than  (E,),  and 
thus  the  reversor  concentration  is  not  reduced  significantly  by  the  binding 
of  the  mercurial.  The  equation  for  the  determination  of  x  is  simplified  to: 

a;=A'  +  a;(R,)  -  (E,)  (R,)  =  0  (7-17) 

It  is  interesting  to  specify  the  conditions  required  to  reverse  the  inhibition 
to  a  determined  value  ij.  Substitution  of  (E^)  (1  —  ij)  for  x  in  the  solution 
to  the  quadratic  equation  leads  to 

(7-18) 

(7-19) 
(E,)  V  K, 

Equation  7-18  gives  the  ratio  of  the  dissociation  constants  of  the  two  mer- 
captides  so  that  the  inhibition  will  be  reduced  to  ij  when  the  enzyme  and 
reversor  concentrations  are  given,  while  Eq.  7-19  shows  how  much  reversor 
must  be  present  relative  to  the  enzyme  to  achieve  a  reactivation  to  if  when 
the  dissociation  constants  are  known.  In  many  experiments  the  reversor  is 
added  at  a  concentration  of  1-10  vaM  and  thus  (R^)/(EJ  is  often  10^  and 
10^,  since  enzyme  concentrations  commonly  run  from  10"^  to  10^^  raM. 
If  we  assume  that  K^  =  K^,  which  is  a  reasonable  approximation  in  many 
cases,  the  values  of  (R^)/(E^)  shown  in  the  following  tabulation  must  be 
used  to  reduce  the  inhibition  to  the  levels  indicated.  Thus  one  would  expect 
that  the  concentrations  of  reversor  often  used  would  completely  abolish 
the  inhibition,  so  that  if  the  enzyme  activity  is  experimentally  not  restored 


Kr 

*/ 

(Rt) 

K, 

(1  —  ifT- 

(E,) 

(Re) 

(1  -  if)' 

Kr 

INHIBITION  OF  ENZYMES  823 

to  normal  values  it  is  likely  that  (1)  some  inactivation  has  occurred,  (2)  in- 
sufficient time  has  been  provided  for  the  reversal,  (3)  the  binding  of  the 
mercurial  to  the  enzyme  is  much  tighter  than  to  the  reversor  (K^"^  K^), 


Final  %  inhibition 

(Ri)/(E,) 

50 

0.5 

20 

3.2 

10 

8.1 

5 

18 

1 

98 

or  (4)  some  secondary  factor  has  complicated  the  situation,  e.g.,  failure  to 
add  a  displaced  cofactor  (page  1-625)  or  the  oxidation  of  enzyme  SH  groups 
by  the  oxidized  thiol  reversor  (page  1-625).  If  the  complete  reversibility 
of  the  inhibition  has  been  established,  and  if  (E^)  and  K^  are  known,  it  is 
possible  to  calculate  the  value  of  K^  from  experiments  in  which  low  con- 
centrations of  reversor  are  added  and  the  ij  determined. 

It  is  relatively  easy  to  treat  equilibrium  conditions  quantitatively,  but 
the  problems  encountered  in  considering  the  rates  of  reversal  are  at  the 
present  mainly  insoluble.  Investigators  have  been  chiefly  interested  in  whe- 
ther reversal  occurs  or  not  and  since  extremely  few  rate  studies  have  been 
made,  there  are  not  adequate  data  upon  which  to  base  accurate  analyses. 
If  the  reversal  occurs  in  two  steps  —  the  dissociation  of  EI  and  the  combi- 
nation of  I  with  R  —  the  individual  reactions  are  not  as  simple  as  with  most 
inhibitions.  For  example,  at  physiological  pH  and  with  most  media  used, 
the  dissociation  of  EI  must  be  written  as: 

EI  +  H+  +  X-  -►  EH  +  IX  (7-20) 

where  X~  represents  any  complexing  anion  in  the  medium.  Likewise,  the 
second  step  must  usually  be  described  by: 

IX  +  RH  ->  RI  +  X-  +  H+  (7-21) 

Now  it  is  not  known  if  the  reactions  occur  in  this  manner,  or  whether  I 
can  be  directly  transferred  from  E  to  R  without  forming  the  IX  complex, 
either  on  the  surface  of  the  enzyme  or  in  solution.  A  study  of  the  effects 
of  (X^)  on  the  reversal  rate  would  help  to  solve  this  problem.  It  is  also 
not  known  if  the  reactive  form  of  the  reversor  is  RH  or  R~,  but  from  the 
effects  of  pH  on  the  rates  of  reaction  of  mercurials  with  simple  thiols  it 
appears  that  at  least  the  form  R"  reacts  much  more  rapidly  than  RH,  as 
would  be  anticipated.  In  any  event,  the  pH  must  be  an  important  factor 


824  7.  MERCURIALS 

in  the  rates  of  reversal,  although  it  may  not  markedly  influence  the  equi- 
librium conditions. 

Despite  the  lack  of  data  on  reversal  rates,  one  obtains  the  impression 
that  reversal  is  often  a  good  deal  faster  than  the  development  of  inhibition. 
Unfortunately  those  workers  who  measured  the  rates  of  inhibition  the  most 
carefully  usually  did  not  examine  the  reversibility  or,  if  they  did,  remarked 
only  that  it  occurs,  without  giving  data  on  the  time  required;  and  in  other 
cases  no  reversal  was  observed.  We  are  thus  in  the  surprising  position  of 
having  essentially  no  information  in  any  one  case  of  mercurial  inhibition 
as  to  the  relative  rates  of  inhibition  and  reversal.  Certainly  in  some  instances 
the  reversal  is  very  rapid,  as  with  3-phosphoglyceraldehyde  dehydrogenase 
(Velick,  1953)  and  hexokinase  (Sols  and  Crane,  1954)  treated  with  cysteine 
after  inhibition  by  p-MB,  as  previously  discussed  (page  1-623  and  Fig.  I- 
13-8).  However,  in  these  cases  the  rates  of  inhibition  are  not  known,  al- 
though they  are  probably  quite  fast;  in  both,  the  enzyme  and  p-MB  were 
preincubated  for  an  arbitrary  time  (for  hexokinase  15  min  at  0°).  It  would 
be  difficult  to  compare  the  rates  of  reversal  by  dialysis  and  reversor  —  it 
seems  never  to  have  been  done  experimentally  —  but  dialysis  is  not  a  very 
efficient  method  due  to  the  fact  that  only  a  small  fraction  of  the  dissociated 
inhibitor  at  any  time  is  able  to  pass  out  through  the  membrane,  most  re- 
combining  with  enzyme,  while  the  presence  of  a  high  concentration  of  re- 
versor throughout  the  medium  ensures  that  free  inhibitor  is  rapidly  bound. 
Therefore,  one  does  not  know  in  reversal  which  step  is  limiting,  or  if  the 
reversor  can  facilitate  the  dissociation  of  the  mercurial  from  the  enzyme. 
It  is  impossible  to  predict  the  relative  rates  of  inhibition  and  reversal  theo- 
retically because  we  do  not  know  the  exact  reactions  involved,  particularly 
the  influence  of  H+  and  complexing  ions.  There  is  only  one  way  to  solve 
these  problems:  to  do  a  few  critical  and  accurate  experiments. 

The  information  to  be  derived  from  simple  reactivation  experiments, 
especially  those  in  which  a  high  concentration  of  reversor  is  used,  is  not 
as  much  or  as  reliable  as  many  seem  to  have  believed  (see  page  1-624).  If 
complete  reversibility  is  obtained,  this  shows  that  no  serious  inactivation 
of  the  enzyme  has  occurred;  it  does  not  imply  that  structural  changes  have 
not  been  induced  by  the  mercurial.  This  information  is  usually  important 
and  must  be  obtained  in  any  quantitative  kinetic  studies,  but  this  is  all 
that  this  type  of  reversal  experiment  will  provide.  Failure  to  recover  the 
activity  can  be  explained  in  a  variety  of  ways,  previously  enumerated  (page 
651).  More  careful  reversal  studies,  especially  determination  of  rates,  the 
effects  of  pH  and  ligand  concentration,  and  the  degree  of  reactivation  with 
low  reversor  concentrations,  would  undoubtedly  be  more  informative. 

Several  hundred  studies  on  mercurial  inhibition  have  included  statements 
relative  to  reversibility  with  thiols.  It  would  serve  very  little  useful  pur- 
pose to  list  these  results.  Summarizing  all  of  them  one  finds  that  complete 


INHIBITION  OF  ENZYMES  825 

reversal  was  obtained  in  59%,  partial  reversal  in  30%,  and  no  reversal  in 
11%.  This  shows  at  least  that  in  most  cases  the  enzyme  is  not  seriously 
altered  or  denatured  by  severe  mercurial  inhibition,  providing  the  contact 
with  the  inhibitor  is  not  prolonged.  On  the  other  hand,  there  are  the  en- 
zymes such  as  papain  which  are  more  stable  when  complexed  with  Hg++. 
Some  of  the  instances  where  partial  or  no  reversal  was  found  are  undoubted- 
ly due  to  inactivation,  but  some  to  the  other  factors  mentioned  above.  In  a 
few  studies,  some  interesting  sidelights  were  noted  and  we  shall  content 
ourselves  with  discussing  some  of  these. 

Two  more  instances  of  very  rapid,  almost  instantaneous  reversal  of  mer- 
curial inhibition  may  be  noted.  Yeast  alcohol  dehydrogenase  is  immediately 
inhibited  by  p-MB  and  likewise  rapidly  reactivated  by  glutathione  (Snod- 
grass  et  al.,  1960).  The  enzyme  at  4.5  X  10"^  M  is  completely  inhibited  by 
2.5  X  10"''  M  Hg++;  glutathione  reverses  this  inhibition  rapidly,  but  par- 
tially —  after  30-sec  exposure  to  Hg++  54%  and  after  10-min  exposure 
only  17%.  This  failure  to  reverse  is  not  due  to  displaced  Zn++,  since  Zn++ 
was  found  not  to  be  released  from  the  enzyme  so  rapidly  and  the  addition 
of  Zn++  does  not  improve  the  reversal.  Heart  lactate  dehydrogenase  is  very 
rapidly  reactivated  from  p-MB  inhibition  by  cysteine,  and  in  this  case  the 
rate  of  inhibition  is  relatively  slow,  about  15  min  being  required  for  maxi- 
mal inhibition  (Neilands,  1954).  This  is  one  of  the  few  instances  in  which 
reversal  definitely  seems  to  occur  faster  than  inhibition.  The  effectiveness 
of  reversors  in  comparison  with  other  methods  for  reactivation  is  seen  with 
succinate  dehydrogenase.  If  the  enzyme  inhibited  by  p-MB  is  dialyzed  for 
3.5  hr  there  is  no  reactivation,  but  addition  of  glutathione  reverses  complete- 
ly (Singer  et  al.,  1956b).  Also  no  reversal  was  observed  following  dilution  of 
the  inhibited  enzyme,  whereas  thiols  reactivate  partially  (Slater,  1949).  The 
rate  of  reactivation  of  erythrocyte  pyruvate  kinase  by  100  mM  glutathione 
after  1-min  contact  of  the  enzyme  with  0.1  mM  p-MB  is  fairly  slow  (see 
accompanying  tabulation);  however,  if  contact  with  the  inhibitor  is  10  min, 


Time 
(min) 


%  Reversal 


10  18 

15  47 

25  59 


only  28%  reversal  is  seen  after  25-min  exposure  to  glutathione  (Solvonuk 
and  CoUier,  1955).  There  is  no  obvious  explanation  why  the  reversal  rates 
with  some  enzymes  are  so  fast  and  in  others  so  slow,  especially  as  this  is 
not  correlated  at  all  with  the  aifinities  of  the  enzymes  for  the  mercurials. 


826  7.  MERCURIALS 

Low  concentrations  of  reversor  in  a  definite  molar  ratio  to  enzyme  or 
mercurial  have  seldom  been  used,  but  in  the  examples  we  have  the  reversal 
is  only  partial.  Succinate  oxidase  preparation  from  pigeon  muscle  is  inhi- 
bited 88%  after  15-min  incubation  with  0.03  mM  Hg++;  reversal  by  di- 
mercaprol  for  30  min  depends  on  the  amount  of  the  dithiol  added  (see  ac- 
companying tabulation)  (Barron  and  Kalnitsky,  1947).  Myocardial  succin- 


Dimercaprol :  Hg++  %  Reversal 

3.3  0 

5  33 

10  63 


ate  oxidase  inhibited  completely  by  p-MB  can  be  31%  reactivated  by  glu- 
tathione at  a  molar  ratio  to  the  mercurial  of  1  :  1  and  65%  at  a  ratio  of  5  :  1 
(Slater,  1949).  Treatment  of  3-phosphoglyceraldehyde  dehydrogenase  with 
33-MB  leads  to  changes  in  the  optical  rotation;  if  the  exposure  to  the  mer- 
curial is  only  1  min,  cysteine  at  a  molar  ratio  of  6  :  1  reverses  these  changes 
around  75%,  but  the  structural  changes  become  progressively  more  irre- 
versible (Elodi,  1960).  To  reverse  the  inhibition  of  glutamate  dehydrogen- 
ase by  p-MB  maximally  (80%)  it  requires  around  60  times  as  much  glu- 
tathione as  mercurial*  (Olson  and  Anfinsen,  1953). 

*  These  experiments  were  done  by  varying  the  p-MB  concentration  from  0.003 
to  3.3  mM  and  keeping  the  reversor,  glutathione,  at  10  mM,  so  that  the  more  normal 
conclusion  is  simply  that  the  reversibility  is  less,  the  higher  the  mercurial  concentration, 
a  phenomenon  commonly  observed  with  other  enzymes.  The  plotting  is  ambiguous; 
the  final  enzyme  activity  is  plotted  as  %  of  the  uninhibited  enzyme,  but  the  results 
with  inhibitor  alone  are  not  given,  although  they  can  be  estimated  from  another  figure, 
so  the  degree  of  reversal  is  not  immediately  apparent.  For  example,  when  p-MB  is 
0.11  mM  the  inhibition  is  given  elsewhere  as  50%;  after  10  mM  glutathione  (molar 
ratio  90  :  1)  the  enzyme  activity  is  given  as  approaching  80%,  which  I  would  call 
60%  reversal  on  one  basis  or  30%  on  another.  However,  this  is  stated  to  be  70% 
reversal  in  the  text.  No  incubation  times  were  mentioned  so  possibly  the  failure  to 
achieve  more  reversal  is  due  to  an  insufficient  time  with  the  reversor.  Other  than  the 
work  on  reversal,  this  investigation  is  an  excellent,  detailed,  and  quantitative  study 
on  an  enzyme,  and  thus  illustrates  a  rather  common  phenomenon  —  the  lackadaisical 
and  disoriented  approach  to  inhibition  reversal.  In  90%  of  the  reports  in  which  re- 
versal is  determined,  apparently  an  arbitrary  (but  high)  concentration  of  reversor  is 
added,  and  the  enzyme  activity  is  measured  after  an  arbitrary  interval,  so  the  conclu- 
sions should  usually  be  taken  as  arbitrary.  If  these  remarks,  and  others  scattered 
throughout  the  book,  can  stimulate  the  performance  of  more  accurate  and  interpretable 
reversal  experiments,  my  aim  will  be  achieved. 


INHIBITION  OF  ENZYMES  827 

Unusual  results  on  the  effects  of  thiols  on  the  inhibition  of  /?-fructofu- 
ranosidase  by  Hg++  were  observed  by  Gemmill  and  Bowman  (1960),  in 
that  dimercaprol  reverses  the  inhibition  to  varying  degrees,  whereas  both 
cysteine  and  glutathione  increase  the  inhibition  (Table  7-12).  The  thiols 
themselves,  even  at  the  highest  concentrations  used,  do  not  significantly 
affect  the  enzyme  activity,  so  the  additional  inhibition  cannot  be  attribut- 
ed to  an  excess  of  the  thiol.  Several  explanations  are  possible:  (1)  The  cys- 
teine and  glutathione  reduce  enzyme  disulfide  groups  to  SH  groups  and 
increase  the  binding  of  the  Hg++  to  the  enzyme;  (2)  the  Hg++  reacts  with 
both  enzyme  SH  groups  and  these  thiols  to  form  E — S — Hg — S — R  com- 
plexes which  are  less  active  than  the  simple  E — S — Hg+  complexes;  or  (3) 
the  R — S — HgX  or  R — S — Hg — S — R  mercaptides  formed  are  inhibitory 
by  a  mechanism  unrelated  to  enzyme  SH  groups.  Whatever  the  explanation 
applicable  here,  one  would  expect  that  these  situations  would  occasionally 
be  important  in  the  reversal  of  other  enzyme  inhibitions,  particularly  when 
the  bifunctional  Hg++  is  used,  and  possibly  failure  to  achieve  reversal  in 
some  cases  may  be  due  to  these  reactions. 

Complexities  arise  in  some  reversal  studies  and  a  few  will  be  mentioned 
briefly.  When  the  aspartate:  or-ketoglutarate  transaminase  from  pig  heart 
is  treated  with  p-MB  there  is  inactivation  but  the  pyridoxal-P  remains 
bound  to  the  apoenzyme  (Turano  et  al.,  1964).  Incubation  for  10  hr  at  pH 
6.4  and  4°  with  0.35  mM  p-MB,  however,  releases  the  coenzyme.  The  ad- 
dition of  pyridoxal-P  restores  the  activity  in  the  sense  that  addition  of 
glutamate  is  followed  by  transamination  to  form  pyridoxamine-P,  but  the 
further  transamination  to  oxalacetate  is  lost.  In  this  instance,  reversal  thus 
occurs  for  only  part  of  the  normal  reaction.  Carboxy peptidase  A  is  activat- 
ed by  Zn++  and  apparently  the  Zn++  is  bound  to  the  SH  groups  of  a  single 
cysteine  residue  and  the  a-amino  group  of  a  terminal  asparagine  (Coombs 
et  al.,  1964).  Zn++  protects  the  enzyme  from  inhibition  by  p-MB,  but  if  the 
apocarboxypeptidase  is  inactivated  by  p-MB,  addition  of  Zn++  does  not 
restore  the  activity. 

Spontaneous  reversal  of  enzyme  inhibition  by  p-MB  occurs  with  3-phos- 
phoglyceraldehyde  dehydrogenase  of  muscle  (Szabolcsi  et  al.,  1960),  and 
has  occasionally  been  observed  in  other  systems,  particularly  homogenates. 
This  has  usually  been  attributed  to  redistribution  of  the  mercurial  from  the 
rapidly  reacting  SH  groups  to  others  which  bind  the  mercurials  more  tight- 
ly, but  in  this  case  it  was  thought  that  intermolecular  rearrangements, 
whereby  eventually  some  of  the  enzyme  is  inactive  and  some  completely 
active,  are  responsible.  This  should  not  be  too  uncommon  a  phenomenon 
if  the  proper  mercurial  concentrations  are  used. 


828 


7.  MERCURIALS 


(M        cci_ 


2 


O 


JS 

-tj 

-u 

'% 

+ 

+ 

fi 

+ 

_o 

^ 

'■S 

W 

,a 

2 

s 

ffi 


^ 


o 

+ 


W 


W 


kd 

^ 

p< 

c 

cS 

HH 

o 

h 

\0 

(B 

0-- 

a 

5 

+ 


w 


o3 

^-v 

03 

d 

« 

iCi 

03 

03 

eS 

o 

1:5 

c 

cS 

CS 

^ 

s 

•3 

^ 

^ 

O 

s 

73 

ffO 

C 

f-t 

cS 

o 

o 

^Zi 

o 

's 

d 

a 

OJ 

OJ 

cS 

o 

? 

+ 

s 

+ 

o 

^ 

c3    a> 


INHIBITION  OF  ENZYMES  829 

Survey  of  Enzyme  Inhibitions 

Some  enzyme  inhibitions  produced  by  the  mercurials  are  given  in  Table 
7-13.  These  represent  only  about  a  fifth  of  the  enzymes  studied.  It  is  a 
difficult  matter  to  decide  which  results  should  be  in  the  table.  I  have  tried  to 
include  principally  two  sorts  of  enzyme,  those  that  are  "important"  and 
those  that  are  inhibited  "potently."  But  which  enzymes  are  important? 
Every  enzyme  is  important  for  a  particular  pathway,  or  a  certain  organism, 
or  the  investigator  who  studies  it.  The  following  groups  of  enzymes  have 
generally  been  chosen:  those  in  the  glycolytic  Embden-Meyerhof  or  pen- 
tose-? pathways,  the  tricarboxylate  cycle,  electron  transport  systems,  phos- 
phate transfer,  and  central  amino  acid  metabolism;  those  often  involved 
in  cell  function  (such  as  choline  acetylase,  cholinesterase,  carbonic  anhy- 
drase,  etc.);  and  certain  classic  SH  enzymes  (such  as  urease).  We  now  must 
consider  how  the  word  "potently"  should  be  defined.  I  have  arbitrarily 
chosen  enzymes  inhibited  significantly  (usually  50%  or  more)  by  concen- 
trations of  mercurial  of  0.01  mM  or  less,  since  such  enzymes  could  usually 
be  considered  as  having  reactive  SH  groups  at  or  near  the  active  center. 
It  is  evident  that  a  certain  enzyme  differs  often  very  markedly  in  suscepti- 
bility depending  on  the  tissue  or  species  from  which  it  is  obtained,  so  that 
one  cannot  speak  of  the  sensitivity  of  NADH  oxidase,  for  example,  in  quan- 
titative terms  without  specifying  which  NADH  oxidase  is  meant.  At  least 
most  will  agree  I  think  that  the  enzymes  included  in  Table  7-13  are  impor- 
tant in  common  metabolic  pathways  and/or  are  inhibited  quite  potently.* 

Another  problem  in  presenting  the  inhibitions  in  Table  7-13  is  that  the 
degree  of  inhibition  by  mercurials  depends  on  a  number  of  factors  to  a 
greater  extent  than  with  most  other  inhibitors.  Particularly  important  are 
the  pH  (mainly  because  of  competition  of  H+  with  the  mercurial  for  the 
S~  group),  the  temperature  (because  of  both  the  high  temperature  coeffi- 
cients of  mercurial  reactions  and  the  possibility  of  secondary  thermal  inac- 
tivation),  the  composition  of  the  medium  (principally  because  of  competi- 
tion of  anions  with  the  S~  group  for  the  mercurials),  the  period  of  the  in- 
cubation with  the  inhibitor  (since  in  many  cases  the  inhibition  does  not 
attain  a  constant  level),  and  the  presence  of  impurities  (most  of  which  can 
complex  with  the  mercurials  and  reduce  their  effectiveness).  There  are  too 
many  of  these  variables  to  include  in  the  table.  The  purposes  of  the  table 
are  to  provide  very  roughly  some  information  on  the  relative  sensitivities 
of  the  more  important  enzymes,  and  to  present  an  experimental  basis  for 
the  appreciation  of  the  lack  of  specificity  of  the  mercurials  when  used  in 
complex  cellular  preparations. 

In  addition  to  these  problems,  it  is  likely  that  many  of  the  most  notable 

*  If  a  reader  wishes  to  obtain  information  on  the  inhibition  of  an  enzyme  not  given 
in  the  table,  I  shall  be  happy  to  try  to  supply  this  upon  receiving  a  written  request. 


830 


7.  MERCURIALS 


P5 


ft 

a 

k4 

c3 

05 

3 

s 

':3 

CO 

c3 

TS 

fcH 

c 

-d 

cS 

eS 

c 

JsJ 

H 


s 

>, 

G 


cc; 


03 


W 


O    O    O    <M 


^ 


o 

CO 

o 

o 

o 

o 

■*    00    o 

»c 

o 

o 

o 

o 

o 

O      O      -H 

o 

to  1^ 


+  m 


m 


o  »o 
d  d 


&q 


fei  iaq  a. 


o 

1 

Si 

^ 

" 

=0 

CO 

to 

CO 

3 

O 
g 

u 
g 

'■*3 

C 

a 

o 

o 

Si 

o 

fl 

0, 

05 

-1-2 

8 

INHIBITION  OF  ENZYMES  831 


w  - 


m 


3        x;   ::^   b 
w        ^  ^  ^ 


e    .2 


-iii 


rt 

C5 

(-< 

c 

« 

i^ 

^ 

tx 

3 

O 

< 

H 

O 

O     ■<*    GO     o   o     o 
O     ■*    00     "O    o     o 


OCDOO'-i  ,_irtO^ 

O         0(M         OOO         OOO-HOO  rtOOOO 


O       OO       OOO       oooooo       ^oooooooooo 


+  ++pqeLi        +m  -i-pqfiHWeqcq        p5         +pqpq 

beg  bC&cSS  "m^  bD^Sf^i^S  S  M^^ 


t3 

-£3 


o 

>, 

>> 

s 

c 

c4 

0) 

a 

o 

■a 

-o 

■-d 

< 

<! 

< 

<3 

832 


7.  MERCURIALS 


<1         % 


T3 

? 

IS 

O 

C 

ce 

T1 

S 

n7i 

>> 

c 

W      fcii 


ft     ? 
S    M     ffi 


S     - 


ft   j2 


-5  a  s 

g  eg  -- 

53  O  ft 

PP  PL|  CC 


O    f-H    O    (M 


lO    iC    O        O 

oo  CO   00       o 


00     CO    OS    l>  O    -H    (M 

O     C0-*CDOO     OOOi— I 
O     OOOO     OOOOr-< 


ooooooooooooooooooo      ooo 


pq 


M 


be 


PQ   W   m   pq 


a      ?5,      a      ?», 


pq  pq 


a      ?5< 


pq 


pq  pq  + 
^  S   t. 


M 


s 

i 


rO      S 


'^^      >^ 


cS 

ft 

> 

n 

PM 

CS 

« 

ki 

+ 

< 

> 

^ 

^ 

^^ 

^-' 

4J 

^ 

^ 

ert 

ai 

^ 

4) 

0) 

n) 

>l 

^ 

« 

^ 


^ 


INHIBITION  OF  ENZYMES  833 


,-^ 

—' 

lO 

o 

. 

o 

Tf* 

i-H 

C5 

>-, 

^^ 

^ 

o 

.2 

5? 

T) 

o 

C 

O 

O 

T) 

o3 

0) 

Xi 

C 

d 

ft 

o 

s 

o 

c 

c3 

bfl 

C 

C 

efl 

c 

^ 

"5     -^    05  _  C5 


CO        r- T    o    '^ 
^    c  ^^ 


---  —     -"--^CJOfOOOS-^ 


lO 

>i5                                           lO 

1— 1 

O    1— 1 

-H       O 

1— 1 

O     ^     i-H                                       O      -H      IC 

2  -' 

o  o 

C^l    fO 

f-H      i-H 

o 

O    O    O                — 1   o    o    o 

2  ==> 

o  o 

^    O    O    -1 

o  o 

o  o 

O      -H 

OOOtN— lOOOO 

o  o 

o  o 

OOOO        OO        OO         Oo-hO        --hOOOOOOOOO        OO 


^  ^  js       s~  rs  o  « 

o     o     o        ^  ^  **  O 

IB     m     03        S  .o  C  if 

tS  c            CO  =0  i2                              P< 


w       tv       '.J  N-i  yj  ^  »i3  '(S-  K^  CO       C3       s  •*.*  fc-l 


S  •«         .S 

ft  ®  ^  T3  ^ 


834 


7.  MERCURIALS 


P5 


H 


H 


U  .-H 


w 


o 


o 

t3 

O 

o 

T3 

I 

c« 

t3 

^ 

S 

o 

cS 

aj 

■ft 

t^ 

3 

3 

M 

C» 

o  o 


o 

O     r-^ 

o  o 


O      -H      P-< 

odd 


<N 

ec 

»C    CD 

<N 

TfH 

(N 

2  o 

O 

o 

(N 

--H   o  o  o 

<6  (6  S  <6 


+  m      + 


m 


PQ 


W 


+  m      +  M 


(h 

M 

ft 

c 

cS 

cS 

« 

S 

c 

p 

s 

w 

O 

CQ 


Pm 


W 


c3 

^ 

;-i 

crt 

o 

PM 

cS 

w 

-t^ 

<; 

=3      ft 


C     S 


Q 


INHIBITION  OF  ENZYMES  835 


B 

T3 

C 

<u 

c3 

a 

C 

Xi 

cS 

JS 

U 

O 

a 

'S 

CO     05 

o 

Tt       ^ 

Oi    ^ 

O 

^-^   *«^ 

T3 

.      B 

C 

e   ~ 

cS 

OC0C5  -H  ICCOCO  ■*'M  cot^ 

oo      oooocoooooooo      o      oooo      oecicooooco 


pq      +  +  PQ     w  m  + 

Sm  Mgg  S§  Sao 


^       -  ""sag 


6    6  6 


836 


7.  MERCURIALS 


P5 


^:2 

a 


^H 

C5 

05 

05 

CO 

>— V 

»c 

1— 1 

(—1 

10 

■* 

05 



^-^ 

■ 

05 

?D 

i-H 

Oi 

(-1 

^^ 

^ 

5 

c8 

M 

kc 

73 

S" 

IM 

C<l 

0 

® 
T3 

s 

43 

cS 

01 

0 
0 

c 

t3 

§ 

g 

10 

OS 

43 
0 

2 

2 

M 

a 

:^ 

'e 

"e 

p 

e 

c3 

c 

o3 

10 
Oi 

0) 

'S 

■« 

C 

_g 

k> 

tH 

^ 

,0 

03 

2 

'S 
2 

3 

c 
3 

is 

(4 

(4 
<» 

be 

03 
"ft 

3 

0 

> 

0 

0 

s 

"o 

0 

0 

cS 

be 

ft 

1 

P5 

« 

0 

S 

S 

§ 

H 

<1 

0 

0 

0   i>   CO 

0 

0000 

t-    lO 

^ 

t^ 

IC 

IC 

-* 

t^ 

»o 

IC   ic   10 

(M    05 

0 

■— ' 

P5 


'« 

TJ 

C 

V 

cS 

>> 

^ 

x> 

3 

H 

W 

W 

^ 


(N     O     --H    ^ 


-* 

!0 

-H    »0    10    CO 

S<)    ■<# 

0 

0    "1 

0      ^      -H                 0000 

0 

0    (N    ^    «D    0 

0    0 

0 

ooo^o   000 

+    +    +0^+    +fq+WPQ 
bcboboi^uibjoi^bci^i^ 


pq 


m     + 

S  bO 


M 


m  pq 


a 

a 

a 

a 

0 

2 

1 

0 

e 
s 

1 

B 

3 
b 

3 

g 

^ 

^ 

>> 

0 

lU 

4) 

OJ 

aj 

(H 

Ih 

4> 

tn 

m 

03 

03 

® 

4) 

t 

0 

« 

00 

C 

C 

C 

C 

c 

c 

s 

+j 

a 

c8 

a 

a 

c3 

a 

a 

03 

a 

j3 

g 

2 

0 

3 

3 

3 

3 

3 

3 

bo 

^ 

03 

w 

w 

w 

K 

w 

w 

w 

S 

Qs 

tf 

p?  s 


03 

>, 

-2 

§ 

j3 

b£ 

71 

C 

0) 

o3 

o 


o     o 


INHIBITION  OF  ENZYMES 


837 


-—  (M 


o 


c 

^  C!.  oi 

t3 

T3 
C 

O 

V     O    '-^ 

« 

o 

^    S    o 

Is 

to 
O 

m 

S  c»  W 

S 

P5 

m 


o     ^ 


CD 

a> 

JD 

05 

T) 

>— < 

c 

— 

2 

j: 

> 

CO 

03 

M 

_C 

C/3 

"^ 
^ 

73 

S 

TJ 

cS 

fi 

73 

Ui 

T3 

cS 

C 

-2 

-is! 

03 

3 
W 

fl 

o 

^ 

fi 

£ 

^ 

o 

> 

OOO        (MO       fCiOO 
lO       ^    O       -^    CO    o 


O    O    O    lO       o 
05    »C    »C    -H       lO 


(M    CO 

o  o 


OOO      ooo      oo 


1— <      »o 
o     o 


OOOOO       MOOO 


»C     ■<#     05 

O    .-H    iM 

<M    rt       OOO 

O    O       OOO 


m 


eq  P5  M 

?!,     a     ?!. 


m 


m 


w 


+  P3  PQ  m    p2 

bc  S   ^   ^      § 

W      i,     a     4         ?!h 


a  (In 


w 


M 


ft    o 


5    s    <^      .2 

J  J  ^     -S 
a,  '^  ^     o 


fll 


m 

3 

s 

S 

S 

■t^ 

■b;> 

^ 

-O 

-o 

,ri 

c« 

c6 

« 

P5 

m 


03 

O 

S 

■^ 

"cS 

o 

Ph 

« 

<» 

03 

T3 
4) 

S 
o 

A 
o 
o 

'-3 
c;i 

CO 

d 

c4 

o 

-o 

-tJ 

O 

0^ 

o 

ft 

ft 

73 

s 

ft 


03        73 


fe 


838 


7.  MERCURIALS 


P5 


■^  =! 


ft 

a. 

o 

eo 

-1-3 

«o 

CO 

Oi 

l-H 

t3 

•^-^ 

s 

bo 

eS 

m 

V 

t*^ 

^ 

a 

3 

o  ^ 


O 


S  O  O 


O 


O      -H 

O      O      -H 

o   o   o   ^ 

O    O    O    O    — 1    Tt< 
OOOOOO-HiM 


oooooooo 


2j  «o 

O    O 

d  d 


o  o 

o  o 

-H  O     O 


CO      O   -^   o  o   o 


W       ffi 


w 


A    w 


eq  m  M 

§  s  ^ 

a  ?i,  4 


>> 

CO 

T3 

'a 

> 

« 

^  ^ 

— ' 

-tj 

>"     cS 

c3 
P5 

^    ce 


03 

cc 

i     6fi 

q 

-S     O 

03 

ctf 

c3 

o    ^ 

.3  -S 

3^ 

INHIBITION  OF  ENZYMES  839 


s      «^ 


CO 
05 


o- — ■hJ'^S        ^  c 


1— iOG0<^^dOc^  I— |i— lO  OtJ<CD         (MOI>XOi 


OOCOO  O     —     C^JO-H  OOOOOOOO-H 

00  MO  rtio-nOOO'OO  Oo'^^'-^'M  OOO  OOOOO 


PQ  pqpq+        pci«pqaaP3mP3        +    +    ^      +   m  m 


to        ^      ,     ^    S    ^    ^ 


+ 

+      ^ 

+ 

pq 

+ 

M 

M  S 

M 

S 

ffi 

W   4 

s 

?:. 

3  -s    >       § 


■g      o    S      * 


000  O 


o  o 


840 


7.  MERCURIALS 


^-  CO 


^     -5  C5 


1^ 


05 


o 


M  O 

CB  O 


00  c<l  ?D  >C  "O 
5^  CO  lO  ^  l> 


O  IC  O  «<l  O  CO 
•*  Oi  O     IC  03 


O  -H       O  ,-H  lO  O 

O  O  -H  O  o  o  o 


CO^OIOIOOOO 


OOO      OoOOOOOOO^-^-^<MOO(MOOOO 


P3 


m 


m 


Ul 

fq 

m 

Ph 

P5 

+ 

S 

S 

§ 

g 

M 

?i, 

?i, 

A 

a 

a 

■*     s     « 


H 

ft 

^ 

Cl 

0) 

<n 

Tl 

-JC 

0) 

•  1 

Ph 

!S 

> 

O 

tf 

0) 

W 

<: 

O 

r 

21 

S 

1*. 

lA, 

O 

i^  Q  ;=  P 


Q       "5 

;3    u_    <3 

PlH     PP     ^ 


o 


rt 

<B 

fl 

v 

ctf 

bfi 

hi 

0 

S 

■J3 

0 

« 

INHIBITION  OF  ENZYMES 


841 


o 


a 
o 

a 

T3 


O 


CQ 


05010       fOiO       05C5       OiOC<l'Mt^Tf 


O    lO    —    GO    lO    t^ 
CO    lO    C»    CC    00    GO 


CO 

_4 

CO 

(M 

o 

CO 

t^ 

r— 1 

lO 

F-H 

lO 

1—1                 -H 

O 

o 

o 

o 

o 

^  o 

o 

*— « 

O     ^      -H 

o 

o 

o 

o 

c 

o 

o  o 

o 

o 

o  o  o 

oo-^     oo     oo     oooooo     ooooo     oooooo     ooo 


m 

i  ^ 

^ 

6J0  1=^ 

^ 

W      ?i, 

pq 


« 


m 


Ph 


pq 


M 


Cm 


P3 


cS 
S3 

■p. 

02 


o 

a. 

50 

s 

o 

w 

« 

o 

CO 

e^ 

V 

o 

"5 

T3 

-i3 


o 


o 


o 


o 


842 


7.  MERCURIALS 


«   2 


A 


<A 

"e 

Ti 

C/J 

C 

T) 

« 

eS 

fl 

rilS 

cS 

>% 

O 

+3 
1 

c 

eS 

0) 

o 

> 

t:^ 

N 


73 

(4 

o 

CI 

OS 

05 

1 

c 

3 

cS 

"e 

cS 

u 

fl 

M 

-1-3 
05 

■» 

g 

S 

3 

0) 

o 
1e 

'3 
k2 

13 

3 

H 

pq 

fe 

02 

§ 

m 

m 

t-    o 

o 

o 

IC    lO    05 

(N    00    '^    <M    tJH 

00  o 

o 

o 

05  oi  00 

fO    i-H    (N    lO    00 

»0    O         O    ^ 

o  o      o  o 


o       ^ 
o       o  ^ 
o      o  o 


o  o      o  o 


O  -H  o 

d      d    -H  d  ^ 


r^      O     O     O     O 


P3     pa 


W  a 


?j,    w 


PQ 


W  P5 


pa 


gi,     Pm   W    4  ?>.  W   a 


pa 


iJ 


^  p^ 


w 


'2     J    s        5"  ^ 


CQ 


^ 

tc 

ft 

^ 

O 

rn 

o 

C^ 

ft 

rC^ 

C 

ft 

<U 

►>> 

O 

o 

3 

o 

T1 

-30. 

a 

a 

ft 


o 


o    o 


INHIBITION  OF  ENZYMES  843 


CO       ^  •— 


O 


2       :::  ^     S 


o 

eS 

T3 

,g 

O  03 


-§11       S       II 

Ct^     Q  M  W  pq  m 


ce 

a 

c 

t3 

3 

O 

O 

O 

^      Tj<      00  ^  W5  Tt      -H 

Tj<mOOOOOOO  OO  o 

-HiMOOOOOOO  00(M  O  OOOOOOOO^ 


pq  +pqpqm  +fLi+  pq 


pq 

+  m 

pq 

S 

h^^ 

§ 

Si, 

W      ?»,   PL, 

a 

1  =     -     I      s     li  -        |ii 


fejpH  K  CMpq  pq  tf  O  f*^'*~' 


W 


o 

^ 

CO 

C 
0^ 

-« 

s 

o3 

3 

3 

<o 

fei 

'K 

s 

o 

t>i 

>i 

U) 

1 

T3 

>^ 

CO 

CO 

c3 

Sh 

!«! 

X 

!«! 

CS 

i 

g 

-O 

2 

a 

"y. 
o 

2 

O 
u 

T3 

2 

C 

CO 

"o 

.2 

to 

c 

'y, 
o 

c 

2 

bj 

CO 

'52. 

?< 

1— 1 

1— 1 

844 


7.  MERCURIALS 


tf 


m   o 


O 


cS 

!>> 

ert 

f-H 

Ti 

05 

<A 



ft 
ft 

W     m 


(M    CO      O    O      O 


05     t^     -^      O 


^ 


o  o    o    o 


o   ^ 
d  d 


d  d 


IC  MM 

^    IC  O    -H    rt    fo 

OO  l>fOOOOO 


pa 


m 


m     m 


m 
S 
a 


M 


Cl, 


fin     W 


fM 


S'    S 


^        o 


fcil 


« 


INHIBITION  OF  ENZYMES  845 


NW^ 

IC 

OS 

(\-i 

^H 

be 

■■ 

a 

tH 

m 

■D 

05 

O 

c 

C  ) 

cS 

^— ^ 

c 

fl 

Ti 

o 

o3 

O   -—■  --^     -H  ^ 

05  ^  CO    ^'        5 


CD  ^  05  ^  .„ 

g>  lo  ^^  -H  fo  E 

g  05  .  ■^-'  CO  cS 

-S2  rl  "^  ^^  "3^  ^ 


'S  L"  ^  ~ 


03       :>       C 


s     ^    j£     S  S  S   -^ 


eo 

o 

»o 

05 

o 

»o 

o 

(^ 

•^    t^ 

1— 1 

o  o 

o 

o  o 

^^^     i-H 

1— 1     O 

o 

o  o 

O    O 

o  o 

rn 

P^ 

m 

M 

pq 

S 

^ 

^ 

§ 

?5, 

a 

?s, 

sa, 

ap  m 

+     + 

+    M 

S  S 

M     bC 

M  S 

55,     ?s, 

ffi     ffi 

W    a 

m    pq    pq  m 

s       1^       s  s 


a  ^ 


<c.  ^ 


s 

s 

03 

+i 

4J 

a; 

^ 

4= 

Si 

<-^ 

<M 

cS 

OJ 

Q) 

w 

tf 

W 

pp 

*  IS   " 

_    _      '^     C^        O        P-i        1"^  i«tf   ?5- 

o        PhPh  pqpqffif^pq  05  w  '^'^ofe) 


00 

% 

A 

u 

u 

e 

tS 

iJQ 

W 

^H 

0 

03 

^"^ 

H 

'O 

e 

-13 

CI 
03 

■^ 

oj 

cS 

0 

'^ 

>^ 

G 

a 

H 

tn 

ti^ 

O  O  -H--HTj<COrtrtOO  00  Oo  00  OOiC  O  -HO 

oiooooooooooo      00      Oo      00      000       o       00 


^   a  W      Ph      Ph  w 


^ 

>. 

K 

0 

£ 

-P 

03 

c3 

C 

0 

^ 

3  f>. 


^ 

•si 

5^ 

0 

s 

"§ 

s 

;■-. 

■4J 

an 

a: 

^ 

846 


7.  MERCURIALS 


P5 


M  ui 


35 

.^^ 

^ 

T3 

T3 

Oi 

CD 

-^^ 

C 

C 

lO 

05 

cS 

c3 

Oi 

fi 

T—t 

^— ' 

O 

c 

a> 

^-' 

m  H 


3     .5 


« 


W 


-H    -H    (M      lO      »C      M 
<M      O    O    O      O      O      CO 


IC  lO    O    — < 

O  -H   ^     o   o   o 

O  O    O     O    O    O     -H 


o  o    o    o  o  o    o 


o^oo    ooo    o^ 


pq    M 

m  m  CM 
§  §  § 

+ 

+ 

?i. 

9,     a 

a    a    ?5, 

a 

i.  a  W 

pq 


pq 


o 


PU      PLH 


m 


^  CO 


pq 


a; 

'bii 

tjj 

c 

n 

>, 

p^ 

X 

0) 

O 

c3 

cS 

T3 

M 


o3     4) 


INHIBITION  OF  ENZYMES 


847 


O 


00     =* 


1^  a    o 


«  X 


S.  o)   ;= 


Xh 

o 

"-' 

;^ 

~J 

Tl 

C 

.,^ 

cS 

«) 

1) 

c 

Cf) 

-M 

JS 

P^    PM    P  ffi    K 


O   lO   o   o   o 

(M    »C    05    O 


>o    O 


t^    O    ffC    o 
O    (M    O 


O       -H 

d  d 


o  ^  ;:: 

O    o    O    -<    CO 


-H       d   d 


IC  ^    1^     ^ 


o  o  o  o 


o  o  o  o 


OOOoO-^O-hO 


pq 

m 

§ 

^ 

S), 

s», 

m 


w  m 


S^     9, 


M 


m 


w  pa  M 


a  a   s*. 


pq  pa  pq  m 
S  §  S  §  ^ 
A    ?^    A  Ph     ?». 


P(5 


pq 


JO       (U 


«    o 


1 

•^ 

^ 

T 

=0 

^ 

o 

»> 
%> 

^ 

o 

« 

rS 

^ 

■? 

►O 

*S 

-<? 

O 

?5, 

'"r^ 

'^  -^  o 


S5, 


5~     e    -s; 
CO    ><i    (^ 


T3 


Is    X 


eS 

fi 

<s 

M 

O 

05 

a 

-fl 

>> 

>> 

X 

c3 

O 

-S 

<g 

>. 

cd 

^ 

^3 

O 

X 

>. 

o 

M 

w 

Q 

O 

< 

<! 

"A 

;?; 

848 


7.  MERCURIALS 


OJ  — 


P5 


m   H 


2   s   -^ 

o 

~    ~    W 


'^         Q    O    "3 


T3 

C 

® 

-M 

1    1 

<D 

j3 

C 

eS 

fl 

<ii   <:   H 


ffi 


^     fl 


w 


W      Q    pq    fL< 


O    M  lO  CD 


CO 

rt      ,-<  o 

O      O      -H    o 

d    d    d  d 


o  o    o  o   o  o 


m 


M    P5 


m   m    +  pq    + 

§      ^        tiC  S        M 
4        ?i,      W      ?5,      W 


P3 


m   m    +  pq 

^      S       M  S 


a   a 


Cm 


fti 

n 

O 

55, 

^^ 

tH 

s 

on 

kd 

+J 

tl) 

«H 

a 

CD 

03 

3 

e 

1 

03 

O 

^ 

tf 

« 

Ph 


pq      pq      oq    Bq    fi^ 


a 

c 


o 

o 
O 

Q 


P    » 


^    o 


INHIBITION  OF  ENZYMES  849 


^*K      ^  TO 


3   J        S   2       ^ 


'^ 

^-' 

73 

t3 

r-«^ 

C 

C 

e 

eS 

C 

C 

<i^ 

cS 

S 

c 
o 

c 

«3 

0) 

3 

o 

W 

i> 

"e 

T) 

cs. 

■«* 

c 

i-H 

^ 

ce 

" — ' 

bi 

_c 

3 

o 

eS 

^ 

Q 

^ 

— 1    o 

»n 

o   o 

m  '^ 

IC 

(M     lO 

OO^"-!!— 1>0^00-^  OO  ^  OOi— lOOO  OfOO(N 

OOOOo'oOOOO^OO  O—iOOOOOO  OOOO 


+   pqS        pqp3        pqS        PQ        P3  W        MPh        M  +eQpq 


Ph  pqp^PLiWpqpq  PhPl,  pq  cqPh 


1^    % 

§1 


T) 

a 

^ 

» 

cd 

d 

o 

« 

? 

T) 

0) 

w 

s 

p 

T3 

^ 

O 

850 


7.  MERCURIALS 


P3 


o 

CO 

^^ 

73 

^ 

cS 

e 

B 

e3 

41 

■» 

cS 

•7S 

3 

05 

O 

o 

^ 

o 

CO 

02 

lO 

i> 

fe 


iS  ^ 


S  W 


>» 

N 

_fi 

o 

e3 

!-l 

-fj 

'a 

.^ 

S 

<D 

cS 

o 

> 

W 

§ 

t3 

T3 

C 

C 

t3 

e3 

cS 

C 

m 

!>. 

tS 

B 

,^ 

-® 

eS 

CO 

u 

■a 

a 

IC    O    O    CO 


iO    Oi      -^    OO    Oi 


—I    o 

d  d 


00     O  -H 

d  d       d  ^ 


o 

o  ^ 
o  o 


«o 

^  <M  CO 

o  o  o 

o  o  o 


O   O     O   lO     o   o   o 


o 


m 


02     c« 

~        CD 


4  a 


ffi 


m  Cl,      pq 


m 


CO  ■— < 


P^    Ph 


ft^ 


(1< 


P5       P5 


PM 


C 

■ft 


a 


^ 

<3i 

CO 

c 
o 
c 

73 

O 

^ 

cS 

cS 

cS 

c3 

03 

c6 

cS 

C 

+J 

s 

-fi 

-tJ 

(- 

-t^ 

<D 

O 

-tJ 

o 

'3 

a- 

o 

-f^ 

o 

a 

3 
T3 

'S 

T3 

3 

^ 

3 

0) 

?^ 

<B 

9 

w 

O 

w 

t4 
O 

W 

t4 

o 

K 

o 

p 

^ 

Q 

^ 

P 

V. 

O 

3 

<i 

'S 

<3 

'k 

<i: 

'x 

<: 

"S 

^ 

o 

^ 

o 

^ 

o 

^ 

o 

f) 

45 

o 

OS 

-»^ 

&> 

3 

T3 

w 

2 

PL| 

O 

P 

■TS 

<Jl 

Xi 

^ 

INHIBITION  OF  ENZYMES 


851 


W 


Pm 


w 


ts^ 


O 


P3 


<N   CO   Q 


9  t^ 


^    — I    — 1       o    ^    o 


o  o  o  o  o 


o  o  o      o  O 


o  o  o  o  o 


W    i,    a      ?5, 


m 


m       pq 


P5 


ea     m 


M 


■5       -^ 


pq 


W 


w 


fej 


Oh 


c 

'a 

CO 


05 


c 

o 

T3 

rt 

a; 

a 

1 

3 

• 

M 

4> 

PM 

O 

fi 

-^ 

<1 

« 

^ 

SI 


s 

'O 

<u 

o 

t4 

(1 

o 

o 

r2 

-C 

'x 

« 

_o 

o 

03 

"o 

W 

Oil 

s 

03 

Pi 

ft 

P 

o 

3 

<: 

c 

« 

^ 

"^ 

^ 

852 


7.  MERCURIALS 


P5 


s 

o    ^ 

s 

fl 

o 

8. 

eS 

M 

■§ 

l> 

<D 

§ 

n 

"I 

H 

H 


M 


Pm 


rS       tS 


05     CO 


p— 1 

a 

,^ 

— ' 

52 

M 

(M 

.-— y 

Ph 

to 

^ 

o 

Ph 

1 

05 

o 

CO 

3 

::i 

O 

M 

T3 

o3 

"e 

73 

O 

o 

"o 

M 

c3 

o 

t3 

c 

o3 

S 

o 

3 

C 

o 

"s 

^ 

3 

o  o 

o 

ao 

05 

O 

o  o 

o 

Oi 

ec 

lO 

o  o     o  o  o 


m 

±   P^ 

w  m  + 

w 

+ 

tc^ 

S    S     M 

§ 

M 

W   4 

a    ?i,  W 

a 

W 

^    .^ 


"^  >'■  m    pC 


Ph 


—^  00 


c 

~' 

cS 

bo 

s 

4) 

M 

^ 

o 

p 

» 

o 

c 

05 

eS 

:;i 

G 

0) 

t-i 

O 

OJ 

^ 

S! 

<     P3 


^W 


SI 

p< 

o< 

s 

o 

«2 

Ph 

T3 

eg 

Xi 

C 

Ti 

^ 

a 

ee 

fl 

O 

c3 

^^ 

c3 

05 

m 

7^ 

.SP 

l-H 

(» 

bO 

'^-^ 

-C 

!> 

^ 

0^) 

M 

O 

0) 

0) 

3 

Q 

w 

h:] 

M 

o  o 

o 

lO 

o  o 

o 

CD 

VO 

OO    O    lO 

-*    O    C<)    (M 

O    O    O    O       -H 


o     ^     ^  »c 


o  o  o  o  o     o  o 


P5 


p:)  P5 
9,  ?i,  P^ 


en     aj 

3     3 
S     S 


Ph  Ph   « 


P5 


O    (M 


+  m    pq    +  « 

txi  ^     §      ao  S 
ffi   a    ?»,    W   ?». 


o    o 
Ph  P^ 


4) 

3 

a 

c 
-2 

3 

II 

;3 

2 

8 

as 

02 

3 

'W) 

3 
a; 

>1 

> 

-2 

.s-i 

o 

M 

Tn 

3 

">. 

1     X 

O 

ft 

2 

O 

S 
o 

o 

^ 

T5 

-3 

-3 

^ 

P4 

O 

>> 

P( 

ft 

ft 

m 

/-} 

J3 

M 

tn 

09 

O 

Ph 

o 

O 

j3 

O 

j3 

Ph 

Ph 

fO 

Ph 

Ph 

Ph 

o 


(Ti 

T3 

-^ 

^— ' 

es 

C 

1 

c 

03 

Z^ 

m 

H 

toll 

o 
O 

§ 

K 

^ 

^ 

X! 

^ 

O 

eS 

-C 

h-l 

« 

o 

INHIBITION  OF  ENZYMES  853 


— ^      ^^        i> 


^  OS  o5       c         ^ 


m 


^ 

05 

-»^ 

cS 

11 

M 

^^ 

c 

O 

3 

"S 
ft 

ft 

X 

GOO  -*00         C5G005t>'*QOaOC5000>OS<lX'— 100 


rt  ^  ■>*   CO  ■<**  CO   Tti 

O  OO^         OOOO         OOoOOOOC<l(MC<IC0OOOOO 

O  OOO        OOOO        OOoOOOOOOOOOOOOO 


P5Wm  pq  pq  pQ+fq+pq+  +pq+pqeq+        pq 


+   fq  +    pq  + 

+   pq  +    pq  pq 

M  S      M  S      M 

^ 

M  S      M  S    S 

te    i.K    i,W 

Ph 

K   a  ffi   i,  3, 

s 

«. 


.a  9i  ^^         eS  S  ■«  <u  s  eS     eS 


PM 

Ph 

f/} 

rrt 

>, 

u 

Ci 

o 

m 

pO 

G 

(4 

o 

1 

js 

O 

ft 

Tl 

o 

$J 

o 


o 

">. 

c 
o 

■ft 


3 

© 

1 

'^ 

(S 

!«1 

T1 

O 

« 

O 

^ 

o 


PM         Ph         P-i  fM 


854 


7.   MERCURIALS 


P5 


^-  C5 


eS 

C 

<l> 

M 

O 

<v 

t-( 

TJ 

t^ 

!>. 

N 
G 

T3 

Cm 


c 

n 

05 
C5 

O 
T3 

Tl 

' — ■ 

C 

fi 

ee 

-i«! 

a 

3 

a 

IS 

n 

o 

o 

ci 

^ 

o 

C-  IS 


b     . 


^       .-.       CO 


m  g 


W     b 


e 

"^ 

es 

m 

fl 

O 

0! 

rQ 

r> 

S 

bX 

CL, 

<t3 

o    o    00 


ooooo     ooooooo 


cq 

W 

pq 

+ 

pq 

pq 

+ 

M 

+ 

S 

§ 

§ 

W) 

^ 

§ 

br 

g 

bO 

?i, 

$5, 

55, 

a 

a 

Si, 

a 

?i, 

W 

pq 


bC 

a 


C55      C^      aj     .S 


c3       B 


f^      (^ 


tH      bD 


3 
P-l 


CM 


O    p^    lO 

o  o  o 
o  o  o 


pq       pq 


&q 


PM 


Ph  PlH  PlH 


INHIBITION  OF  ENZYMES 


855 


C5  ^-' 


00     ;r: 


10-=  -< 


^^ 


05 

0 

OJ 

10 

•^-^ 

^— ' 

Oi 

to 

45 

-H 

->J 

10 

3 

0 

c3 

0 

^ 

b 

> 

t3 

e 

TJ 

C 

■^^ 

c 

'^ 

d 

%) 

eS 

c2 

c8 

i 

03 

^    10 


Ui  m  Q 


M 


W   fe   o 


X   § 


^        t« 


o 


pq 


O    t-    t^    o 

O    l>    00    o 


O     O     LO    o     t^     ''^ 

05      O       05     O      '32       t^ 


(M 

10 
0 

>o 

»c 

CO 

CO 

^ 
"o 

0 

0 

0 

0 

r- 

1— ( 

C<l 

M 

•* 

<N 

0 

(^ 

0 

^^ 

0 

c 

^ 

2 

»c 

(N 

»o 

«o 

1— ^ 

0 

*--( 

•"H 

(M 

Tf 

>— 1 

0 

>— 1 

0 

0 

0 

0 

0 

0 

5, 

X 

r:^ 

t^ 

0 

0 

0 

0 

0 

0 

■^^ 

■"*_ 

0 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

n 
0 

»o 

0 

OS 

d 

—1    d 

d 

d 

d 

d 

d 

d 

d  — 1 

d 

m 


m  13  m 


m 


ffi 


m 


M  m  pq 


?S<      a      ?5. 


^  ^  s 


^ 


w 


> 

03 

-2^ 

0 

CO 

(D 

k^ 

>^ 

^ 

0 

C 

^ 

^ 

0 

0 

C 

ee 

>> 

3 
^ 

ft 

S2 

w 

w 

W 

c 
'ft 


a,   05 


'^    Q,    Ph 


-^ 


P3 


3 

CO 


856 


7.  MERCURIALS 


P5 


^^ 

^-^ 

O 

CO 

— ^ 

ff 

»o 

CO 

lO 

lO 

t^ 

^— ^ 

« 

•* 

■* 

05 

Oi 

lO 

i-H 

^ 

2 

05 

,-^ 

^^ 

Oi 

05 

c 

a; 

CO 
05 

O 

13 

05 
C5 

'2. 

3 

CO 
05 

C 

(35 

2 

> 

o 

o 

e8 

T3 

1 

S 

2 
S 

S 

c 
o 

o3 

u 

o 

CO 

O 
»-5 

J2 

C 

® 

>> 

i 

00 
J2 

W 

^ 

^ 

w 

W 

C5 

CQ 

1 

w 

^ 

W 

C 


■^OCOiCCOOOOO-^'-^OO 
C^C0-HO5O5lOlO'OO5C5'^f- 


rt  O  -<       O      -H 

o    ■<**■-'■*  o   o  o 
o   o   o  o  o   o  o 


oooioooooooo 


-H  (M  »0 

O  O  O  -H 

o  o  o  o 

<6  <Z>  Oi  c^ 


m 
S 

m 
S 

a  a 

W   ?i,  a 

55, 

?», 

g  :=  s  =!  -«  -^  ^ 

S^„  -k^  -tJ  +J  -1-3  -*i  -4-i 

gf  eS  c3  cS  o3  eS  cS 

a,  P5  D5  Pi  P5  tf  P5 


^     pq 


P3 


a> 

cS 

.13 

TJ 

c 

s 

>> 

(U 

^ 

U) 

a> 

O 

0 

TS 

;.4 

TJ 

"ab 

3 

0) 

>. 

o 

s 

OQ 

f^ 

00 

-»^ 

e8 

Ci 

« 

9. 

S 

C 
o 

-o 

CO 

^ 

bO 

0^ 

C 

3 

-2 

1 

>> 

o 

to 

'S 

'S 

X 

« 

C 

cS 

o 

b4 

eg 

^ 

OQ 

H 

H 

oo 

05      ^^ 


c  2  i 


INHIBITION  OF  ENZYMES  857 


2  =5      S 


■  -^  »--i  wj  >       w  w^  ■^^  f«»i,  ;::,     '-/j 


^  °  i   s      s     111     i 

Ills  I  t     :3   I         1      =-      ^        g 


73  S 


e 

4^    -t^     c 

fl 

0 

f^ 

hear 
hear 
brai 

■3 

0 

M     M    "^ 

sg 

-u 

=? 

S     S     P5 

pq 

a. 

tf 

QO     CO 

IC  OOO<M00OiKI 


0000       o       o        ^  00       OOOOO 


pqpqeSffi  m  pQfqpQ 


w 

M 

cc 

m 

eq 

pa 

a^ 

pq 

CM 

+ 

fL, 

m 

^ 

§ 

§ 

S 

s 

+ 

S 

p. 

?i. 

4 

a 

i, 

s. 

K 

A 

•—        qS    ••-■  OJ        .—        .« 


bO 


S      c3  OS         eS  c3  TO  0)       . 


-S 


ea 


s 

-5 
eg 
> 

3 

-iS 

s5 

3 

0 

2 

<» 

(D 

c> 

^ 

eg 

M 

-H! 

s 

u 
4^ 

> 

bi 
J 

'6 
c 

a 

0 

0 

c 

3 

0 

"3d 

'3d 

03 
c3 

"s 

s 

£ 

5 

o 


u 

"a 

>!>, 

c 

0 

cr! 

0 

ft 
>> 

3 

0 

H 

858 


7.  MERCURIALS 


03 


ic         — 


O 


P. 

B 
o 

H 


O 

o 

Tj*     CO     O 

t- 

o 

-H      ^      lO      O 

o 

o 
o 

o 

o 

o 
o 

o 
o 

o 

o 

i 

O 

o 

o  -^  o 

o 

o 

o  o  o  o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

M 


W 


m 


c^ 


X 


w 


pq 


m 


pq 


pq 


&q 


fei 


p^ 


s 

c 


ft 

o 

o 

ft 

j5 
ft 

s 

X 

Pi 

H 

Is  s 

o   ^ 

d       CD 

ft  2 


-C 

_rt 

ft 

"3 

o 

M 

-1-^ 

-■    >> 

ft 

CM 

PM  ^ 

^ 

Q 

S  -S 

H 

P 

t) 

INHIBITION  OF  ENZYMES  859 


g  lO  o  g* 

O  §  2  .S 

T3  -  -  « 

i  ^  -^  -^ 


X  J;^  ft  s-^3.2S3 


OCOr^iOfOiOCD       00        GO  O        O        GO-hiOOOOOO 

05C5G5'-HfCiCOC)       <N        05  C5        O        COiOOO-^OOO 


OOOOO  OOTtt— ifOO  «3  o  «3  "Zt 

OOOOO  OOOOOO  O  O-^OtN  o 

OOOOO       OO       OOOOO    oo       o        O  O        O        OOOOOC^I-^ 


CO 

m  pq+  +  pq-[-fQMm+  m  Ph 


oq 


o 

?» 

"p 

ki 

o 

_o 

» 

to 

CO 

S3 

'-3 

_>; 

•§ 

cS 

o3 

ki 

V 

Si 

0) 

-*i 

^ 

s 

02 

'2 

-ii 

_u 

■is 
o 

o 

►-5 

O 

O 

O 

:3 

c 

<D 

&D 

a 

2 

bX) 
o 

O 

.1 

J5 

-c 
-§ 

CO 

O 

« 

« 

fl 

ft 
O 

2 

c6 
C 

13 

ft 

a 

03 

ee 

"3 

O 

T3 

'>. 

'>, 

^ 

X 

X 

X 

860  7.  MERCURIALS 

inhibitions  constitute  mutual  depletion  systems  and  are  in  zones  B  or  C 
(see  page  1-66).  The  inhibition  is  independent  of  K^  in  zone  C  —  i.e.,  i  = 
(I;)/(E^)  —  and  so  the  degree  of  inhibition  indicates,  not  the  affinity  of  the 
enzyme  for  the  mercurial,  but  only  the  concentration  of  enzyme  (or  of  other 
substances  binding  the  mercurials).  When  we  see  in  the  table  that  0.001  raM 
p-MB  inhibits  an  enzyme  50%,  does  this  mean  that  K^  is  approximately 
0.001  vciM,  or  that  the  enzyme  concentration  is  0.002  mM  (assuming  pure 
enzyme)?  It  is  actually  more  likely  to  be  the  latter,  and  this  is  one  reason 
why  the  values  in  the  table  should  not  be  taken  too  seriously.  These  ques- 
tions will  be  considered  in  the  following  section. 

It  is  rather  difficult  to  find  many  enzymes  which  are  insensitive  to  the 
mercurials.  The  following  enzymes  might  be  considered  as  relatively  resis- 
tant ( <  10%  inhibition  at  1  mM  or  above):  adenylate  kinase,  most  alkaline 
phosphatases,  a-amylase,  potato  apyrase,  cellobiase,  coagulases,  copropor- 
phyrinogen  oxidase,  dinitrophenol  reductase,  elastase,  /5-glycerophospha- 
tase,  glycolate  oxidase,  kynurenine  formamidase,  certain  lipases,  maltase, 
neuraminidase,  nitrite  reductase,  oxalate  decarboxylase,  pepsin,  peroxid- 
ases, phospholipases,  many  proteases  and  peptidases  (especially  bacterial, 
fungal,  and  venom),  some  pyrophosphatases,  most  RNases  and  DNases, 
certain  sulfatases,  thiamine  diphosphatase,  and  uricase.  These  certainly  do 
not  constitute  as  a  whole  a  very  extensive  or  particularly  important  group 
of  enzymes.  Actually,  the  majority  of  enzymes  are  inhibited  in  an  inter- 
mediate fashion  between  these  and  the  examples  in  Table  7-13,  i.e.,  50- 
100%  by  concentrations  in  the  range  0.05-1  mM,  although  some  of  these 
would  undoubtedly  exhibit  a  much  greater  sensitivity  if  examined  under 
appropriate  conditions  (in  a  pure  form,  at  physiological  temperatures  and 
pH,  and  in  the  absence  of  high  concentrations  of  ligands).* 

Comparison  of  Mercurials 

Since  the  introduction  of  p-MB  some  30  years  ago,  it  and  the  related  p- 
MPS  have  been  used  for  enzyme  inhibition  more  and  more  frequently  at 
the  expense  of  Hg++.  It  is  interesting,  therefore,  to  look  into  the  results 
which  have  been  obtained  with  the  inorganic  and  organic  mercurials.  Of 
the  total  of  160  reports  on  enzyme  inhibition  using  both  Hg++  and  p-MB, 
25%  do  not  allow  an  accurate  comparison,  due  mainly  to  different  concen- 

*  A  word  should  perhaps  be  said  against  the  common  practice  of  reporting  only 
an  inhibition  of  100%  with  a  single  mercurial  concentration,  particularly  if  this  is 
relatively  high,  since  such  results  are  not  very  meaningful  and  the  enzymes  cannot 
be  accurately  classified  as  to  sensitivity.  For  example,  to  state  that  1  mM  p-MB 
inhibits  100%  is  bad  on  two  counts:  One  does  not  know  the  true  sensitivity  of  the 
enzyme,  since  0.01  mil/  might  also  inhibit  100%,  and  such  a  result  does  not  provide 
much  evidence  for  the  importance  of  SH  groups  at  the  active  center,  although  it 
is  often  so  interpreted. 


INHIBITION  OF  ENZYMES  861 

trations  being  used,  or  statements  that  both  at  certain  concentrations  in- 
hibit 100%.  Of  the  remaining  reports,  Hg++  is  more  potent  in  65%,  p-MB 
in  29%,  and  in  6%  they  are  of  equal  potency.  One  must  admit  that  Hg++ 
is  generally  more  effective.  In  some  cases  it  is  of  much  greater  inhibitory 
potency  than  p-MB  or  the  other  organic  mercurials.  One  might  expect  Hg++ 
to  be  more  potent  than  p-MB  because  (1)  it  is  smaller  and  might  be  able 
to  penetrate  and  react  with  SH  groups  inaccessible  to  the  larger  molecule, 
and  (2)  it  could  possibly  in  some  instances  induce  dimerization  of  the  en- 
zyme (or  even  polymerization),  since  it  is  bifunctional.  On  the  other  hand, 
p-MB  might  be  considered  to  shield  off,  sterically  or  electrostatically,  a 
greater  area  on  the  enzyme  surface  once  it  has  combined  with  the  SH  groups, 
due  to  its  greater  size  and  the  charged  COO"  group.  The  result  in  any  case 
is  probably  a  balance  of  these  and  other  factors.  The  fact  that  Hg++  is 
often  more  inhibitory  than  p-MB  does  not  immediately  imply  that  it  is  a 
better  or  more  reliable  inhibitor  to  use  for  the  purpose  of  detecting  SH 
groups  on  enzymes,  but  it  does,  I  think,  suggest  that  Hg++  has  been  un- 
necessarily neglected  by  many  workers.  It  might  be  proposed  that  both 
mercurials  be  used,  since  not  only  will  the  detection  of  SH  groups  be  made 
more  certain,  but  occasionally  interesting  information  on  the  nature  of  the 
inhibition  may  be  obtained. 

The  organic  mercurials  themselves  have  not  often  been  used  in  the  same 
investigation,  but  in  19  reports  using  both  p-MB  and  PM,  I  have  found  PM 
to  be  more  potent  in  79%.  The  differences  between  them  are  seldom  very 
marked,  however.  One  would  not  expect  much  difference  between  p-MB 
and  2?-MPS,  and  examination  of  the  eight  reports  using  both  bears  this  out, 
in  two  p-MB  being  the  more  potent,  in  two  p-MPS  being  the  more  potent, 
and  in  four  the  potencies  being  the  same.  There  is  some  reason  for  believ- 
ing that  the  smaller  uncharged  alkyl  mercurials,  such  as  MM,  might  be 
better  for  enzyme  study  than  any  of  the  other  mercurials,  but  there  has 
been  so  little  comparison  that  nothing  can  be  said  definitely  about  their 
relative  effectiveness  at  this  time. 

Meaning  of  K,-  and   Methods  of  Expressing  Inhibition   by  the   Mercurials 

The  values  for  Kj  have  occasionally  been  reported  for  mercurial  inhibi- 
tion; e.g.,  for  the  inhibition  of  phosphoribosyl-ATP  pyrophosphorylase  by 
p-MB,  pZ,  =  5.15  (Martin,  1963),  and  for  the  inhibition  of  heart  lactate 
dehydrogenase  by  p-MB,  pK^  =  4.10  (Gruber  et  al.,  1962).  In  the  latter  case 
the  binding  of  the  mercurial  to  the  noncatalytic  SH  groups,  causing  the 
spontaneous  reversal  of  the  inhibition,  is  characterized  by  a  p^j  of  5.40. 
Most  of  the  values  given  for  p^,  are  in  the  range  4-6.  However,  these  values 
were  obtained  by  simply  taking  the  concentration  to  produce  50%  inhibi- 
tion, which  would  be  valid  if  (1)  the  inhibition  is  classically  noncompetitive 
(which  in  most  cases  probably  it  is  not),  and  (2)  the  system  is  in  zone  A 


862  7.  MERCURIALS 

(which  it  seldom  is).  If  the  system  is  in  zone  B  or  C,  K^  may  be  a  good 
deal  smaller  than  PI50,  and  we  have  seen  that  in  zone  C  there  is  no  way 
kinetically  of  determining  K^.  The  only  valid  calculation  of  a  true  dissocia- 
tion constant  for  a  mercurial  complex  with  an  enzyme,  of  which  I  am  aware, 
is  that  of  Madsen  and  Gurd  (1956)  for  muscle  phosphorylase  and  p-MB. 
They  used  an  ultracentrifugal  method  to  measure  the  concentration  of  free 
p-MB  after  equilibration  and  determined  K^  from  a  plot  of  1/r  against 
r/(p-MBy),  where  r  is  the  molar  ratio  of  p-MB  bound  to  protein  and  p-MBy 
is  the  free  mercurial.  A  value  of  p^,  =  6  was  found.  It  is  likely  that  in 
most  cases  in  which  an  enzyme  is  potently  inhibited  by  a  mercurial,  a  p^, 
of  6  or  less  would  be  found,  and  in  this  range  it  is  very  difficult  to  deter- 
mine the  constant  by  the  usual  plotting  procedures.  It  will  be  recalled  that 
W.  L.  Hughes  (1950)  found  a  p^  of  4.46  for  the  complex  of  mercaptalbumin 
and  MM  (page  759).  We  may  now  inquire  into  what  values  of  K^  would  be 
predicted  under  ordinary  circumstances.  Equation  7-3  gives  the  relation 
between  the  experimental  constant  and  the  dissociation  constant  for  a  mer- 
curial complex,  and  if  we  alter  it  to  correspond  to  inhibition  of  enzymes 
we  have 

p^/  =  vKi  —  pKa  -  p-fi'x  , 

where  ipK/  is  the  experimental  or  apparent  dissociation  constant.  If  p^,  is 
taken  as  21,  \)K„  as  8.7,  and  ipK^  (for  Cl~)  as  6.5,  all  of  these  being  approxi- 
mations, Y>K/  turns  out  to  be  around  5.5.  Since  pK^  certainly  varies  from 
20  to  22,  pK^  from  7.5  to  9.5,  and  p^^,  (depending  on  the  ligand)  from  6  to 
9,  it  is  clear  that  pK/  may  vary  over  a  wide  range,  but  at  least  the  values 
experimentally  determined  are  of  the  correct  order  of  magnitude.  The  ex- 
perimental i)K/  is  thus  a  good  deal  less  than  the  true  p^,  because  of  the 
competitive  effects  of  H+  and  the  ligand  X~. 

The  question  of  how  best  to  report  mercurial  inhibitions  is  a  difficult 
one.  First,  the  concentration-inhibition  curves  are  often  very  steep  (Fig. 
7-31)  so  that  giving  the  results  of  a  single  concentration  may  be  quite  mis- 
leading. Therefore  one  can  suggest  that  in  all  studies  a  range  of  mercurial 
concentration  be  used,  such  as  to  provide  different  degrees  of  inhibition, 
preferably  from  0  to  100%.  Second,  values  of  K^,  which  are  so  useful  in 
other  inhibitions,  are  difficult  if  not  impossible  to  determine  by  the  usual 
procedures,  especially  when  the  systems  are  in  zones  B  or  C,  in  which  case 
plgo  may  vary  greatly  depending  on  the  enzyme  concentration.  It  is  evident 
that  for  work  with  pure  enzymes  it  is  best  to  state  the  amount  of  inhibitor 
present  in  terms  of  //moles  per  milligram  of  enzyme,  or  if  the  molecular 
weight  of  the  enzyme  is  known  to  express  this  as  a  molar  ratio.  However, 
when  impurities  are  present,  and  especially  when  preparations  such  as  ho- 
mogenates  are  used,  this  method  is  not  as  useful  and  even  a  designation 
such  as  //moles  of  mercurial  per  milligram  of  total  protein  is  not  very  mean- 
ingful. Third,   as  we  have  discussed   previously,   mercurial  inhibition  is 


INHIBITION  OF  ENZYMES 


863 


strongly  dependent  on  several  factors,  such  as  pH,  temperature,  and  me- 
dium composition,  so  that  these  conditions  should  be  stated  accurately 
and  completely,  and  it  should  always  be  realized  that  the  results  reported 
apply  only  to  these  particular  conditions. 


Fig.  7-31.  Inhibitions  of  glutamate  dehydro- 
genase, showing  the  relative  potencies  of  the 
various  inhibitors.  Glutamate  =11  mM,  NAD 
=  0.17  mM,  and  pH  7.6.  (From  Olson  and 
Anfinsen,    1953.) 


Inhibition  of  ATPase 

The  results  of  the  actions  of  the  mercurials  on  ATPase  were  not  included 
in  Table  7-13  because  they  are  complex  and  warrant  more  detailed  treat- 
ment, particularly  as  the  effects  of  mercurials  on  mitochondrial  and  myosin 
ATPase  are  of  importance  in  the  work  on  oxidative  phosphorylation  and 
muscle  contraction,  respectively.  Some  of  the  reported  inhibitions  of  ATPase 
are  shown  in  Table  7-14;  the  stimulation  of  ATPase  under  certain  conditions 
has  been  omitted  since  it  was  presented  in  Table  7-11.  One  immediately 
notes  a  very  great  variation  in  results.  This  is  due  partly  to  the  different 
sources  of  the  enzyme,  but  also  to  the  different  conditions  under  which 
the  experiments  were  run.  The  response  to  mercurials  depends  on  the  state 
of  activation  of  the  enzyme,  whether  Ca++  or  Mg++  is  present,  the  pH,  and 
the  temperature,  as  well  as  the  obvious  factors  of  buffers  and  nonenzymic 
protein.  The  pH  actually  determines  whether  stimulation  or  inhibition  will 


864 


7.  MERCURIALS 


03 


O 


4 


ft3 


O  S 


5 

fl 

1 

.2 

(-C 

'-2 

C8 

o 

ft 

cS 

£ 

i 

a^ 

4J 

03 

o 

-C 

o 

M 


S  «^      ^ 


CQ 


OS 

^■^ 

to 

CO 

OS 

CO 

1— 1 

OS 

" — ' 

l-H 

<a 

TS 

•S 

c 
o 

o 

05 

OS 

G 

CO 
05 

(4 

i 

a 

o 

^3 

'r* 

>. 

C 

fM 


c« 

es 

O 

S 

o 

C 

eS 

eS 

a 

<U 

»H 

<U 

M 

ft 

^ 

o 


OS     e 


ft 

n 

n 

M 

m 

t8 

•E 

S 

§ 

§ 

01 

a 

?s, 

A 

?j, 

S 

4    W 


-hcOC<IOS(N'-^-hOS 

cocoi>i>xcoi>t> 


H 


W 


^  I 
"ft  3 
O    .« 

J-     <v 


I 


bC 


INHIBITION  OF  ENZYMES  865 


S^SSii'-'^  ^  —  OS 

2   2      §     ^  --• 


^        r-        -^ 


h-;hp5:2;cc       ^h-i         ocoPm         <!iPh 


^-^ 

05 

o 

>c 

^_^ 

^^ 

o> 

cc 

B 

»Ci 

^ 

^^ 

05 

« 

en 

B 

" — ■ 

C5 

>, 

C 

c 

^ 

-a 

o 

:0 

o 

02 

CO 

t-  o   o 
00  o   t- 

o 

CO 

^3 

^^ 

c 

<u 

ei 

be 

>. 

^ 

-a 

cS 

ITt 

w 

02 

(M      -HGOT}<t^      t^O<M05      OOI> 


O    -H    cq  »o 

O    O    O  O    -H    ic 


OO      ^      OOOOM  oo      o  -^ooo-^oooooo-^ 


MMPQM  pq  +pqp5 

■  ■    s    § 


Wa.©.©,?),      PLia         ffi?;,?!,         plhK 


?!.      A 

54, 

;0     t> 

O 

-o 

4) 

^ 

tijj 

C3 

Ph 

X 

ui 


s    ^ 


»       5r!  at, 


866 


7.  MERCURIALS 


^ 


a- 


Tt<    oi         ri         o 

2  ^      a      ^ 


M 


O 


1^    ^ 


W 


W     O     O  ^ 


3    ^ 

H 


;ir      u 


«0    OO      CO    »o 


o  -<  o 
odd 


O     CO   M     o   o 
d    d  -^    d  d 


a    Ph 


o^ 


CO 
o 

o 

o 

CO    00 
CD    CD 
O    O 

O 

o 

o 

O    O 

o 

m 
§ 

m 
S 

^ 

A 

a 

a 

a 

S 

a 

ft 


»0    lO 

00     03 


3       +s 


pq 


P3 


PL, 


INHIBITION  OF  ENZYMES 


867 


be  exhibited  over  a  wide  range  of  mercurial  concentration  (Fig.  7-32).  The 
fairly  sjonmetrical  curves  for  myosin  ATPase,  the  maximal  stimulation 
being  observed  at  a  pH  around  7.5,  are  most  likely  the  result  of  changes  in 
ionizable  groups  at  or  near  the  active  center,  whereas  the  more  complex 
curve  for  mitochondrial  ATPase  perhaps  arises  from  additional  factors  relat- 
ed to  mitochondrial  structure  or  the  effects  of  intramitochondrial  compo- 
nents on  ATPase.  It  may  also  be  mentioned  that  Tonomura  and  Furuya 
(1960)  found  essentially  the  same  behavior  for  myosin  B  ATPase,  stimul- 
ation being  maximal  at  pH  7.8  and  absent  at  5.7  and  10. 


+  250- 


•200 


+  100 


+  50 


-100 


MITOCHONDRIA 


8 


Fig.  7-32.  Effects  of  pH  on  the  actions  of  p-MB  on 
ATPase.  Liver  mitochondrial  ATPase  treated  with  p-MB 
at  0.1  mM  (Myers  and  Slater,  1957  b).  Myosin  ATPase 
curve  #  1  treated  with  p-MB  at  0.04  /<mole/mg  (Stracher 
and  Chan,  1961),  and  curve  #2  treated  with  p-MB  at 
0.0872  /<mole/mg   (Blum,   1962  a). 


When  ATP  is  added  to  a  preparation  of  myosin  ATPase,  there  is  an  initial 
burst  of  phosphate  release,  followed  by  a  steady  level  of  hydrolysis.  The  ef- 
fects of  p-MB  on  these  two  phases  of  activity  have  been  shown  to  be  quite 
different  by  Tonomura  and  Kitagawa  (1957,  1960).  There  is  a  progressive 
depression  of  the  magnitude  of  the  initial  burst  as  the  SH  groups  are  ti- 
trated, but  the  steady  rate  is  accelerated  until  around  80%  of  the  groups 
have  been  combined  (Fig.  7-33).  The  rate  of  the  initial  burst  may  be  stim- 
ulated but  the  amount  of  ATP  split  during  this  period  is  reduced.  How 
these  observations  correlate  with  the  various  theories  of  how  mercurials 
activate  ATPase  (page  819)  is  not  known;  for  example,  does  the  initial 


868 


7.  MERCURIALS 


burst  have  anything  to  do  with  the  postulated  structural  rearrangements 
of  the  enzyme  following  addition  of  ATP  ?  In  any  event,  there  is  evidence 
that  mercurials  can  alter  not  only  actomyosin  stability  itself  (Gergely  et  al., 
1959)  but  the  structure  (Kominz,  1961)  or  flexibility  (Levy  et  al,  1962)  of 
the  enzyme  near  the  active  center. 

The  configurational  changes  are  apparently  complex  inasmuch  as  meas- 
urements of  the  optical  rotatory  dispersion  indicate  that  p-MB  up  to  4 
moles/ 10^  g  myosin  A  increases  the  helical  content  several  per  cent,  but 
addition  of  8  moles/10^  g  decreases  the  hehcal  content  (Tonomura  et  al., 
1963  a).  It  was  suggested  that  these  small  changes  in  the  helical  structure 

200 


150- 


100 


-   50 


UJ   < 

tr  > 


^ 

\      STEADY 

/ 

/ 

\       RATE 

\ 

\ 

^/titration 

\ 

\           ^ 

^         OF    SH      GROUPS 

N.       INITIAL 

\^^ 

\burst 

y^ 

/> 

\V 

0  0  02  0.04 

p-MB    (/iMOLES/MG)     


006 


0.08 


0.10 


Fig.  7-33.  Titration  of  myosin  B  ATPase  with  p-MB  and 

the  effects  on  the  initial  and  steady-state  rates  at  pH  6.7 

and  20°.  The  SH  reaction  measured  by  absorption  at  255 

m/<.   (From  Tonomura  and  Kitagawa,  1960.) 


induce  modifications  at  the  active  site.  If  the  inactivated  enzyme  is  treated 
with  /5-mercaptoethanol  the  mercurial  is  removed,  the  activity  is  restored, 
and  the  rotatory  dispersion  returns  to  normal  (Tonomura  et  al.,  1963  b). 
One  might  then  assume  that  the  effects  of  the  p-MB  are  completely  rever- 
sible; however,  it  was  found  that  the  substrate  inhibition  at  high  concentra- 
tions of  ATP  no  longer  occurs,  and  that  EDTA  now  does  not  inhibit.  It 
may  be  that  the  treatment  with  ^j-MB  removes  divalent  cations,  possibly 
the  tightly  bound  Ca++.  It  was  indeed  shown  that  the  progressive  depression 
of  the  ATPase  activity  is  accompanied  by  a  loss  of  the  binding  of  Ca++  and 
Mg++  (Martonosi  and  Meyer,  1964). 


INHIBITION  OF  ENZYMES  869 

The  interesting  treatment  of  the  kinetics  of  myosin  ATPase  inhibition 
with  p-MB  recently  reported  by  Walter  (1963)  should  be  consulted  for  the 
theoretical  development  of  the  equations  and  appropriate  plotting  proce- 
dures, but  in  connection  with  our  present  subject  it  is  worth  noting  that 
the  kinetics  are  complicated  by  two  factors  which  should  be  kept  in  mind 
in  such  work  on  all  enzymes.  First,  the  reaction  of  the  first  SH  groups 
does  not  lead  to  inactivation  immediately.  Second,  this  initial  reaction  re- 
duces the  mercurial  concentration  so  that  the  more  slowly  reacting  groups 
are  exposed  to  a  much  lower  concentration  than  was  originally  added.  The 
calculated  bimolecular  rate  constant  for  these  catalytically  important  SH 
groups  is  thus  quite  different  than  that  obtained  from  the  initial  rate. 

The  relationship  between  the  effects  of  mercurials  and  2,4-dinitrophenol 
is  very  interesting  and  the  results  indicate  that  SH  groups  are  involved  in 
the  stimulation  of  ATPase  activity  by  the  latter  substance.  Lardy  and 
Wellman  (1953)  noted  that  0.04  mM  p-MB  almost  completely  abolishes 
the  activation  of  mitochondrial  ATP  splitting  by  DNP,  and  others  not 
only  have  confirmed  this  in  general  but  have  shown  that  the  DNP-activated 
enzyme  is  only  inhibited  by  mercurials  (Greville  and  Needham,  1955;  Gil- 
mour  and  Griffiths,  1957;  Pullman  et  al.,  1960).  This  is  shown  for  myosin 
ATPase  in  Fig.  7-30.  However,  many  factors  can  influence  these  interac- 
tions. Myers  and  Slater  (1957  b)  found  that  p-MB  inhibits  the  DNP-acti- 
vated mitochondrial  ATPase  from  pH  6  to  8,  but  stimulates  further  at  pH  9, 
and  Cooper  (1958  b)  has  emphasized  the  importance  of  Mg++,  addition  of 
this  ion  decreasing  the  inhibition  by  p-MB  of  DNP-activated  enzyme,  as 
originally  shown  by  Lardy  and  Wellman.  Not  all  ATPases  may  behave  in 
this  fashion,  and  the  activity  of  a  particulate  preparation  from  Rhodospi- 
rillum  rubrum  can  still  be  stimulated  by  DNP  in  the  presence  of  p-MB  and 
Mg++  (Cooper,  1958  a).  Some  investigators  have  postulated  that  DNP  and 
the  mercurials  activate  ATPase  by  similar  mechanisms,  but  it  seems  doubt- 
ful if  the  evidence  is  sufficient  to  draw  this  conclusion.  It  has  recently  been 
found  that  COg  stimulates  mitochondrial  ATPase  markedly  and  this  occurs 
in  the  presence  of  0.005  mM  Hg  ++,which  itself  has  broght  about  activation; 
indeed,  Hg++  activates  about  the  same  in  the  absence  or  presence  of  COg 
(Fanestil  et  al.,  1963).  These  stimulations  thus  appear  to  be  approximately 
additive.  The  K+-Na+-activated  membrane  ATPase  contains  SH  groups 
which  seem  to  be  specially  involved  in  this  activation  and  react  readily 
with  mercurials  (Skou,  1963). 

The  interesting  effects  of  the  mercurials  on  the  P,-ATP  and  ADP-ATP 
exchange  reactions  occurring  in  mitochondria  and  the  relations  to  ATPase 
will  be  discussed  later  under  oxidative  phosphorylation  (page  872). 


870  7.   MERCURIALS 

ELECTRON   TRANSPORT 
AND   OXIDATIVE   PHOSPHORYLATION 

The  majority  of  dehydrogenases  are  quite  sensitive  to  the  mercurials, 
and  inspection  of  Table  7-13  shows  that  50%  inhibition  is  commonly  pro- 
duced by  concentrations  of  0.001-0.05  mM.  If  we  define  NADH  dehydro- 
genase as  the  enzyme  component  responsible  for  the  transfer  of  electrons 
to  a  variety  of  acceptors,  it  must  be  placed  in  the  same  category  with 
respect  to  sensitivity,  and,  indeed,  NADH-cytochrome  c  reductase  is  usual- 
ly even  more  susceptible,  often  being  completely  inhibited  by  concentra- 
tions of  0.001-0.01  mM.  One  is  thus  tempted  to  attribute  the  inhibition 
of  various  oxidations  by  the  mercurials  to  an  action  early  in  the  electron 
transport  chain,  at  least  pre-cytochrome.  Furthermore,  Barron  and  Singer 
(1945)  had  reported  that  the  oxidation  of  reduced  cytochrome  c  by  a  cyto- 
chrome oxidase  preparation  is  not  affected  by  p-MB,  and  this  has  more 
recently  been  observed  with  Arum  (Simon,  1957)  and  Penicillium  (Sih  et 
al.,  1958)  cytochrome  oxidases.  Finally,  some  have  found  that  certain  oxi- 
dases and  the  corresponding  dehydrogenases  are  inhibited  equally  by  mer- 
curials, although  the  different  conditions  of  testing  in  such  cases  make  ac- 
curate comparison  difficult. 

This  simple  picture  of  inhibition  in  the  electron  transport  sequence  has, 
however,  been  questioned  by  workers  at  the  Institutum  Divi  Thomae,  who 
from  1946  to  1957  obtained  increasing  evidence  that  the  cytochrome  system 
may  not  be  as  immune  to  mercurials  as  generally  imagined.  Their  results 
may  be  summarized  in  the  four  following  categories.  (1 )  Cytochrome  oxidase 
activity,  as  determined  by  the  oxidation  of  ascorbate,  is  inhibited  rather 
potently,  50%  depression  being  observed  with  0.006-0.012  mM  PM  and 
0.032  mM  p-MB  (Cook  et  al,  1946;  Kreke  et  al,  1950).  (2)  Succinate  oxi- 
dase is  much  more  sensitive  to  mercurials  than  is  succinate  dehydrogenase. 
It  requires  around  10  times  the  concentration  to  inhibit  rat  heart  succinate 
dehydrogenase  compared  to  the  oxidase  (Cook  et  al,  1946;  Kreke  et  al, 
1949;  Smalt  et  al,  1957).  (3)  The  inhibitions  of  cytochrome  oxidase  and 
succinate  oxidase  are  not  reversed  by  thiols  (Cook  and  Perisutti,  1947; 
Kreke  et  al,  1949,  1950).  This  led  them  to  suppose  that  the  inhibition  might 
not  involve  SH  groups,  but  this  conclusion,  as  we  have  seen,  is  not  valid. 
(4)  No  evidence  for  reaction  of  the  mercurials  with  ascorbate  or  cytochrome 
c  could  be  obtained  by  spectroscopic  or  preincubation  techniques  (Cook  et 
al,  1946;  Kreke  et  al,  1950).  It  may  also  be  mentioned  that  Boeri  and  Tosi 
(1954)  found  no  reaction  of  p-MB  with  cytochrome  c,  and  that  Strittmatter 
and  Velick  (1956)  likewise  found  no  change  in  microsomal  cytochrome  spec- 
tral absorption  after  incubation  with  1  mM  p-MB.  All  of  these  data  have 
been  interpreted  as  indicating  that  the  mercurials  may  exert  a  major  part 
of  their  effect  on  cytochrome  oxidase. 


ELECTKON    TRANSPORT  871 

Slater  (1949)  had  also  observed  that  succinate  oxidase  is  inhibited  more 
strongly  than  the  dehydrogenase  by  p-MB  (and  also  by  o-iodosobenzoate 
and  oxidized  glutathione),  although  the  difference  was  not  as  great  as  re- 
ported by  Cook,  Kreke,  and  their  co-workers,  and  attributed  this  in  the 
particulate  preparations  used  to  an  effect  on  some  link  between  the  dehy- 
drogenase and  the  oxidase,  presumably  occurring  before  cytochrome  c  in 
the  chain.  This  effect  might  be  a  structural  disorganization  of  the  complex 
to  interrupt  the  flow  of  electrons.  Nevertheless,  Slater  observed  some  inhibi- 
tion of  cytochrome  oxidase.  Seibert  et  al.  (1950)  made  a  solubilized  deoxy- 
cholate  preparation  of  cytochrome  oxidase  and  found  by  both  manometric 
and  spectrophotometric  tests  that  it  is  inhibited  to  the  same  degree  as  the 
crude  preparation:  they  also  demonstrated  shifts  in  the  spectral  bands  of  the 
oxidase  following  treatment  with  the  mercurials.  The  final  conclusion  of  the 
Institutum  Divi  Thomae  group  is  that  the  actions  of  the  mercurials  on  heme 
enzymes  may  be  nonspecific,  may  involve  denaturation  (which  could  ac- 
count for  the  spectral  shifts),  and  do  not  involve  SH  groups,  but  I  doubt 
if  there  is  sufficient  evidence  for  any  of  these  statements.  However,  their 
data,  which  are  definite  and  consistent,  must  be  explained  on  some  basis. 
It  is  important  to  realize  that  the  inhibitions  reported  for  "cytochrome 
oxidase"  were  all  obtained  with  ascorbate  (and  occasionally  hydroquinone) 
as  the  substrate.  Now  neither  ascorbate  nor  hydroquinone  is  oxidized  di- 
rectly by  cytochrome  oxidase  and  the  electron  transfer  occurs  through  a 
series  of  components.  It  has  usually  been  assumed  that  ascorbate  reduces 
cytochrome  c^  or  c,  in  which  case  the  action  of  the  mercurials  could  be  on 
some  component  or  link  in  the  cytochrome  sequence,  rather  than  on  cyto- 
chrome oxidase  itself.  It  will  be  remembered  that  the  work  quoted  at  the 
beginning  of  this  section  showed  that,  when  cytochrome  c  is  used  as  sub- 
strate, the  mercurials  do  not  inhibit.  Is  it  possible  that  there  is  a  com- 
ponent which  might  be  designated  as  ascorbate  dehydrogenase,  which  is 
sensitive  to  the  mercurials?  Seibert  et  al.  (1950)  actually  observed  relative- 
ly little  inhibition  of  the  purified  system  when  determined  spectrophoto- 
metrically  with  cytochrome  c  as  the  substrate. 

There  are  several  ways  of  explaining  the  differential  inhibitions  of  suc- 
cinate dehydrogenase  and  oxidase.  Since  the  activities  of  these  two  sys- 
tems are  determined  very  differently  —  the  dehydrogenase  usually  by 
methylene  blue  reduction  and  the  oxidase  manometrically  —  one  must 
question  if  this  could  be  responsible  for  the  different  sensitivities  observed. 
The  dehydrogenase  activity  associated  with  methylene  blue  reduction  might 
not  be  exactly  the  same  as  in  the  normal  transfer  of  electrons  to  the  cyto- 
chromes; i.e.,  a  region  of  the  enzyme  surface,  or  another  component  in  the 
chain,  might  be  involved  in  the  normal  transfer  but  not  in  the  dye  reduc- 
tion, and  this  part  of  the  system  could  be  sensitive  to  the  mercurials.  If 
we  look  into  the  details  of  the  procedures  (Kreke  et  al.,  1949),  we  find  that 


872  7.   MERCURIALS 

in  the  dehydrogenase  test  the  pH  was  7.2  and  the  succinate  concentration 
0.33  rciM,  whereas  in  the  oxidase  test  the  pH  was  7.4  and  the  succinate 
88  mM;  the  former  was  done  in  strong  phosphate  buffer,  whereas  the  latter 
medium  contained  0.7  mM  Ca++  and  A1+++;  in  addition,  the  times  for 
equilibration  and  incubation  were  different.  When  the  conditions  are  so 
diverse,  it  is  impossible  to  compare  these  two  systems  quantitatively.  The 
structural  interference  theory  of  Slater  must  also  be  considered  and  has  as 
much  evidence  as  the  other  explanations  (i.e.,  none).  It  would  be  important 
to  know  just  how  much  effect  mercurials  can  exert  on  the  cytochrome  sys- 
tem, inasmuch  as  it  has  obvious  bearing  in  considerations  of  the  actions 
on  various  oxidations,  mitochondrial  systems,  and  respiration. 

A  comparable  situation  with  NADH  dehydrogenase,  NADH:  cytochrome 
c  reductase,  and  NADH  oxidase  has  been  noted  by  Minakami  et  al.  (1963). 
The  total  oxidase  and  the  cytochrome  c  reductase  are  very  sensitive  to  p-MB 
whereas  the  dehydrogenase,  as  determined  by  ferricyanide  reduction,  is  not 
as  sensitive.  It  was  postulated  that  two  types  of  SH  group  are  involved  in 
NADH  oxidation,  one  readily  accessible  to  mercurials  and  functioning  be- 
tween the  dehydrogenase  active  site  and  the  distal  respiratory  chain  (this 
SH  group  is  not  required  for  ferricyanide  reduction),  and  a  second  con- 
cealed in  the  dehydrogenase  complex  as  isolated,  and  exposed  on  degrada- 
tion to  the  cytochrome  c  reductase.  Such  an  explanation  could  apply  to  the 
succinate  oxidase  as  well,  as  was  suggested  above  relative  to  methylene 
blue  as  an  acceptor  for  the  determination  of  dehydrogenase  activity. 

Oxidative  Phosphorylation 

The  results  summarized  in  Table  7-15  show  that  mercurials  are  not  par- 
ticularly specific  or  effective  uncouplers  of  oxidative  phosphorylation  in 
mitochondria,  but  that  a  fair  degree  of  uncoupling  can  occur  under  certain 
circumstances.  It  is  especially  interesting  that  high  toxic  doses  of  the  mer- 
curial diuretics  and  HgClg  can  often  reduce  the  P  :  0  ratio  in  the  mitochon- 
dria of  excised  kidneys  several  hours  after  the  administration,  without  sim- 
ultaneously affecting  oxidative  phosphorylation  in  the  liver,  but  this  is 
undoubtedly  due  to  the  higher  concentration  of  mercurial  in  the  kidney. 
The  P  :  0  ratio  is,  however,  not  altered  significantly  by  the  ordinary  diuretic 
doses,  so  that  it  is  questionable  if  this  action  is  related  to  diuresis.  I  know 
of  no  instance  in  which  the  mercurials  augment  O2  uptake  while  simul- 
taneously reducing  the  P,  incorporation,  so  that  they  are  not  true  uncou- 
plers in  the  same  sense  as  the  nitrophenols. 

The  Pj^^-ATP  exchange  is  quite  potently  inhibited  by  mercurials  in  the 
mitochondria  obtained  from  mosquitoes  (Avi-Dor  and  Gonda,  1959),  pig 
liver  (Chiga  and  Plant,  1959),  and  rat  liver  (Plaut,  1957;  Cooper  and  Leh- 
ninger,  1957;  Lehninger  et  al.,  1958;  Low  et  al.,  1958).  For  rat  liver  mito- 
chondria the   exchange  is  sometimes   inhibited  50%    by   concentrations 


ELECTRON    TRANSPORT 


873 


P5 


^-  lO      05 


W   d   W 


"V    lo 


^ 

OJ 

Ti 

~^ 

a 

B 

cS 

■^ 

(4 

03 

C 

O 

3 

W   o 


>o 

o 

0) 

^^ 

05 

O 

JH 

o 

t-H 

05 

o 

CO 

^■^ 

c 

C5 

O 
o 

o 

2 
S 

'e 

'-s 

c» 

-o 

"S 

TJ 
c 

-^ 

S 

>» 

c3 

S 

M 

0) 

o 

o 

o 

S 

o 

M 

< 

fLi 

(M 

o 

d 

d 

A 

1 

t 

CO 

CO 

so 

fO 

^  p 


t  t    t 


IM      ^         O 


CD 

05 

t^ 

o 

O 

o 

-^ 

t 

t 

0) 

t- 

lO 

o 

lO 

lO 

"A 

d 

d 

i<      (M 


"2       fl       !-■ 
O       QJ       C 

O    2  ^ 


be    M 

ho 

bC 

'^  =^ 

^ 

^ 

'm  ^ 

"bC 

^ 

so 
so 

CD 

3 

ffi    W 

W 

w 

3 

s 

-a 

bC    bfi 

2   S 

bc 

s 

bC 

g 

^ 

d    d 

o 

t- 

a 

o 

«<i 

Q 

i>  t- 

o 

so 

o 

d  d 

(M 

d 

-H     (N 

m 

m 

pq 

+ 

+ 

^ 

S 

<D 

S 

M 

+ 
bC 

4 

?s, 

S 

i. 

ffi 

W 

w  § 


m      pq 


Q  c3       c3     i3     K     'o  3 

^  a    g    O    '=a  CO      Ph 


5  w 


Pm 


pq 


874  7.  MERCURIALS 

around  0.002  mM,  and  completely  by  0.01  mM.  The  ADP^s-ATP  and 
ADP-C^*-ATP  exchanges  are  also  inhibited,  but  perhaps  not  as  strongly 
(Wadkins  and  Lehninger,  1958;  Chiga  and  Plant,  1959;  Kahn  and  Jagen- 
dorf,  1961).  These  exchange  reactions  are  related  intimately  with  oxi- 
dative phosphorylation.  Indeed,  Wadkins  and  Lehninger  (1958)  postulated 
that  the  Pj-ATP  exchange  is  a  measure  of  the  two  terminal  reactions  in 
oxidative  phosphorylation: 

Carrier  — ^  X  +  Pi  ^  carrier  +  P  ^ — ^  X 
P  —  X  4-  ADP  ±5  ATP  +  X 

where  X  is  perhaps  the  enzyme  protein,  while  the  ADP- ATP  exchange 
measures  only  the  last  reaction.  They  further  suggest  that  the  mercurials 
inhibit  this  last  reaction  principally,  whereas  2,4-dinitrophenol  acts  on  the 
penultimate  step.  If  the  mercurials  act  solely  on  the  transfer  of  phosphate 
to  ATP  they  would  be  good  uncouplers,  but  actions  elsewhere  in  the  elec- 
tron transport  chain  limit  their  efficiency.  It  is  interesting  that  Griffiths 
and  Chaplain  (1962)  have  found  evidence  for  a  new  phosphorylated  deriva- 
tive of  NAD  following  incubation  of  heart  mitochondria  with  succinate 
and  P(^2.  ATP  can  be  formed  from  the  intermediate  and  this  reaction  is 
completely  blocked  by  p-MB  at  0.01  mM.  This  observation  is  compatible 
with  the  scheme  of  Wadkins  and  Lehninger. 

FERMENTATION  AND  GLYCOLYSIS 

The  first  impression  from  surveying  the  studies  of  mercurial  action  on 
fermentation  and  glycolysis  is  that  these  pathways  are  often  surprisingly 
insensitive  to  this  group  of  inhibitors.  In  many  cases  it  requires  concentra- 
tions greater  than  1  mM  to  depress  glycolysis  significantly  in  cellular  sys- 
tems. Reference  to  Table  7-13  shows  that  several  enzymes  in  the  Embden- 
Meyerhof  pathway  are  quite  readily  inhibited  by  mercurials,  e.g.,  hexo- 
kinase,  aldolase,  3-phosphoglyceraldehyde  dehydrogenase,  enolase,  and  lac- 
tate dehydrogenase,  concentrations  of  0.001-0.05  mM  usually  inhibiting 
50%  or  more  in  muscle,  although  little  is  known  about  the  sensitivities  of 
the  yeast  enzymes.  Since  several  enzymes  in  the  pathway  are  susceptible, 
one  might  anticipate  that  the  sequential  inhibition  by  mercurials  at  con- 
centrations above  0.05  mM  would  produce  a  very  strong  over-all  depression 
of  anaerobic  COg  or  lactate  formation.  Three  explanations  for  the  failure  to 
do  so  are  immediately  apparent:  (1)  The  mercurials  do  not  penetrate  into 
the  cells  readily;  (2)  the  glycolytic  enzymes  are  protected  in  the  cell  (e.g., 
by  substrates  or  coenzymes);  and  (3)  the  large  amount  of  nonenzymic  ma- 
terial in  cellular  preparations  binds  much  of  the  mercurial.  Cleland  (1949) 
found  that  1  mM  PM  inhibits  glycolysis  in  oyster  eggs  only  17%  at  0-45 
min  and  48%  at  45-90  min,  whereas  glycolysis  in  egg  homogenate  (with 


FERMENTATION  AND   GLYCOLYSIS  875 

glycogen,  ATP,  and  NAD  added)  is  completely  blocked.  Similar  results 
were  obtained  with  sea  urchin  eggs  by  Cleland  and  Rothschild  (1952  a), 
1  mM  PM  inhibiting  lactate  formation  17%  in  whole  eggs  and  97%  in 
extracts.  These  results  were  interpreted  to  indicate  poor  penetration  by  the 
mercurial,  but  the  other  two  explanations  given  above  are  probably  as 
likely,  and  undoubtedly  all  contribute  to  some  extent.  It  may  be  noted  that 
Cleland  found  endogenous  respiration  to  be  inhibited  more  potently  in 
whole  eggs  than  in  homogenates,  which  is  more  difficult  to  explain.  There 
are  instances  of  cellular  glycolysis  quite  sensitive  to  the  mercurials;  in  asci- 
tes carcinoma  cells  there  is  50%  inhibition  by  0.0032  mM  Hg++  (Schom 
et  at.,  1961).  The  glycolysis  in  spleen  slices  is  also  fairly  sensitive,  although 
it  requires  4-5  hr  to  reach  maximal  inhibition  (Fig.  1-12-24)  (Jowett  and 
Brooks,  1928).  The  interesting  questions  of  the  penetration  of  mercurials 
and  the  effects  exerted  on  cell  membranes  will  be  considered  later  (page  892). 

Yeast  Fermentation 

Since  the  early  work  of  Schulz  (1888),  who  reported  an  initial  stimulation 
of  fermentation  by  low  concentrations  of  Hg++  (0.005-0.008  mM)  and  inhi- 
bition by  higher  concentrations  (>  0.02  mM),  there  have  been  many  stud- 
ies of  yeast  fermentation  with  variable  results.  The  stimulation  observed 
by  Schulz  has  seldom  been  confirmed.  Joachimoglu  (1922)  could  never  dem- 
onstrate acceleration  of  CO2  formation  by  Hg++,  concentrations  of  0.0031- 
0.037  mM  exerting  no  effect  and  0.074  mM  inhibiting  around  70%.  Meier 
(1926)  found  even  more  potent  inhibition  of  aerobic  fermentation,  0.009  mM 
Hg++  depressing  72%,  while  Kostytschew  and  Berg  (1930)  never  observed 
stimulation,  inhibition  beginning  at  0.0185  mM  Hg++  and  reaching  42% 
at  0.2  mM.  More  recently,  some  have  found  potent  inhibition  by  Hg++ 
(e.g.,  Hurwitz  and  Chaffee,  1954),  but  others  have  not  (e.g.,  Weitzel  and 
Buddecke,  1959);  in  the  latter  work,  1  mM  Hg++  inhibited  only  70%  in 
fermenting  yeast.  Organic  mercurials  have  not  been  often  used,  but  Spiegel- 
man  et  al.  (1948)  reported  69%  inhibition  by  0.01  mM  PM  and  32%  inhi- 
bition by  0.05  mM  p-MB,  indicating  these  mercurials  to  be  fairly  effective. 
Certainly  much  of  the  variation  in  the  results  is  due  to  the  different  densi- 
ties of  yeast  suspension  used,  the  media  employed,  and  the  state  of  the 
yeast  (by  which  is  meant  its  fermentative  activity  and  past  history).  One 
would  expect  mercurials  to  attack  surface  hexokinase  and  the  initial  phos- 
phorylation of  glucose,  as  occurs  in  muscle,  so  one  can  explain  the  exam- 
ples of  weak  inhibition  only  on  the  basis  of  relatively  dense  yeast  suspen- 
sions. 

Muscle  Glycolysis 

The  results  obtained  on  muscle  glycolysis  with  the  mercurials  have  been 
quite  inconsistent  and  even  more  difficult  to  explain  than  those  with  yeast. 


876  7.  MERCURIALS 

Gemmill  and  Hellerman  (1937)  found  that  Hg++,  p-MB,  and  PM  all  block 
glycolysis  in  extracts  of  frog  muscle,  but  the  concentrations  used  were  too 
high  and  usually  unspecified.  Separated  fibers  of  cockroach  muscle  treated 
with  1  mikf  p-MB  show  no  change  of  COg  formation  and  a  rather  marked 
increase  in  lactate  formation  if  only  glucose  is  added,  but  in  the  presence 
of  glucose  +  ATP,  COg  production  is  inhibited  67%  and  there  is  no  effect 
on  lactate  (Barron  and  Tahmisian,  1948).  This  behavior  is  quite  different 
from  that  of  iodoacetate,  which  inhibits  only  in  the  absence  of  added  ATP. 
The  authors  felt  that  the  failure  to  depress  lactate  formation  in  any  case 
is  perhaps  a  characteristic  of  invertebrate  muscle,  since  Harting  (1947)  had 
observed  1  mM  p-MB  to  produce  only  stimulation  of  glycolysis  in  strips  of 
scallop  and  thyone  muscle.  However,  Krueger  (1950)  has  shown  that  2  min 
perfusion  of  frog  muscle  with  37  mM  Hg++  essentially  doubles  the  lactate 
formation.  The  only  serious  study  of  muscle  glycolysis  was  done  by  Bailey 
and  Marsh  (1952)  on  rabbit  psoas  homogenates.  Here  p-MB  produces  def- 
inite inhibition  (see  accompanying  tabulation),  but  the  concentration  is  so 


Control 

p-MB  4  mM 

zIpH 

-  0.28 

-  0.08 

A  Fructose-diP 

+27 

+  6 

A  Triose-P 

+  3 

-   1 

zl  ATP 

-14 

-  5 

ATP  resynthesis 

45 

13 

ATPase  inhibition 

77% 

high  that  it  is  remarkable  that  the  inhibition  is  not  much  greater.  It  was 
believed  that  aldolase  inhibition  is  responsible  for  the  results  but  from  the 
data  it  is  not  possible  to  localize  the  site  of  action  so  closely.  The  authors 
pointed  out  that  3-phosphoglyceraldehyde  dehydrogenase  is  not  so  readily 
inhibited  by  p-MB  as  by  iodoacetate.  The  transfer  of  phosphate  from  crea- 
tine-P  to  ADP  is  immediately  and  completely  blocked  by  p-MB,  so  that 
creatine-P  remains  at  its  initial  level,  and  this  must  also  be  a  factor  in  the 
inhibition,  since  it  would  prevent  regeneration  of  ATP.  It  is  thus  impossible 
in  this  study  to  determine  what  effects  p-MB  might  have  directly  on  the 
Embden-Meyerhof  pathway.  All  of  the  results  on  intact  muscle  tissue  seem 
to  be  incompatible  with  the  demonstration  by  Demis  and  Rothstein  (1955) 
that  glucose  uptake  by  diaphragm  is  very  sensitive  to  Hg++,  being  almost 
completely  inhibited  by  0.2  mM.  However,  respiration  and  anaerobic  lac- 
tate formation,  being  dependent  on  endogenous  substrate,  are  much  less 
sensitive  and  are  only  slowly  inhibited.  This  will  be  considered  in  greater 
detail  when  the  effects  of  mercurials  on  respiration  are  discussed  (page  884). 


TRICARBOXYLATE  CYCLE  877 

Stimulation  of  glycolysis  by  the  mercurials  is  not  confined  to  yeast  and 
muscle.  Hg++  below  0.11  mM  stimulates  glycolysis  in  guinea  pig  blood  and 
inhibits  in  higher  concentration  (Fuentes  and  Rubino,  1923),  while  in  hu- 
man blood  Hg++  stimulates  anaerobic  glycolysis  from  0.0185  to  1.85  vaM 
although  at  18.5  mM  there  is  almost  complete  inhibition  (Rubino  and 
Varela,  1923).  Glucose  utilization,  COg  release,  and  lactate  formation  in 
human  erythrocytes  are  all  stimulated  by  p-MB  up  to  5  //moles/ml  of 
erythrocytes  (Jacob  and  Jandl,  1962).  No  explanation  for  these  results 
is  immediately  evident. 

TRICARBOXYLATE  CYCLE 

Despite  the  fact  that  no  analysis  of  the  effects  of  mercurials  on  the  cycle 
or  on  the  operation  of  mitochondria  has  been  made,  one  would  predict 
quite  potent  inhibition  on  the  basis  of  the  sensitivities  of  the  individual 
enz\Tnes  (Table  7-13).  Mercurial  concentrations  in  the  neighborhood  of 
0.01  mM  should  depress  several  enzjTnes  very  significantly  (e.g.,  pjTuvate 
oxidase,  isocitrate  dehydrogenase,  or-ketoglutarate  oxidase,  succinate  de- 
hydrogenase, malate  dehydrogenase,  and  some  ancillary  enzymes,  such  as 
acetate  kinase),  and  concentrations  of  the  order  of  0.1  mM  should  block 
completely.  However,  since  we  have  already  noted  that  glycolysis  is  often 
not  inhibited  as  much  as  one  would  expect  from  studies  of  the  individual 
enzymes,  we  must  be  very  careful  in  considering  inhibitions  of  the  cycle 
in  cellular  preparations.  The  utilization  of  pyruvate  and  acetate  by  a  va- 
riety of  cellular  and  subcellular  preparations  has  been  shown  to  be  readily 
inhibited  by  mercurials  (Table  7-16),  but  in  no  case  was  the  operation  of 
the  entire  cycle  tested,  so  that  the  entire  inhibition,  as  far  as  one  knows, 
might  be  on  the  initial  enzyme  reaction  (p>Tuvate  oxidase  or  acetate  kinase). 
If  the  cycle  is  operating  by  regenerating  oxalacetate,  much  stronger  inhi- 
bition would  undoubtedly  be  observed.  In  work  with  mitochondria,  homo- 
genates,  or  cell  suspensions,  however,  one  must  always  remember  the  role 
of  nonenzyme  protein  in  reducing  the  mercurial  available  for  inhibition, 
so  that  concentrations  such  as  those  in  Table  7-16  are  not  of  much  quanti- 
tative significance,  but  show  definite  interference  with  cycle  activity. 

In  order  to  answer  some  of  these  questions  relative  to  the  action  of  the 
mercurials  on  the  cycle,  Dr.  Yang  kindly  consented  to  examine  the  effects 
of  Hg++  on  the  Oj  uptake  of  rabbit  heart  mitochondria  by  the  same  tech- 
niques used  in  a  previous  study  of  iodoacetate  (Yang,  1957).  The  changes 
over  a  60  min  period  obtained  from  results  on  three  preparations  are  shown 
in  the  accompanying  tabulation.  These  data  show  clearly  that  several  steps 
in  the  cycle  are  inhibited  rather  strongly  as  the  concentration  is  raised  from 
0.003  mM  to  0.01  mM,  and  that  at  0.1  mM  the  cycle  activity  is  essentially 
completely  blocked.  The  stimulation  observed  with  a-ketoglutarate  as  the 


878 


7.  MEKCURIALS 


A 


X 

« 

•— N 

-i-= 

•^ 

eS 

^^ 

PL| 

-tJ 

01 

b^ 

W. 

O 

!?; 

o 

H 

5 

3 

iz 

Ph 


m 


d   d 


M 


W 


fL,   Ph 


tiq 


^-     IC 


03     OJ      OJ 


o 


bn 

-d 

C 

CI 

r/j 

r! 

-o 

o 

C 

fcH 

p  c:  ^ 


05        '-' 


O    — 


c3  f^ 


c 

,^ 

crt 

SU 

<u 

S 

0) 

01 

ffi 

« 

d   d 


d   d 


o  o 
d  d 


O    -H    (M  lO 

d   d   d  d 


P5  m 


?!,    ?S, 


m  P5 
9,  a 


P3   P5   PP 


55.    ?5, 


(Coo 


P4 


PM 


c   .ti 


o)   ;^   .« 


a   =   ^ 


O   O 


^ 


A      A 


S  § 


RESPIRATION  879 


°o  Changes  from  Hg++  at: 


0.001  mM 

0.003  mM 

0.01  ml/ 

0.1 

mM 

Pyruvate  +  malate 

-10 

-79 

99 

a-Ketoglutarate 

+24 

+5 

-74 

- 

100 

Succinate 

0 

+5 

-41 

— 

100 

Malate 

-   6 

— 

-79 

— 

99 

substrate  seems  to  be  real  since  it  was  consistently  obtained.  The  inhibition 
with  pyruvate  +  malate  as  substrates  and  Hg++  at  0.01  vcvM  is  about  50% 
at  10  min  and  then  increases  more  slowly  until  it  is  100%  at  60  min.  The 
figures  in  the  tabulation  are  mean  inhibitions  over  the  60  min  period  and 
even  at  0.01  milf  the  activity  was  almost  all  lost  in  aU  cases  by  60  min. 


RESPIRATION 

The  effects  of  the  mercurials  on  the  0.,  uptake  of  tissues  vary  considerably 
and  depend  on  the  mercurial  used,  the  substrate,  the  pH,  the  species,  and 
many  other  factors  (Table  7-17).  One  factor  about  which  little  is  known, 
but  which  could  be  very  important,  is  the  thickness  of  the  tissue  when  the 
preparation  is  a  strip,  section,  or  slice,  inasmuch  as  the  mercurial  possibly 
does  not  penetrate  equally  throughout  but  acts  primarily  on  the  outer 
layers  of  cells.  Cascarano  and  Zweifach  (1962)  examined  rat  diaphragm 
after  exposure  to  Hg+"^  by  determining  the  ability  of  the  tissue  to  reduce 
a  tetrazolium  dye,  and  found  that  only  a  well-defined  band  of  surface  fibers 
had  lost  the  ability,  the  central  portions  retaining  activity.  Measurements 
of  respiratory  inhibition  in  such  cases  do  not  provide  true  values  (see  page 
1-479);  in  the  extreme  case  the  inhibition  may  relate  only  to  the  fraction 
of  the  tissue  affected,  and  progressively  developing  inhibition  may  exhibit 
time  relations  dependent  only  on  the  rate  of  penetration  through  the  tissue. 
This  would  apply  not  only  to  respiration,  of  course,  but  to  all  measurements, 
metabolic  or  functional,  made  on  all  intact  tissues.  Failure  to  reach  all  of 
the  cells  equally  must  be  one  reason  for  the  low  degree  of  inhibition  often 
observed,  lower  than  would  be  predicted  from  the  effects  on  glycolysis, 
the  cycle,  and  the  enzymes  involved. 

One  notes  several  examples  wherein  respiration  is  stimulated  by  the  mer- 
curials, more  often  at  low  concentration  but  in  one  instance,  yeast  respir- 
ing endogenously  (Shacter,  1953),  the  stimulation  appears  only  at  high 
concentrations  of  1-2  vaM.  There  are  other  reports  of  stimulation  not  in- 
cluded in  the  table.  For  example,  Gremels  (1929)  found  that  when  mersalyl 
induces  diuresis  in  a  heart-lung-kidney  preparation,  the  kidney  respiration 


880 


7.  MERCURIALS 


P5 


P5 


t) 

O     £ 

o 

O    - 

B3 

H 

S 

w 

H 

ft 

w 

H 

>^ 

, 

m 

.2 

'C 

^ 

3 

o 

o 

M 

tH 

fi 

« 

^-.  ■<* 


CO  c 

lO  t3 

05  ^ 

3  S 

^  eS 


o  00 

O    O 


o   02   ffi   o   ^2; 


o   ^ 


O    o  o    ® 


O      O      lO 


o 
^    d 


M  Mia 
w  w  H 


K 


o  o 


CO      CO      n      OS 


s 

OO 

00 

s 

c 

£? 

p 

S      3 
02     02 


3      3 
02     02 


SO 

•<^ 

cb 

1— 1 

>-^ 

0) 

-«  ^ 

,,— .s. 

g  M 

CO 

^2 

§ 

73    ^ 

^-^ 

S      ^ 

bl 

CO     fli 

<D 

^^ 

+s 

J<      o 

O 

O     eS 

eS 

O    JH 

_Q 

O   02 

OJ 

o  >o 

■* 

O 

o 

S  <=> 

05 

ec 

I— 

o 

(M    ■* 

Tjf    O      O      O      iM      to 
t^    t>      t^      l>      t^      tJ< 


05 

o 
o 

00 

o 

CO 

o 

CO 

o 

o 

9 

,^ 

^ 

lO 

© 

d 

d 

d 

d 

d 

d 

d 

-H  d 

d 

to 

1 

PQ 

+     + 

S 

1-1    +     + 
^     bc     tti 

a 

(iH    W    W 

(1h   W 


O 


PQ 


Pm   sj. 


0)  a> 
n  oo 
O    O 


o  o 


0)        V        V        0) 

ft    p-    Oh    a 


SQ 


oq    cq    ft?    rq    S-, 


RESPIRATION  881 


"^ 

0 

-§ 

<B 

eS 

t8 

ffi 

c 

S 

05 

C 
<3 

05 

05 

§ 

2 

X 
35 

.2 

OS 

TS 

"»* 

5 

— H 

C5 

^~^ 

M 

^..^ 

B 

CO 
05 

' — ' 

■ — ' 

4^ 

2 

c3 

2 

"» 

"5 

;^ 

d- 

>. 

ft 

■3 

ki 

ft 

g 

0) 

C5 

0 

4J 
"ft 
ft 

a 

o 

CO 

0 

iC 

t- 

_^ 

0 

>c 

10 

0 

0 

<5 

0 

0 

^ 

»o 

(-J 

0 

0 

0 

0 

0 

0 

5<1 

0 

0 

CO 

X 

C5 

05 

05 

M 

-f3 

0 

CO 

«<1 

t- 
■* 

X 

CO 

X 

10 

Tj< 

X 

cq 

-f9 
00 

-t3 

00 

02 

CO 

C<I 

-^ 

c 

0 

0 

0 

0 

0 

-^ 

0 

0 

0 

0 

-H 

CO 

CO 

>c 

IM 

— < 

— < 

-H 

- 

- 

d 

— 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

-' 

d 

d 

-^ 

d 

^ 

o 

o 

0 

0 

10 

»o 

IC 

K5 

10 

<M 

<N 

10 

1 

1 

i> 

t^ 

t- 

t> 

d 

d 

d 

CO 

d 

X 

X 

t- 

+   +   +  t   +   +  "^  pQCQpamea 


V  V  <»  « 

09  -<5  K  CO 

^oS^^oo  o  0000 


jg 


ft 

a 
0 

g 

e8 

s 

0 

c 

?^ 

^ 

^ 

& 

3 

c 

3 

CC 

02 

» 

ss 

s 

5 

C 

« 

n 

c 

C! 

bO 

n 

eS 

oc 

a> 

C 

ai 

<a 

73 

iS 

ft 
S 

S 
0 

fa 

OQ 

^s< 

H 

!» 

!rt 

■«» 

w 

=0 

1. 

aj 

^ 

•^ 

"» 

s 

^ 

e 

e 

1 

5» 

® 

0 

§ 

? 
0 
« 

i ' 

0 

n 

0 

a 

0 

3 

a. 

t*. 

^ 

882 


7.  MERCURIALS 


tf 


^  o  S 


Ol 

13 

2 

o 

3 

b 

^ 

^^ 

N 

0) 

o 

GO 

m 

X. 

o 

C5 

'9- 

a 

w 

S 

""rt 

■c 

'e 

cS 

c 

c 

>. 

cS 

e3 

"S 

S 
o 

o 

2 

c 

b 

3 

0^ 

cS 

s 

^ 

O 

m 

^  Ci  05 

'  f-H      ^^      *44 


2      ^ 


o 


—- 

■e 

!0 

r. 

"o 

"S 

:;^ 

"o 

cS 

L^ 

^-^ 

-C 

g 

c 
1 

c 

03 
4) 

o 

o 


C 


oo  o  00  t-  o  00 

r^   o   o   >c   o   GO 


o    S 


O      — '  -H 


o  -^  o  — 
d  —  d  d  d  d  — ' 


^  d  d  -H  o  d 


d  d  -<  d  —  d 


O    ■*    lO    (M 
;C     I>     t^     GO 


PQ  pq 

^  ^  ^ 
?5,    ?5,  (In 


Ph 


Ph   P-( 


P5  pp  + 

+    + 

pq 

^    S     M 

+    + 

§ 

a  Si,  W 

ffi    W 

?>, 

pp 


§   S   g   £ 

-S  ^    o    o 

a  o  ;^  ;^ 


a> 

0) 

0; 

aj 

(D 

aj 

(» 

<D 

C 

c 

c 

c 

c 

C 

c 

a 

o 

o 

o 

o 

o 

o 

o 

o 

;?; 

^ 

f^ 

^ 

^ 

;^ 

"^ 

^ 

ft       ;= 


PL, 


e8 

C 

c 

.2 

Oj      g 

02 

.2 

_S 

M     O 
O      N 

C    ^j 

ft 

5 

C 
4) 

O     c6 

to 
3 
to 

ft 

ft 

^   a3 

to 

CO 

M 

3 

^     ft 

6C 

CB 

M 

W   M 

H 

;«s» 

00 

5 
^ 

5S 

Si. 

o 

Co 

=0 

S 

-2 

=0 

.e 

1 

o 

•(» 

H 

-< 

-o 

J^ 

O 

V) 

i« 

^ 

^ 

t^ 

'^^ 

o 


c3 

ce 

ft 
2 

»c 

c4 

^ 
B. 

M 

o 

ft 

o 

<a 

O 

S 

to 

-o 

> 

S 

+3 

c« 

CC 

-S 

c« 

o 

<A 

to 

J3 

« 

0^ 

QJ 

_® 

1 

M 

c 

o 

•'^ 

"o 

"o 

3 

-3 

4^ 

« 

c 

o 

j5 

j3 

to 

2 

be 

ft 

% 

1 

^ 

^ 

'2, 

o3 

e 

e 

'^  E>q  ^    *  ^  p^ 


RESPIRATION  883 


C5  CO 


C     =3 


O    — '  IC  X    S<l    o 

ooooo  —  co'occo-^oc;  —  L-:  —  oM»^oo 


t^   I~-   t^ 


c  c 


O  O  O                HH 

2  "  "                              S-,                                         ^                                                    r- 

o  o  c      =52.2;             ;z;             o  o      c  <          o      o 

EC 


"TS  02 


>  :2 


M  O  Hj;4  ^  QPKffi  W 


p  O  e8  *  ^ 


884  7.  MERCUEIALS 

is  increased.  There  has  been  no  study  of  the  mechanism  whereby  such  stim- 
ulation is  produced.  One  might  postulate  that  the  mercurials  can  increase 
membrane  permeability  so  that  substrates  can  enter  cells  more  rapidly,  but 
although  this  may  be  a  factor  it  is  clear  that  endogenous  respiration  can  be 
stimulated,  as  in  the  work  of  Shacter.  It  is  also  known  that  subcellular 
preparations,  and  indeed  certain  enzymes  themselves,  are  stimulated  (page 
815),  so  that  one  cannot  expect  to  provide  a  comprehensive  theory  based 
only  on  cell  and  tissue  responses.  Shacter  felt  that  the  mercurials  might 
react  with  certain  thiols  which  regulate  metabolism,  but  despite  all  the 
discussion  of  such  regulators  in  the  past,  there  seems  to  be  little  evidence 
at  present  for  their  importance.  Of  the  several  mechanisms  suggested  pre- 
viously (page  1-453),  one  is  at  a  loss  to  select  any  that  would  apply  partic- 
ularly to  the  mercurials.  Since  mercurials  have  been  shown  to  reduce  the 
P :  0  ratio  in  various  isolated  systems,  it  is  possible  that  in  the  cell  an  un- 
coupling action  might  increase  Og  uptake  in  a  certain  limited  range  of  con- 
centration, and  it  is  also  possible  that  the  mercurials  by  a  membrane  effect 
might  alter  ion  movements  and  concentrations,  thus  secondarily  bringing 
about  metabolic  stimulation,  but  there  is  no  direct  evidence  for  either  of 
these  mechanisms.  Intracellular  changes  are  undoubtedly  so  complex  that 
metabolic  effects  usually  defy  analysis.  Consider  the  situation  described  by 
Estler  et  al.  (1960)  in  yeast  treated  with  Hg++  (Fig.  7-34),  the  levels  of  all 
the  adenine  nucleotides  increasing  at  0.2  mM,  although  Og  uptake  is  scarcely 
affected,  while  at  higher  concentrations  the  nucleotides  change  in  a  com- 
plicated fashion  and  respiration  is  severely  depressed.  Unfortunately  stim- 
ulation was  not  recorded  here,  but  it  is  only  by  thorough  studies  of  this 
type  that  one  can  hope  to  penetrate  into  the  mysteries  of  inhibitor  stim- 
ulation. 

Although  Meier  (1926)  reported  that  aerobic  fermentation  in  yeast  is 
more  potently  inhibited  than  respiration  by  Hg++  —  at  0.009  mM  the 
former  was  inhibited  72%  and  the  latter  not  at  all  —  subsequent  work  on 
a  variety  of  cells  has  indicated  no  general  relationship  between  the  sensi- 
tivities of  glycolysis  and  respiration,  and  even  in  yeast  Weitzel  and  Bud- 
decke  (1959)  found  both  to  be  inhibited  similarly,  at  least  at  high  concentra- 
tion (1  mM)  of  Hg++.  The  respiration  of  oyster  eggs  is  inhibited  more  than 
glycolysis  by  PM  (Cleland,  1949),  whereas  in  oyster  spermatozoa  the  res- 
piration is  inhibited  by  PM  when  glycolysis  as  measured  by  lactate  for- 
mation is  increased  (Humphrey,  1950).  The  utilization  of  pyruvate  in  these 
cells  may  be  more  sensitive  to  mercurials  than  the  glycolytic  pathway. 
Certainly  the  inhibition  of  respiration  does  not  imply  a  reduction  in  glu- 
cose uptake:  In  diaphragm,  1  mM  p-MB  depresses  Og  uptake  15%,  lowers 
the  glycogen  content,  and  yet  increases  glucose  utilization  somewhat  (Haft 
and  Mirsky,  1952).  In  most  cases  (e.g.,  yeast  and  Chlorella),  glucose  respira- 
tion is  more  sensitive  than  endogenous  respiration  to  mercurials,  but  this 


RESPIRATION 


885 


has  not  been  investigated  sufficiently  to  draw  valid  conclusions.  Of  course, 
in  tissues  such  as  most  muscle  and  heart,  in  which  endogenous  substrates 
are  responsible  for  the  bulk  or  all  of  the  respiration  for  some  after  time  ex- 
cision, glucose  would  not  be  expected  to  have  much  effect  on  the  inhibition 
by  mercurials.  Little  is  known  about  the  effects  of  mercurials  on  the  pen- 
tose-? pathway  or  other  routes  of  glucose  degradation.  In  crude  extracts 
of  Pseudomonas  converting  gluconate-6-P  to  3-phosphoglyceraldehyde  and 


12- 


09 


1000 


-  800 


0.6-1 


^  0.3  - 


-  600 


-400 


0  i-o 


Fig.  7-34.  Effects  of  Hg++  on  the  respiration  and  the 

levels  of  substances  in  yeast.  Concentrations  given  as 

/*moles/g  dry  weight.  Run  at  pH  6.8  for  60  min.  (From 

Estler  et  al.,  1960.) 


pyruvate,  p-MB  at  1  mM  inhibits  completely  (Kovachevich  and  Wood, 
1955).  The  site  of  the  inhibition  is  not  clear,  but  it  is  presumably  not  on 
the  gluconate-6-P  dehydrase,  which  was  purified  and  found  to  be  only 
moderately  sensitive  to  p-MB.  The  pentose-P  pathway  is  operative  in  ex- 
tracts of  tobacco  leaves,  the  oxidations  being  NADP  specific,  and  p-MB 
at  0.1  mM  almost  completely  blocks  the  reduction  of  NADP  by  both  glu- 
cose-6-P  and  fructose- 1,6-diP  (Clayton,  1959).  The  utilization  of  pentose-P 
by  extracts  of  Lactobacillus  hrevis  is  inhibited  only  20%  by  0.1  milf  p-MB 


886  7.   MERCURIALS 

(Eltz  and  Vandemark,  1960).  There  are  no  data  for  comparing  the  relative 
sensitivities  of  the  Embden-Meyerhof  and  pentose-P  pathways. 

As  appears  to  have  been  the  case  with  most  of  the  workers  who  have 
examined  the  effects  of  the  mercurials  on  respiration,  I  find  little  to  say 
that  seems  worthwhile.  The  most  interesting  aspects  of  respiratory  inhibi- 
tion probably  pertain  to  the  metabolic  basis  of  certain  cellidar  functions, 
e.g.,  gastric  secretion  (page  914)  or  renal  transport  (page  917).  The  site  of 
action  to  inhibit  respiration  is  unknown  and  multiple  sites  are  likely.  We 
know  little  of  the  penetration  of  mercurials  into  cells  and  the  intracellular 
concentrations  attained,  and  the  information  is  lacking  to  evaluate  the  im- 
portance of  nonenzymic  effects.  These  gaps  in  our  knowledge  apply  not 
only  to  respiration  but  essentially  to  all  cellular  activities.  One  can  at  least 
state  with  fair  certainty  that  the  mercurials  do  not  act  like  other  SH  re- 
agents, such  as  iodoacetate  or  the  arsenicals,  i.e.,  their  pattern  of  inhibition 
is  quite  different. 

VARIOUS   METABOLIC   PATHWAYS 

In  this  section  we  shall  consider  briefly  some  of  the  important  types  of 
metabolism  which  are  readily  inhibited  by  the  mercurials.  Only  the  more 
interesting  aspects  and  interpretable  investigations  will  be  mentioned.  The 
effects  of  the  mercurials  on  metabolism  are  complex  and  vague  in  all  cases, 
so  it  is  essential  to  emphasize  those  studies  in  which  clear-cut  results  have 
been  obtained,  even  though  the  work  is  limited  to  only  a  certain  phase  of 
the  over-all  pathway  and  the  exact  site  or  mechanism  of  action  is  unknown. 
The  few  systems  discussed  will  at  least  point  out  clearly  the  manifold  potent 
inhibitions  which  can  be  exerted  by  the  mercurials  and  will  serve  to  establish 
the  fact  that  specific  effects  on  metabolism  can  seldom,  if  ever,  be  achieved 
in  cellular  systems.  Perhaps  with  the  increasing  knowledge  of  the  detailed 
actions  of  the  mercurials,  there  will  arise  situations  in  which  selective  blocks 
can  be  produced  under  controlled  conditions,  but  at  the  present  time  there 
is  not  much  reason  for  optimism. 

Lipid  Synthesis 

The  long  sequence  of  reactions  in  the  biosynthesis  of  sterols  seems  to  be 
strongly  inhibited  by  mercurials  at  different  sites.  The  total  incorporation 
of  mevalonate-C^*  by  Lactobacillus  casei  over  4  hr  is  inhibited  59%  by  0.1 
mif  p-MB  and  96%  by  1  mM  (Thorne  and  Kodicek,  1962).  The  conversion 
of  farnesyl-PP  and  mevalonate  to  squalene  by  various  fractions  of  rat  liver 
is  depressed  50%  by  p-MB,  p-MPS,  and  Hg++  at  concentrations  near  0.05 
mM,  and  essentially  completely  by  concentrations  much  above  0.1  mM 
(Popjak  et  al.,  1958;  Anderson  et  al.,  1960;  Goodman  and  Popjak,  1969). 
The  further  conversion  of  squalene  to  sterols  is  97%  blocked  by  0.33  mM 


VARIOUS    METABOLIC    PATHWAYS  887 

p-MB  (Goodman,  1961).  The  sensitive  enzymes  are  probably  aU  located  in 
the  microsomes.  One  of  the  enzymes  on  the  pathway  from  mevalonate  to 
farnesyl-PP,  the  isopentenyl-PP  isomerase,  is  inhibited  completely  by  0.1 
vnM  p-MB  (Agranoff  et  al.,  1960),  so  this  could  account  for  part  of  the  block 
in  squalene  formation,  but  there  are  undoubtedly  other  sensitive  steps. 

Fatty  acid  biosynthesis  from  acetate  in  mammary  gland  homogenates  is 
inhibited  95%  by  0.1  mM  Hg++  (Popjak  and  Tietz,  1955),  and  from  acetyl- 
CoA  and  malonyl-CoA  in  purified  fractions  from  pigeon  liver  95%  by  0.075 
mM  p-MB  (Bressler  and  Wakil,  1962).  The  inhibition  is  probably  early  in 
the  sequence,  since  various  acyl-CoA's  protect,  but  it  is  not  on  the  NADPH: 
acetoacetyl-CoA  oxidoreductase.  The  incorporation  of  acetate- 1-C^^  into 
lipid  by  chloroplast  suspensions  is  also  strongly  depressed:  22%  by  0.001  mM 
p-MB,  50%  by  0.01  mM,  and  88%  by  0.1  mM  (Mudd  and  McManus,  1964). 
Fatty  acid  oxidation  is  potently  inhibited  by  the  mercurials,  and  one  likely 
site  is  the  initial  activation  by  ATP  (with  or  without  Co  A),  catalyzed  by 
fatty  acid  thiokinase,  since  this  is  completely  inhibited  by  0.1  mM  p-MB 
(Jencks  and  Lipmann,  1957).  The  incorporation  of  P,^^  into  mycobacterial 
phospholipids  is  not  depressed  so  readily,  1  mM  p-MB  inhibiting  only  24% 
(Tanaka,  1960),  although  the  synthesis  of  phospholipid  in  rat  liver  mito- 
chondria from  of-glycerophosphate  is  completely  blocked  at  this  concentra- 
tion (Wojtczak  et  al.,  1963). 

The  direct  actions  of  the  mercurials  on  lipid  biosynthesis  combined  with 
other  actions  which  would  secondarily  inhibit  these  pathways,  e.g.,  the  re- 
actions with  coenzyme  A  or  the  depletion  of  available  ATP,  must  lead  to 
serious  interference  in  the  formation  of  fatty  acids  and  sterols  in  proliferat- 
ing microorganisms  and  contribute  to  the  suppression  of  growth,  and  it  is 
interesting  to  speculate  whether  they  play  a  role  in  chronic  mercurial  poi- 
soning in  animals. 

Protein  Synthesis 

In  the  preparations  which  have  been  examined  it  appears  that  protein 
synthesis  is  not  particularly  sensitive  to  the  mercurials.  The  incorporation 
of  leucine-C^*  into  chloroplast  protein  is  inhibited  only  30%  by  5  mM  mer- 
salyl  (Stephenson  et  al.,  1956)  and  into  reticulocyte  protein  only  9%  by 
0.1  mM  p-MB,  although  1  mM  inhibits  almost  completely  (Borsook  et  al., 
1957),  while  the  incorporation  of  phenylalanine-C^*  into  rat  liver  soluble 
proteins  is  inhibited  85%  by  1  mM  p-MB  (Haining  et  al.,  1960),  of  amines 
into  guinea  pig  liver  soluble  proteins  100%  by  1  mM  p-MB  (Clarke  et  al., 
1959),  and  of  amino  acids  into  the  acid-soluble  proteins  of  frog  egg  super- 
natant fractions  100%  by  0.77  mM  p-MB  (Burr  and  Finamore,  1963). 
Although  these  results  do  not  conclusively  indicate  the  exact  sensitivity  of 
protein  synthesis  to  the  mercurials,  one  is  somewhat  surprised  to  find  that 
such  high  concentrations  apparently  must  be  used  to  inhibit  effectively. 
The  only  instance  of  potent  inhibition  of  which  I  am  aware  is  that  found  in 


888  7.  MERCURIALS 

Pseudomonas  aeruginosa  by  DeTurk  and  Bernheim  (1960),  the  induction 
of  enzymes  for  the  oxidation  of  putrescine,  benzoate,  and  }^-aminobutyrate 
being  reduced  50%  by  0.0028  mM  p-MB.  The  enzymes  themselves  are  not 
inhibited  at  this  concentration.  Partial  protection  by  Fe++  when  it  is  added 
with  the  mercurial  or  shortly  after  was  observed.  It  is  now  known  that 
enzyme  induction  is  not  a  valid  system  for  estimating  the  effects  of  in- 
hibitors on  protein  synthesis  in  general,  because  there  are  many  other  fac- 
tors involved.  In  the  inhibition  cited,  it  was  in  fact  postulated  that  some 
transport  process  in  the  membrane  requires  Fe++  and  that  this  is  the  site 
of  attack  by  the  mercurial. 

Porphyrin  Synthesis 

The  formation  of  porphyrins  from  glycine  and  a-ketoglutarate  by  Rhodo- 
pseudomonas  spheroides  is  completely  blocked  by  0.04  mM  p-MB,  and  from 
aminolevulinate  by  0.1  mM  (possibly  by  lower  concentrations  since  they 
were  not  tested)  (Lascelles,  1956).  The  formation  of  aminolevulinate  from 
glycine,  phosphoenolpyruvate,  and  succinyl-CoA  is  completely  prevented 
by  0.44  mM  p-MB  (Gibson  et  al.,  1958).  It  would  thus  appear  that  steps 
both  pre-  and  post-aminolevulinate  are  vulnerable.  The  report  of  Granick 
(1958)  that  1  mM  p-MB  does  not  interfere  with  protoporphyrin  synthesis 
from  glycine  and  or-ketoglutarate  in  chicken  erythrocytes  is  surprising,  but 
may  be  attributed  to  the  high  density  of  the  cell  suspension  (around  45% 
by  volume)  and  the  consequent  binding  of  the  mercurial  to  nonenzyme 
proteins.  The  condensation  of  porphobilinogen  to  uroporphyrinogen  is  al- 
most totally  blocked  by  0.02  mM  Hg++  and  0.1  mM  p-MB  (Lockwood 
and  Benson,  1960),  and  the  subsequent  conversion  of  uroporphyrinogen  to 
coproporphyrinogen  is  again  essentially  blocked  by  0.012  mM  Hg++  and 
0.7  mM  p-MB  (Mauzerall  and  Granick,  1958),  if  the  results  on  the  isolated 
enzymes  catalyzing  these  reactions  can  be  applied  to  cellular  preparations. 
The  incorporation  of  Fe++  into  protoporphyrin  to  form  heme  is  not  so  sen- 
sitive, in  chicken  erythrocyte  hemolyzate  being  inhibited  64%  and  58% 
by  1  mM  Hg++  and  p-MB,  respectively  (Kagawa  et  al.,  1959).  The  purified 
chelating  enzyme  from  rat  liver  is  inhibited  75%  by  0.1  mM  Hg++  (Labbe 
and  Hubbard,  1961),  the  greater  effect  probably  being  due  to  the  relative 
purity  of  the  preparation.  Again  one  can  speculate  that  a  depression  of 
porphyrin  synthesis  may  be  of  some  significance  in  growth  studies  or 
chronic  poisoning. 

Biolumlnescence 

One  of  the  very  few  thorough,  quantitative,  and  interesting  investiga- 
tions on  mercurial  inhibition  was  made  by  Houck  (1942),  who  studied  the 
effects  of  Hg++  on  the  luminescence  of  Achromobacter  fischeri.  The  standard 
conditions  were  as  follows:  pH  7.3,  temperature  25°,  25  mM  glucose  as 


VARIOUS    METABOLIC   PATHWAYS  889 

substrate,  and  a  suspension  density  of  4  X  10^  cells/ml.  Both  respiration 
and  luminescence  of  these  cells  are  inhibited  potently  by  Hg++,  the  latter 
being  somewhat  more  sensitive  (Fig.  7-35).  The  rate  of  inhibition  is  much 
more  rapid  than  with  most  cellular  activities,  half  maximal  inhibition  being 
reached  in  about  30-40  sec  (Fig.  7-36).  It  is  not  immediately  evident  why 
the  inhibition  is  more  potent  in  the  rate  experiments  than  in  the  studies 
on  the  effect  of  concentration.  The  effects  of  cell  density  on  the  inhibitions 
are  very  striking  (Fig.  7-37).  The  initial  suspension  here  contained  6  X  10^ 
cells/ml  and  this  was  diluted  as  indicated  in  the  graph.  At  0.001  niM  no 

lOOi 


LUMINESCENCE 


0  015 


0.02 


0  025 


0.03 


Fig.  7-35.  Effects  of  Hg++  on  the  respiration  and  luminescence  of 
Achromobacter.  (From  Houck,   1942.) 


inhibition  is  observed  until  sufficient  dilution  is  made  and  at  low  cell  den- 
sities the  inhibition  is  complete.  These  curves  illustrate  very  well  what  es- 
sentially must  occur  in  all  cell  or  tissue  preparations,  whatever  activity  is 
measured.  The  meaninglessness  of  statements  that  such  and  such  a  con- 
centration of  mercurial  produces  a  certain  degree  of  inhibition  of  some  cel- 
lular process  is  all  too  clear;  in  this  case  with  0.001  mM  Hg++,  one  might 
observe  any  inhibition  from  0  to  100%  depending  on  the  cell  density  chosen. 
Inhibition  was  determined  with  0.001  mM  Hg++  at  three  values  of  the  pH, 
and  was  greatest  at  5.3  and  8.4,  and  least  at  7.3  (one  can  estimate  the  mean 
per  cent  inhibitions  at  1  min  to  be  91%,  60%,  and  86%  at  pH  5.3,  7.3, 
and  8.4,  respectively).  The  light  intensity  is  much  greater  at  pH  7.3  and 
this  may  possibly  be  related  to  the  rate  of  glucose  uptake.  The  effects  of 
temperature  have  already  been  illustrated  (Fig.  1-15-9)  and  discussed  (page 


890 


7.  MERCURIALS 


250 

SEC 


Fig.  7-36.  Effects  of  Hg++  at  different  concentrations  on  lumines- 
cence of  Achromohacter .  (From  Houck,  1942.) 


LUMINESCENCE 
RESPIRATION 


1/4  1/9  1/16  1/32  1/64 


DILUTION      OF     CELL     SUSPENSION 


Fig.  7-37.   Effects  of  dilution  of  the  Achromobacter  sus- 
pension on  the  inhibitions  of  respiration   and   lumines- 
cence by  Hg++.  (From  Houck,  1942.) 


VARIOUS   METABOLIC   PATHWAYS  891 

1-786).  The  increase  of  the  inhibition  with  rise  of  temperature  was  inter- 
preted by  Houck  in  terms  of  an  equilibrium  between  active  and  denatured 
forms  of  the  attacked  enzyme,  especially  luciferase.  Actually  from  this 
work  one  cannot  locate  the  site  of  the  inhibition,  or  even  be  certain  it  is 
on  the  bioluminescent  reactions  themselves,  since  interference  with  glucose 
uptake  or  oxidation,  or  the  supply  of  ATP,  could  be  responsible.  However, 
it  has  been  found  that  Achromohacter  luciferase  is  markedly  inhibited  in  the 
range  of  Hg++  concentrations  found  to  inhibit  luminescence  (Table  7-13), 
so  it  may  well  be  that  luciferase  is  the  major  site  of  action.  This  is  somewhat 
substantiated  by  the  fact  that  luminescence  in  extracts  of  Renilla  reniforniis, 
the  sea  pansy,  is  strongly  inhibited  by  p-MB  (Cormier,  1960). 

Photosynthesis   and    Photophosphorylation 

The  marked  inhibition  of  certain  phases  of  photosynthesis  by  iodoacetate 
and  iodoacetamide  (III-1-156)  indicates  the  necessity  of  SH  groups,  so  that 
one  would  expect  the  mercurials  to  be  effective  inhibitors,  and  this  is  borne 
out.  The  photoreduction  of  various  dyes  in  isolated  chloroplasts  or  grana 
(Hill  reaction)  is  very  sensitive.  In  spinach  chloroplasts  it  is  inhibited  90% 
by  0.005  mi/  Hg++  (Macdowall,  1949).  The  dye  reduction  may  be  me- 
diated through  NADPH,  which  is  the  initial  acceptor.  The  photosjmthetic 
NADP  reductase  from  spinach  is  inhibited  50%  by  0.012  mM  and  90%by 
0.016  mM  p-MB  (San  Pietro  and  Lang,  1958)  and  the  photoreduction  of 
NADP  in  chloroplasts  is  similarly  inhibited,  although  slightly  less  potently 
(J.  S.  C.  Wessels,  1959).  The  photoreduction  of  cytochrome  c  and  NADP 
by  the  chloroplast  enzyme  is  50%  reduced  by  0.004  mM  p-MPS  and  the 
enzjTne  is  bleached  by  the  mercurial  (Keister  and  San  Pietro,  1963).  In 
Chromatium,  illumination  causes  a  blue  fluorescence  presumably  due  to 
bound  NADH,  indicating  that  here  there  is  a  photoreduction  of  NAD. 
This  fluorescence  change  is  completely  abolished  by  0.02  mM  PM  (Olson 
et  al,  1959).  Finally,  a  NADPH  diaphorase  from  chloroplasts,  possibly  in- 
volved in  the  reduction  of  the  Hill  dyes  by  NADPH,  is  inhibited  53%  by 
0.023  mM  Hg++  and  50%  by  0.13  mM  p-MB  (Avron  and  Jagendorf,  1956). 
The  initial  photoreductive  changes  upon  illumination  are  thus  quite  po- 
tently inhibited  by  the  mercurials,  and  this  must  certainly  be  one  site  of 
action  on  over-aU  photosynthesis.  Other  evidence  for  a  primary  interference 
with  the  photolysis  of  water  was  obtained  by  Damaschke  and  Liibke  (1958), 
who  showed  that  Chlorella  under  anaerobic  conditions  produces  a  sudden 
burst  of  Hg  upon  illumination  and  that  this  is  completely  inhibited  by  0.2 
mM  p-MB  (lower  concentrations  not  tested),  and  by  Whittingham  (1956), 
who  found  that  althoagh  0.12  mM  p-MB  does  not  inhibit  the  initial  evo- 
lution of  Oo  by  illuminated  Chlorella,  the  steady-state  formation  of  Og  is 
strongly  depressed.  It  may  be  mentioned  that  even  high  concentrations  of 
Hg++  do  not  react  with  chlorophyll  (Macdowall,  1949). 


892  7.  MERCURIALS 

Photophosphorylation  to  form  ATP  is  not  necessarily  coupled  with  NADP 
reduction  (J.  S.  C.  Wessels,  1959),  but  nevertheless  one  might  predict  that 
it  would  be  reduced  by  mercurials.  It  has  been  found  that  the  incorporation 
of  Pj  into  ATP  in  illuminated  chlorophasts  is  inhibited  around  50%  by 
p-MB  at  0.05-0.1  mM  (Arnon  et  al,  1956;  J.  S.  C.  Wessels,  1958,  1959; 
Jagendorf  and  Avron,  1959),  and  similar  effects  were  reported  for  Rhodo- 
spirillum  rubrum  (Smith  and  Baltscheffsky,  1959).  Photophosphorylation 
is  not  inhibited  as  potently  as  photoreduction. 

The  photochemical  fixation  of  C^^Og  by  chloroplasts  is  inhibited  14%  and 
88%  by  0.01  and  0.05  mM  p-MB,  respectively  (Gibbs  and  Calo,  1959  b), 
but  a  reconstructed  system  (extract  +  chloroplast  fragments)  is  more  sen- 
sitive, 61%  and  94%  inhibition  being  exerted  by  these  concentrations  of 
p-MB  (Gibbs  and  Calo,  1960  b).  It  is  not  known  if  this  implies  some  barrier 
to  penetration  in  the  intact  chloroplast.  Both  the  initial  and  steady-state 
rates  of  fixation  of  CO2  in  illuminated  dahlia  leaves  are  only  slightly  re- 
duced (15-25%)  by  0.5  mM  p-MB,  even  though  plenty  of  time  is  provided 
for  penetration  (Massini,  1957),  and  in  Scenedesmus  obliquus  photosynthesis 
is  inhibited  only  50%  by  1  mM  p-MB  after  260  min  exposure  (Horwitz, 
1957).  The  failure  to  inhibit  more  potently  in  these  cases  can  at  present 
be  explained  only  on  the  basis  of  inadequate  penetration  into  the  cells  or  a 
certain  structural  integrity  of  the  photosynthetic  apparatus  which  makes 
it  difficult  for  a  mercurial  to  exert  such  inhibition  as  is  observed  with  isolat- 
ed chloroplasts.  No  detailed  study  of  the  effects  of  mercurials  on  the  rapidly 
labeled  C  compounds  has  been  made,  but  Miyachi  (1960)  has  found  that 
p-MB  decreases  the  level  of  what  he  calls  the  primary  photogenic  agent 
(measured  by  3-sec  C^^Og  fixation)  in  ChloreUa,  although  it  does  not  inter- 
fere with  the  participation  of  this  substance  in  the  subsequent  photosyn- 
thetic pathway.  Nonphotosynthetic  C^^Og  fixation  is  usually  inhibited 
strongly  by  mercurials,  e.g.,  the  autotrophic  fixation  by  Hydrogenomonas 
facilis  (McFadden  and  Atkinson,  1957)  or  the  fixation  associated  with  sul- 
fide oxidation  in  Thiobacillus  thiooxidans  (Iwatsuka  et  al.,  1962),  both  being 
inhibited  around  50%  by  0.01  mM  p-MB  —  which  is  not  surprising  con- 
sidering the  sensitivity  of  the  various  enzymes  usually  involved  in  COg 
fixation.  This  dark  fixation  is  possibly  related  in  some  manner  to  photo- 
synthesis, and  it  has  occasionally  been  pointed  out  that  the  same  inhibitors 
are  effective  in  both. 

THE  CELL  MEMBRANE  AS  A  SITE  FOR  MERCURIAL  ACTION 

In  the  following  section  we  shall  discuss  the  effects  of  the  mercurials  on 
permeability  and  membrane  transport  systems,  as  a  background  for  under- 
standing the  responses  of  tissues  to  these  inhibitors,  but  it  may  serve  to 
clarify  the  problem  if  we  take  up  the  theory  of  the  role  of  the  cell  membrane 


THE    CELL    MEMBRANE    AS    A    SITE    FOR   MERCURIAL   ACTION  893 

in  heavy  metal  ion  inhibitions  as  an  introduction.  Although  many  workers 
have  considered  the  effects  of  heavy  metal  ions  on  membranes,  the  concepts 
presented  here  will  be  mainly  those  of  Rothstein  and  his  group  at  Rochester, 
since  they  have  been  actively  engaged  for  over  10  years  in  studying  this 
problem.  Although  much  of  the  evidence  is  based  on  work  with  copper, 
molybdate,  and  uranyl  ions,  and  the  theory  is  meant  to  apply  to  heavy 
metal  ion  action  in  general,  Hg++  has  been  used  frequently  and  it  is  im- 
possible to  discuss  the  effects  of  the  mercurials  on  cells  and  tissue  without 
considering  this  aspect  of  their  actions.  The  basic  concepts  of  Rothstein 
(1959)  may  be  summarized  as  follows.  (1)  The  cell  membrane  is  exposed 
directly  to  the  heavy  metal  ions  in  the  medium  and  is  that  part  of  the  cell 
which  reacts  initially  when  the  heavy  metal  ions  are  added.  Ligand  groups 
at  the  surface  or  within  the  membrane  will  combine  with  the  heavy  metal 
ions  as  they  diffuse,  and  hence  the  membrane  will  experience  the  first  ef- 
fects, and  certain  changes  in  cellular  metabolism  or  function  may  at  this 
time  relate  to  a  selective  membrane  action.  (2)  The  cell  membrane  usually 
presents  a  barrier  to  the  penetration  of  the  heavy  metal  ion  into  the  cell 
and  thus  protects  the  cytoplasmic  enzymes.  (3)  Nonenzymic  or  nonfuc- 
tional  ligand  groups  in  the  membrane  or  within  the  cell  combine  with  the 
heavy  metal  ions  and  thereby  protect  the  active  sites  by  reducing  the 
amount  of  heavy  metal  ion  which  is  free  to  react.  (4)  As  a  result  of  the 
second  and  third  postulates,  enzymes  within  cells  are  less  readily  attacked 
by  metal  ions  than  when  they  are  isolated  from  the  cells.  (5)  As  a  result 
of  the  first  statement  and  the  fourth  deduction,  it  would  be  likely  that  the 
major  site  of  heavy  metal  ion  action  on  cells  and  tissues  is  often  the  cell 
membrane,  rather  than  the  enzyme  and  metabolic  systems  within  the  cell. 
(6)  The  most  important  active  sites  in  the  membrane  are  enzymes  or  other 
components  involved  in  the  transport  of  substances  across  the  membrane. 
Much  of  the  toxicity  would  therefore  be  due  to  interference  with  the  move- 
ments of  substrates  or  ions  into  or  out  of  the  cell. 

Glucose  Uptake  and  Respiration  of  Diaphragm  Muscle 

The  uptake  of  glucose  by  rat  diaphragm  is  almost  completely  abolished 
within  20-30  min  by  0.2  milf  Hg++,  but  the  respiration  is  not  affected  be- 
fore 30  min  and  is  inhibited  only  30%  maximally  after  2  hr  (Fig.  1-12-31) 
(Demis  and  Rothstein,  1955).  It  requires  2  mM  Hg++  to  inhibit  the  res- 
piration 90%  and  this  occurs  after  1.5  hr.  Thus  glucose  uptake  is  depressed 
much  more  rapidly  and  is  more  sensitive  than  respiration  by  at  least  a 
factor  of  10.  These  results  might  imply  that  Hg++  acts  initially  on  the  mem- 
brane to  block  glucose  transport,  and  later  on  intracellular  respiratory  sys- 
tems; depression  of  glucose  uptake  does  not  in  itself  inhibit  respiration  since 
the  latter  is  dependent  on  endogenous  substrate.  Of  some  confirmatory  evi- 
dence is  the  fact  that  cysteine  will  reverse  the  inhibition  of  glucose  uptake 


894 


7.  MERCURIALS 


but  will  not  restore  the  respiration  once  it  is  inhibited;  i.e.,  the  surface- 
bound  Hg++  is  available  to  the  cysteine,  but  penetration  of  the  amino  acid 
into  the  cells  is  inadequate  to  remove  the  Hg++  responsible  for  reducing 
the  respiration. 

Uptake   of  Hg++   by    Diaphragm    Muscle 

Logarithmic  plots  of  Hg++  uptake  with  different  initial  concentrations 
in  the  medium  are  shown  in  Fig.  7-38.  There  appears  to  be  two  components, 
a  fast  phase  with  a  half-time  of  12  min  and  a  slow  phase  with  a  half-time 
of  around  60  min.  The  uptake  essentially  ceases  after  20-30  min  at  low 


Fig.  7-38.  The  uptake  of  Hg++  by  rat  diaphragm, 

at  pH  7.4  and  38°,  with  time,  as  determined  by  the 

Hg++  remaining  in  the  medium.   (From   Demis  and 

Rothstein,  1955.) 


initial  concentrations.  It  was  assumed  that  the  fast  phase  corresponds  to 
the  diffusion  of  Hg++  into  the  extracellular  space  and  binding  to  the  plasma 
membranes,  the  slow  phase  to  the  penetration  into  the  cells.  The  time  re- 
lations point  to  a  correlation  between  the  membrane  binding  and  the  inhi- 
bition of  glucose  uptake,  and  between  penetration  and  respiratory  kihi- 
bition. 


THE    CELL    MEMBRANE   AS    A   SITE    FOR    MERCURIAL    ACTION  895 

There  are  certain  aspects  of  these  results  which  are  puzzling  and  seem 
to  me  to  be  difficult  to  reconcile  with  the  simple  theory  presented.  Why 
does  the  Hg++  uptake  cease  after  the  fast  phase  for  the  lower  initial  con- 
centrations (0.2  mM  or  below)  (Fig.  7-38)?  Let  us  estimate  how  much 
Hg++  is  taken  up  by  the  diaphragm  in  the  fast  phase  (see  accompanying 
tabulation).  If  this  represents  Hg++  bound  to  membranes  and  as  much  as 


Initial  Hg+  + 

Quantity  of  Hg++ 

concentration 

%  Uptake 

taken  up 

(mif) 

(/^moles/g  tissue) 

0.9 

63 

9.4 

0.6 

77 

7.7 

0.2 

85 

2.8 

0.05 

84 

0.70 

0.025 

78 

0.33 

9.4  //moles/g  of  tissue  can  be  taken  up,  why  does  uptake  stop  when  so 
little  is  bound  at  the  lower  concentrations?  One  would  expect  all  the  Hg++ 
to  disappear  from  the  medium  when  the  initial  amount  of  Hg++  is  less 
than  that  required  to  saturate  the  membranes.  The  maximal  Hg++  bound 
finally  at  the  highest  concentration  was  stated  to  be  about  15  ^/moles/g 
of  tissue.  Inasmuch  as  the  plasma  membranes  cannot  contribute  more 
than  1%  of  the  tissue  mass,  how  is  it  that  they  can  bind  over  half  this 
amount  ? 

The  amount  of  Hg++  diffusing  into  the  extracellular  space  and  existing 
there  unbound  must  be  negligible,  since  the  diaphragms  weighed  0.6  g  and 
the  total  medium  volume  was  10  ml,  so  that  the  extracellular  space  would 
be  approximately  1  %  of  the  total  volume.  We  have  mentioned  that  Casca- 
rano  and  Zweifach  (1962)  found  diaphragm  exposed  to  Hg++  to  show  evi- 
dence of  dehydrogenase  inhibition  only  in  the  outer  few  layers  of  cells 
(page  879).  Thus  Hg++  does  not  appear  to  penetrate  readily  throughout 
the  tissue.  One  must  ask  if  the  fast  phase  of  uptake  may  be  correlated  with 
binding  only  to  the  membranes  of  the  outermost  layer  of  cells.  A  rough 
estimate  of  the  protein  contained  in  the  outermost  membranes  on  both 
sides  of  diaphragms,  assuming  a  generous  membrane  thickness  of  200  A, 
gives  1.5  X  10"*  //mole/g  of  tissue.  If  all  the  Hg++  taken  up  w-ere  bound  by 
these  membranes,  there  would  be  62,000  Hg++  ions  bound  per  protein 
molecule  (of  assumed  molecular  weight  100,000),  and  since  this  value  is 
impossibly  large,  one  must  conclude  that  most  of  the  Hg++  must  be  bound 
deeper  in  the  tissue.  Demis  and  Rothstein  (1955)  assumed  that  the  Hg++ 
is  not  bound  entirely  to  the  outermost  ceUs,  but  to  the  plasma  membranes 
throughout  the  diaphragm.  It  is  difficult  to  estimate  the  amount  of  protein 


896  7.   MEKCURIALS 

in  the  total  membrane,  but  it  seems  very  unlikely  that  it  could  accommo- 
date all  the  Hg++  taken  up  at  the  higher  concentrations,  especially  if  pene- 
tration deep  into  the  tissue  does  not  occur. 

The  uptake  data  by  themselves  could  be  explained  in  a  variety  of  ways. 
Binding  to  proteins  often  shows  different  phases  due  to  the  different  reac- 
tivities of  the  various  types  of  SH  group,  and  in  cellular  systems  one  must 
perhaps  also  consider  ligands  other  than  SH  groups.  But  how  can  one  inter- 
pret the  results  on  glucose  utilization  and  respiration,  especially  as  they 
seem  to  be  correlated  in  time  with  the  Hg++  uptake  phases?  Particularly, 
why  is  there  such  a  long  lag  period  before  respiration  is  depressed?  It  may 
be  noted  that  a  lag  period  is  not  always  observed  in  other  tissues  or  cell 
suspensions.  One  possibility  which  cannot  be  ignored  is  that  the  Hg++  en- 
ters the  cells  early  but  is  initiaUy  and  preferentially  bound  to  SH  groups 
not  involved  with  respiration.  In  muscle  cells  this  might  be  more  evident 
than  in  other  tissues  because  of  the  large  amounts  of  actin  and  myosin, 
each  of  which  possesses  numerous  SH  groups;  only  when  these  groups  be- 
come saturated  with  Hg++  would  effects  on  the  oxidation  enzymes  be  ob- 
served. It  is  unfortunate  that  the  effects  on  muscle  contraction  were  not 
determined,  since  if  this  explanation  is  valid,  contractile  activity  should 
be  reduced  during  the  fast  phase  of  uptake.  In  this  case  the  fast  phase 
would  refer  to  the  binding  to  membrane  and  actomyosin  (and  any  other 
reactive  ligands),  the  membrane  contributing  only  slightly.  The  kinetics 
of  the  effects  of  mercurials  on  diaphragm  contraction  have  apparently  not 
been  studied,  but  one  notes  that  the  diaphragms  exposed  to  1  mM  p-MB 
for  30  min  by  Kono  and  Colowick  (1961)  were  stated  to  be  in  contracture. 
On  the  other  hand,  the  results  obtained  with  rat  atria  exposed  to  0.05  mM 
p-MB  indicate  that  no  effect  on  the  contractile  amplitude  occurs  during  the 
initial  22  min,  although  effects  on  the  membrane  are  evident  (decrease  in 
magnitude  and  duration  of  the  action  potential),  and  that  depression  of  the 
contraction  proceeds  subsequently  (Webb  and  Hollander,  1959).  These  re- 
sults on  atria  thus  would  fit  into  the  theory  of  Rothstein.  However,  it  must 
be  remembered  that  in  obtaining  transmembrane  potentials  one  examines 
only  the  outermost  cells,  and  that  contractile  amplitude  involves  the  entire 
tissue;  for  this  reason  one  would  expect  a  delay  in  contractile  response.  A 
decision  cannot  be  made  until  direct  experiments  on  respiratory  and  con- 
tractile response  are  made  in  diaphragms.  It  must  be  emphasized  that  any 
modification  of  the  concepts  of  Rothstein  suggested  here  are  not  necessarily 
applicable  to  other  heavy  metal  ions  or  other  cells  (especially  yeast),  but 
relate  to  mercurials  only. 

Another  factor  which  must  be  considered  in  tissue  uptake  studies  with 
the  mercurials  is  the  possibility  of  damage  to  the  external  layers,  mani- 
fested by  increased  permeability  and  exposure  of  reactive  SH  groups,  espe- 
cially with  the  higher  concentrations  often  used.  The  high  degree  of  bind- 


THE    CELL   MEMBRANE    AS   A   SITE    FOR   MERCURIAL   ACTION  897 

ing  observed  by  Demis  and  Rothstein  (1955)  with  0.6-0.9  mM  Hg++  might 
be  due  in  part  to  this,  the  outer  layers  of  cells  picking  up  the  Hg++  not 
only  in  the  membranes  but  within  the  cells.  One  might  try  to  estimate 
roughly  the  amount  of  Hg++  required  for  saturation  of  membrane  sites  by 
determining  the  initial  concentration  so  that  all  the  Hg++  is  removed  from 
the  medium.  If  one  plots  as  accurately  as  possible  the  amount  of  Hg++ 
remaining  in  the  medium  after  the  fast  phase  against  the  initial  concentra- 
tion, one  finds  that  the  nearly  linear  curve  passes  almost  exactly  through 
the  origin.  All  this  shows  is  that  it  must  require  very  little  Hg++  to  saturate 
the  ligands  involved  in  the  fast  phase  uptake. 

It  is  of  some  interest  to  attempt  an  estimate  of  the  concentration  of 
membrane  SH  groups  in  certain  cellular  suspensions  in  order  to  obtain  some 
idea  of  the  order  of  magnitude.  In  the  experiments  of  Houck  (1942)  the 
suspensions  contained  4  x  10^  cells/ml  oi  Achromobacter  fischeri  under  stand- 
ard conditions.  Achromohacter  is  a  rod  with  dimensions  0.9  X  1.8  ;/  and 
thus  the  surface  area  of  a  single  cell  is  around  3.8  X  10~^  cm^.  If  one  as- 
sumes that  the  membrane  is  200  A  thick  (which  is  probablj^  too  high),  that 
the  membrane  is  50%  water  (since  it  is  perhaps  more  condensed  than  the 
cytoplasm),  that  the  membrane  solids  include  65%  protein  (values  of  this 
magnitude  have  been  obtained  for  the  membranes  of  other  bacteria),  that 
protein  specific  gravity  is  1.4.,  that  the  mean  molecular  weight  of  the  mem- 
brane proteins  is  100,000,  and  that  there  are  approximately  10  reactive  SH 
groups  on  a  protein  of  this  molecular  weight,  one  can  calculate  that  the 
concentration  of  membrane  SH  groups  is  close  to  10"^  rcvM.  The  lowest 
Hg++  concentration  to  produce  reduction  of  luminescence  was  10~^  m.M, 
so  that  even  at  this  lowest  concentration  the  Hg++  was  around  100  fold  in 
excess  of  the  membrane  SH  groups.  Of  course,  ligands  other  than  SH  groups 
may  occur  in  the  membrane.  This  suspension  of  Achromohacter  is  fairly 
dilute  relative  to  most  suspensions  used,  since  calculation  of  total  cell  vol- 
ume indicates  that  the  cells  occupy  0.046%  of  the  total  volume.  In  more 
concentrated  suspensions,  such  as  are  often  used,  the  situation  can  be  quite 
different.  A  10%  suspension  of  human  erythrocytes  (1.16  X  10^  cells/ml, 
cell  surface  area  =  1.4  X  10~®  cm^,  membrane  thickness  =  82  A,  and  10  SH 
groups/protein  molecule  of  molecular  weight  100,000)  would  be  0.059  mM 
with  respect  to  membrane  SH  groups,  so  that  an  appreciable  amount  of 
Hg++  might  be  bound  by  the  membranes  in  this  case.  In  any  study  relat- 
ing to  a  theory  of  membrane  binding  of  heavy  metal  ions,  it  would  be  well 
to  make  some  reasonable  estimates  of  the  concentration  of  membrane  lig- 
and  groups.  Although  such  calculations  cannot  be  very  accurate,  the  experi- 
mental results  may  be  of  an  entirely  different  order  of  magnitude,  which 
should  impel  the  investigator  to  question  the  validity  of  the  theory. 


898  '  7.  MERCURIALS 

Comparison  of  Effects  of  Hg++  on  Intact  Diaphragm  and  Homogenates 

The  endogenous  respiration  of  diaphragm  homogenates  fortified  with  ATP 
was  claimed  by  Demis  and  Rothstein  (1955)  to  be  inhibited  faster  and  less 
potently  than  the  respiration  of  intact  diaphragm  by  Hg++.  Actually,  from 
the  data  given,  it  is  not  evident  that  the  rate  of  inhibition  in  homogenates 
is  much  faster;  at  10  min  after  adding  Hg++,  for  example,  there  is  no  sig- 
nificant difference  in  the  rates  judged  from  the  points  presented,  although 
from  then  on  the  rate  in  intact  diaphragm  falls  off,  so  that  the  inhibitions 
are  not  equivalent  again  until  50  min.  It  was  stated  that  it  requires  10 
times  the  concentration  of  Hg++  to  inhibit  homogenate  respiration  compar- 
ed to  intact  tissue  [in  a  later  review  Rothstein  (1959)  stated  200  times], 
but  no  data  on  this  point  are  given  (the  only  experiment  reported  is  with 
the  extremely  high  concentration  of  9  m.M),  and  in  any  case  it  depends 
on  what  time  is  chosen  to  compare  the  inhibitions  (e.g.,  up  to  50  min, 
homogenate  respiration  is  inhibited  more  strongly  by  9  mM  Hg++).  It  is, 
furthermore,  very  difficult  to  interpret  differences  in  inhibitions  of  intact 
cells  and  extracts,  since  the  substrates  utilized,  the  pathways  taken,  and 
the  states  of  the  enzymes  are  probably  very  different.  Mercurial  inhibition 
has  usually  been  found  to  be  more  potent  in  cell  extracts  than  intact  cells, 
e.g.,  Nakayama  (1959)  reported  that  while  0.077  mM  p-MB  inhibits  ethanol 
oxidation  9%  in  Acetobacter,  it  requires  only  0.0077  mM  to  inhibit  14%  in 
extracts.  How  much  role  the  membrane  plays  in  any  of  these  observations 
is  impossible  to  determine. 

Binding  of  Hg++  to  Yeast  Cells  and    Loss  of  K+ 

The  efflux  of  K+  from  yeast  is  accelerated  by  Hg++  as  it  is  from  most 
cells.  Although  the  effects  of  the  mercurials  on  permeability  and  active 
transport  will  be  taken  up  later,  the  work  done  by  Rothstein  and  his  co- 
workers will  be  treated  here  since  it  has  bearing  on  the  concept  of  differen- 
tial membrane  binding.  Rothstein  and  Bruce  (1958)  studied  the  efflux  of 
K+  into  a  K+-free  medium  flowing  through  a  yeast  cell  column;  since  the 
pH  of  the  medium  was  3.5,  and  lowering  the  pH  enhances  the  efflux  rate, 
it  was  assumed  that  the  process  is  mainly  a  K+-H+  exchange.  The  loss  of 
K+  from  the  cells  is  very  sensitive  to  Hg++,  0.001  mM  producing  a  slight 
effect  after  a  long  lag  period,  and  0.003  mM  producing  at  least  a  tripling 
of  the  rate;  at  the  highest  concentration  used,  0.1  mM,  80%  of  the  cell  K+ 
is  lost  in  1  hr.*  Passow  and  Rothstein  (1960)  used  a  different  technique  in 
that  the  rate  of  K+  loss  into  a  medium  (distilled  water  adjusted  to  pH  3 
with  HCl)  from  a  suspension  of  yeast  cells  was  measured.  The  minimal  ef- 
fective concentration  of  Hg++  to  accelerate  the  efflux  was  found  to  be  0.2 

*  Dr.  Rothstein  informed  me  that  Fig.  6  of  the  paper  by  Rothstein  and  Bruce 
(1958)  presents  the  cumulative  K+  loss  rather  than  the  rate  of  K+  loss  as  stated. 


THE    CELL   MEMBRANE    AS   A    SITE    FOR    MERCURIAL    ACTION  899 

ToM,  and  1.6  mM  produces  essentially  a  complete  loss  of  the  cell  K+  in 
2  hr.*  It  is  possible  from  the  results  with  the  yeast  columns  that  Hg++  at 
low  concentrations  has  a  specific  effect  on  K+  permeability  without  depres- 
sing active  transport,  and  this  is  borne  out  in  the  work  with  erythrocytes. 
It  was  stated  that  the  curve  obtained  by  plotting  log(IIg++)  against  max- 
imal K+  loss  is  sigmoid,  which  fits  a  "normal  distribution"  (presumably  of 
susceptibility  of  different  yeast  cells  to  Hg++),  and  that  the  loss  of  K+  is 
probably  an  all-or-none  phenomenon,  this  being  confirmed  by  determina- 
tions of  staining  by  certain  dyes  in  Hg++-treated  cells.  Although  yeast  cells 
undoubtedly  show  a  variation  in  the  sensitivity  to  Hg++,  it  is  doubtful  if 
the  evidence  is  sufficient  to  categorize  the  K+  loss  as  all-or-none,  especially 
since  sigmoid  curves  of  this  type  (they  are  not  given  so  one  cannot  directly 
evaluate  them)  are  also  compatible  with  graded  effects  and,  in  fact,  are  the 
commonest  relations  observed  in  the  actions  of  most  inhibitors  on  cell  me- 
tabolism or  function.  There  is  an  increase  in  general  membrane  permeability 
produced  by  Hg++,  as  proved  by  the  loss  of  a  variety  of  substances  from  the 
cells  and  a  greater  penetration  of  dyes,  and  this  could  be  a  graded  phenom- 
enon occurring  simultaneously  with  the  alterations  in  K+  efflux,  without 
the  need  for  assuming  cytolysis  as  the  necessary  concomitant  of  K+  loss. 
Hg++  is  bound  relatively  rapidly  to  yeast  cells,  the  half-time  being  2-4 
min  and  maximal  binding  occurring  in  15-20  min.  Passow  and  Rothstein 
(1960)  determined  the  uptake  of  both  Hg++  and  Cl~,  and  found  that  ini- 
tially only  Hg++  is  bound,  the  Cl~  entering  when  the  concentration  of  Hg++ 
is  sufficiently  high.  The  binding  at  low  concentrations  was  thus  claimed  to 
represent  "binding  of  Hg++  rather  than  HgClg."  Since  the  concentration  of 
the  Hg++  ion  is  actually  extremely  small,  it  seems  more  likely  that  HgClg 
or  other  chloride  complexes  react  with  the  yeast  cell  wall  and  membrane, 
releasing  the  CI  which  diffuses  into  the  medium.  When  the  concentration 
of  the  mercurial  becomes  great  enough  to  lead  to  a  significant  increase  in 
permeability,  Cl~  then  enters,  either  alone  or  with  Hg++.  The  general  con- 
clusion is  that  the  membrane  effect  of  Hg++  is  not  specific  for  K+  but  is  a 
more  or  less  nonspecific  breakdown  of  the  membrane,  caused  by  the  "mol- 

*  The  approximately  1000-fold  difference  in  sensitivity  observed  in  these  two  types 
of  experiment  deserves  some  comment  and  Dr.  Rothstein  has  kindly  provided  me  with 
the  reasons.  In  the  suspension  experiments  the  yeast  density  was  60  mg/ml  and  at 
0.4  mM  Hg++  the  maximum  binding  of  the  metal  would  be  about  7  millimoles/kg 
of  cells.  In  the  column  experiments  with  600  mg  of  cells  and  a  flow  rate  of  5  ml/min, 
the  maximum  binding  in  30  min  at  0.05  mM  Hg++  would  be  only  0.015  millimole/kg. 
Thus  the  yeast  in  the  column  would  be  much  more  readily  affected  since  less  of  the 
Hg++  is  removed.  Second,  the  suspension  experiments  measure  the  steady-state  flux 
and  the  net  loss  of  K+,  whereas  the  column  experiments  measure  the  rate  of  efflux 
into  K+-free  medium.  It  is  therefore  difficult  to  compare  the  results  by  the  two  tech- 
niques on  a  quantitative  basis. 


900  7.  MERCURIALS 

ecular  stress"  brought  about  by  the  formation  of  S — Hg — S  bridges  in  the 
membrane;  when  this  stress  reaches  a  critical  level,  the  membrane  disinte- 
grates and  cellular  components  are  released  (Rothstein,  1959).  Little  con- 
sideration is  given  to  the  possible  effects  of  Hg++  on  the  active  transport 
mechanisms  by  which  K+  is  accumulated  and  emphasis  is  placed  on  the 
structural  changes  occurring  in  the  membrane.  Most  of  the  studies  on  K+ 
loss  from  tissues  have  been  interpreted  in  terms  of  an  inhibition  of  active 
transport  (page  907),  and  it  seems  that  this  would  be  the  more  direct  and 
logical  explanation.  It  should  also  be  pointed  out  that,  as  in  all  studies  of 
the  effects  of  substances  on  transmembrane  fluxes,  it  is  very  difiicult  to 
distinguish  between  actions  on  the  membrane  and  within  the  cells,  and 
that  therefore  these  results  in  themselves  cannot  be  taken  as  evidence  for 
a  direct  or  specific  membrane  effect. 

Erythrocyte  Permeability  and  Hemolysis 

Organic  mercurials  increase  erythrocyte  fragility  and  promote  hemolysis, 
often  at  quite  low  concentrations,  but  the  effects  of  Hg++  are  more  complex, 
hemolysis  being  either  favored  or  inhibited  depending  on  the  conditions,  of 
which  the  concentration  of  Hg++  and  the  type  of  hemolysis  are  the  most 
important.  If  hemolysis  in  isotonic  glycerol  is  studied,  Hg++  can  markedly 
delay  the  hemolysis.  Human  erythrocytes  hemolyze  rapidly  in  isotonic  gly- 
cerol at  pH  7.2;  as  the  concentration  of  Hg++  is  increased,  inhibition  is  first 
observed  at  0.025  mM  and  very  strong  inhibition  at  0.05  mM  (Wilbrandt, 
1941).  This  was  interpreted  as  an  inhibition  of  glycerol  entry  into  the  cells 
by  Hg++.  On  the  other  hand,  if  hypotonic  hemolysis  of  human  erythrocytes 
is  examined  (i.e.,  hemolysis  in  Tyrode  solution  diluted  to  varying  degrees), 
Hg++  can  either  accelerate  or  slow  hemolysis  (Fig.  7-39)  (Jung,  1947).  In 
normal  or  weakly  diluted  medium,  Hg++  favors  hemolysis,  but  at  low  con- 
centration it  suppresses  hemolysis  in  markedly  hypotonic  media.  Jung  be- 
lieved that  the  resistance  to  osmotic  effects  is  mediated  through  a  denatur- 
ation  of  the  membrane.  Arbuthnott  (1962)  has  recently  confirmed  the  dual 
action  of  Hg++,  hemolysis  of  rabbit  erythrocytes  being  promoted  by  low 
concentrations  and  inhibited  by  concentrations  around  1  mM.  Organic  mer- 
curials {p-MB,  ethyl-Hg+,  and  thimerosal),  however,  are  only  lytic,  even 
at  high  concentrations.  Arbuthnott  related  this  to  the  number  of  charges 
on  the  mercurials,  although  it  is  more  likely  a  matter  of  the  ability  of 
Hg+"'"  to  form  S — Hg — S  bridges  which  increase  the  stability  of  the  mem- 
brane. These  effects  of  the  mercurials  on  erythrocytes  may  or  may  not  de- 
pend on  metabolic  inhibition,  but  they  are  important  nexvertheless  in  un- 
derstanding the  actions  of  the  mercurials  on  cell  membranes  in  general, 
since  the  mammalian  erythrocyte  presents  an  especially  simple  system  for 
investigation  and  has  been  well  studied. 

Hg++  appears  to  have  greater  lytic  potency  than  the  organic  mercurials. 


THE    CELL   MEMBRANE    AS    A   SITE    FOR   MERCURIAL   ACTION 


901 


Minatoya  et  al.  (1960)  reported  the  ED50  for  the  lytic  action  on  rabbit 
erythrocytes  to  be  0.0034  m.M  for  Hg++  and  0.0174  m.M  for  mersalyl,  and 
Arbuthnott  (1962)  found  that  lysis  can  occur  in  1  hr  with  0.017  vnM  Hg++ 
whereas  it  requires  1  mM  p-MB  or  ethyl-Hg+.  The  effectiveness  depends  on 
the  temperature  and  must  also  depend  on  the  medium  used,  since  a  much 
less  potent  action  of  Hg++  on  rabbit  erythrocytes  was  observed  by  Joyce 
et  al.  (1954),  lysis  occurring  in  2  hr  with  0.13  mikf.  p-MB  is  much  more 
lytic  to  rat  erythrocytes  than  is  PM,  0.1  vaM  of  the  former  lysing  almost, 
completely  in  40-60  min,  whereas  at  this  time  0.5  mM  PM  produces  only 
about  50%  hemolysis  (Moore,  1959),  and  Hg++  is  about  3  times  as  potent 
as  p-MB,  50%  hemolysis  being  given  by  0.4  raM  Hg++  and  1.2  mM  p-MB 
(these  values  estimated  from  data  given)  in  90  min  (Tsen  and  Collier,  1960). 
It  is  obviously  difficult  to  compare  results  obtained  by  different  investiga- 
tors, even  when  the  same  species  is  used,  but  the  definite  difference  in  po- 
tency between  the  various  mercurials  is  clear.  Although  the  role  of  SH 
groups  in  erythrocytic  membrane  structure  and  function  is  important,  ex- 
actly how  they  operate  in  this  capacity  is  unknown,  so  it  is  difficult  to 
speculate  on  either  the  mechanisms  of  hemolysis  by  the  mercurials  or  the 
reasons  for  the  differences  between  the  mercurials.  Other  cells  do  not  lyse 
so  easily  in  the  presence  of  mercurials,  but  this  does  not  necessarily  prove 
that  SH  groups  are  of  more  importance  for  the  erythrocytic  membrane, 
since  the  inherent  stability  may  be  less. 


100 


80 


60 


20 


Fig.  7-39.   Hemolysis  of   human    erythrocytes   by    Hg++   at  different 

fractional  dilution  of  Tyrode  solution  {r).  The  control  curve  shows  the 

hemolysis  in  the  absence  of  Hg++.   (From  Jung,   1947.) 


902  7.  MERCURIALS 

It  is  interesting  to  inquire  into  how  much  Hg++  must  be  bound  to  the 
erythrocytic  membrane  to  cause  hemolysis.  The  data  of  Meneghetti  (1922) 
indicate  about  1.5  X  10'  atoms/cell,  but  Jung  (1947)  believed  this  to  be 
too  low  and  revised  the  figure  on  the  basis  of  his  results  to  1.4  X  10^  atoms/ 
cell.  The  data  of  Vincent  and  Blackburn  (1958)  allow  a  rough  calculation 
that  K+  loss  is  induced  by  Hg++  at  binding  levels  around  2  X  10'  atoms/ 
cell,  although  no  hemolysis  occurs,  while  maximal  K+  loss  and  inhibition 
of  glucose  uptake  in  human  erythrocytes  were  found  by  Weed  et  al.  (1962) 
to  be  produced  by  3.6-4.5  X  10^  atoms/cell.  If  there  are  10  reactive  SH 
groups  for  each  membrane  protein  of  molecular  weight  100,000,  one  can 
estimate  there  to  be  around  3  X  10'  SH  groups  per  erythrocyte  membrane. 
However,  although  stromal  SH  groups  have  a  greater  affinity  for  Hg++, 
hemoglobin  SH  groups  account  for  around  85%  of  the  total  binding,  so  the 
figures  given  above  should  be  reduced  if  only  membrane  binding  is  desired. 
All  one  can  say  is  that  the  amount  of  Hg++  to  alter  membrane  properties 
is  of  the  same  order  of  magnitude  as  the  estimated  SH  content  of  the  mem- 
brane. On  the  other  hand,  the  number  of  molecules/ceU  of  the  organic  mer- 
curials required  for  hemolysis  is  greater  than  necessary  to  cover  the  sur- 
face of  the  sheep  erythrocyte  (Benesch  and  Benesch,  1954).  For  PM  there 
is  a  4-fold  excess  and  for  mersalyl  a  24-fold  excess.  Of  course,  the  organic 
mercurials  probably  do  not  lie  flat  on  the  membrane,  but,  more  important, 
it  is  not  known  how  much  of  the  mercurial  is  bound  to  hemoglobin  or  other 
nonmembrane  components. 

The  kinetics  of  mercurial  hemolysis  are  generally  characterized  by  a  lag 
period,  the  duration  of  which  is  dependent  on  the  mercurial  concentration, 
followed  by  a  rather  sudden  hemolysis  (Fig.  7-40).  For  sheep  erythrocytes 
there  is  a  lag  period  of  around  80  min  when  treated  with  0.45  milf  PM 
(Benesch  and  Benesch,  1954),  and  for  human  erythrocytes  the  lag  period 
is  90  min  at  37^  when  exposed  to  0.5  mM  p-MB  (Sheets  et  al,  1956  a).  The 
temperature  is  an  important  factor,  since  in  the  latter  case  the  lag  period 
is  around  200  min  at  25°.  The  lag  period  is  partly  due  to  the  slow  binding 
of  these  organic  mercurials.  Washing  the  erythrocytes  1  min  after  exposure 
to  p-MB  protects  completely,  after  30  min  protects  partially,  and  after 
60  min  there  is  no  protection  (Sheets  et  al.,  1956  a).  On  the  other  hand, 
there  is  maximal  uptake  of  Hg^"^  by  erythrocytes  within  5  min  (Weed  et 
al.,  1962).  The  osmotic  fragility  is  altered  after  3-min  exposure  to  Hg++: 
At  5.2  X  10'-5.5  X  10^  atoms/cell  there  is  a  decrease  in  the  fragility,  at 
4.5  X  10^  atoms/cell  there  is  an  increased  fragility  and  hemolysis.  The  ki- 
netics for  Hg++  and  the  organic  mercurials  appear  to  be  quite  different.  The 
very  marked  effects  of  PM  concentration  on  the  kinetics  of  hemolysis  may 
be  seen  in  Fig.  7-40  for  sheep  erythrocytes,  and  a  similar  dependence  has 
been  noted  in  rat  erythrocytes  (Moore,  1959).  This  might  indicate  a  rather 
critical  level  of  membrane  binding  to  produce  hemolysis.  The  uptake  of 


THE    CELL   MEMBRANE   AS   A   SITE    FOR   MERCURIAL   ACTION 


903 


Hg+''"  by  erythrocytes  or  ghosts  is  very  rapid  but  the  situation  with  chlor- 
merodrin  is  different,  in  that  binding  to  ghosts  is  rapid  but  uptake  into 
erythrocytes  continues  for  2  hr  or  more;  the  binding  of  chlormerodrin  is 
also  perhaps  more  specific  for  certain  SH  groups  (Rothstein,  1964).  Chlor- 


100 


Fig.  7-40.  Hemolysis  of  sheep  erythrocytes  in  a  2%  suspension  by  PM 
at  pH  7.4  and  37°,  showing  the  marked  differences  over  a  narrow  con- 
centration range.   (From  Benesch  and  Benesch,   1954.) 


merodrin  thus  might  be  useful  in  separating  the  effects  of  membrane  and 
internal  binding. 

Is  hemolysis  by  the  mercurials  in  any  way  related  to  effects  on  glucose 
uptake  or  metabolism?  This  question  cannot  be  satisfactorily  answered 
since  there  has  been  little  work  where  metabolic  and  hemolytic  actions  can 
be  compared,  and  the  results  available  are  divergent.  The  utilization  of 
glucose  by  human  erythrocj'tes  is  inhibited  moderately  within  a  range  of 
Hg++  concentration,  the  inhibition  disappearing  as  the  amount  of  Hg++ 
bound  is  increased  (Fig.  7-41)  (Weed  et  al.,  1962).  In  the  reversal  range,  a 
change  in  the  hemoglobin  was  observed  and  some  agglutination  of  the  cells 
occurred.  The  question  arises  as  to  whether  the  effects  on  K+  loss  and  glu- 
cose utilization  result  from  some  action  on  all  the  cells  or  are  due  to  hemol- 
ysis of  a  few  cells.  Weed  et  al.  (1962)  assumed  that  high  Hg++  concentra- 
tions denature  the  membrane,  causing  a  decrease  in  permeability,  which 
could  explain  the  reversal  of  the  effect  on  K+  loss,  but  is  difficult  to  reconcile 
with  the  disappearance  of  the  effect  on  glucose  utilization.  It  is  interesting 
to  compare  their  results  on  osmotic  fragility  with  these  actions.  At  the  Hg++ 


904 


7.  MERCURIALS 


level  producing  the  maximal  K+  loss  and  inhibition  of  glucose  utilization 
(6.0-7.5  X  10"^^  mole  Hg/cell),  it  is  claimed  that  the  fragility  is  decreased, 
which  it  is  in  very  hypotonic  media,  but  examination  of  the  curves  shows 
that  some  hemolysis  (probably  around  2-5%)  has  occurred  in  normal  me- 
dium. At  a  higher  level  of  bound  Hg++,  where  the  effects  on  K+  and  glu- 
cose have  been  partly  reversed  (7.6  X  10~^^  mole  Hg/cell),  fragility  is 
definitely  increased,  and  some  20%  hemolysis  has  occurred  in  the  normal 
medium  within  3  min.  It  seems  clear,  therefore,  that  the  K+  loss  and  sup- 
pression of  glucose  utilization  at  lower  levels  of  bound  Hg++  (left  of  the 


50- 


40 


30 


20 


10 


1000 


MOLES     Hg     BOUND/RBC    «  10 


Fig.  7-41.   Effect  of  Hg++  on  the  loss  of  K+  from  human  erythrocytes 
and  the  uptake  of  glucose.   (From  Weed  et  al.,  1962.) 


maximum  in  Fig.  7-41)  are  due  almost  entirely  to  effects  on  all  the  cells 
and  not  on  lysis  of  a  fraction  of  the  cells.  However,  at  higher  levels  of 
bound  Hg++,  lysis  must  contribute  to  both  K+  loss  and  interference  with 
glucose  utilization;  e.g.,  at  7.6  X  10~^^  mole  Hg/cell  bound,  there  is  20% 
lysis  and  around  20%  loss  of  K+.  If  this  is  so,  the  disappearance  of  the 
effect  on  K+  loss  from  intact  cells  must  occur  even  more  precipitously  than 
appears  in  the  figure  (the  reversal  is,  of  course,  really  not  precipitous,  since 
it  is  a  logarithmic  scale;  to  reverse  these  effects  appreciably  it  requires  the 
binding  of  about  10  times  that  amount  of  Hg++  necessary  for  maximal  K+ 
loss).  The  results  of  Jacob  and  Jandl  (1962),  also  on  human  erythrocytes, 
are  quite  different,  since  they  observed  that  p-MB  does  not  inhibit  glucose 
uptake  or  lactate  formation  —  indeed,  stimulates  these  somewhat  —  up  to 
5  //moles  p-MB/ml  of  cells,  which  is  around  2  x  10"^^  mole  2>-MB/cell. 
There  is  also  no  reaction  of  p-MB  with  the  intracellular  glutathione.  These 
results  point  to  a  failure  of  p-MB  to  penetrate  through  the  membrane. 
Hemolysis  occurs  and  is  presumably  due  to  an  action  on  the  membrane. 
However,  it  is  rather  strange  that  sufficient  mercurial  can  be  bound  to  the 


THE    CELL   MEMBKANE    AS   A   SITE    FOR    MERCURIAL   ACTION  905 

membrane  to  cause  lysis  and  yet  have  no  inhibitory  effect  on  glucose  uptake. 
One  might  conclude  that  Hg++  penetrates  into  the  erythrocytes  more  readi- 
ly than  2?-MB,  which  is  undoubtedly  the  case,  but  it  is  also  possible  to  spec- 
ulate that  the  bifunctional  Hg++  can  distort  the  membrane  pores  by  form- 
ing S — Hg — S  bridges  in  such  a  way  that  glucose  penetration  is  slowed 
while  K+  permeability  is  increased,  as  in  the  concept  of  critical  pore  sizes 
formulated  by  Mullins  (1960).  Another  problem  is  how  these  results  can 
be  reconciled  with  those  of  LeFevre  (1948),  who  showed  that  glucose  utiliz- 
ation by  human  erythrocytes  is  inhibited  by  0.002  mM  p-MB  and  abolished 
by  0.01  mM. 

The  very  rapid  loss  of  K+  from  human  erythrocytes  observed  by  Weed 
et  al.  (1962)  —  maximally  50%  of  the  total  cell  K+  in  3  min  at  7.5  X  10-i« 
mole  Hg/cell  —  is  not  seen  in  rabbit  erythrocytes,  from  which  there  is  a 
slow  loss  of  K+  in  the  presence  of  Hg++,  a  rapid  loss  occurring  only  upon 
hemolysis,  a  result  which  is  quite  reasonable  (Joyce  et  al.,  1954).  It  is  dif- 
ficult to  compare  the  results  of  these  two  groups  of  investigators  because 
the  Hg++  concentrations  are  expressed  differently.  However,  it  is  possible 
to  estimate  that  when  7.5  X  10~^^  mole  Hg/cell  is  bound,  the  free  Hg++ 
concentration  is  roughly  0.01  mM  (see  Fig.  2  of  Weed  et  al.).  In  the  work 
of  Joyce  et  al.,  0.032  mM  Hg++  caused  a  30%  loss  of  total  K  in  4  hr,  so 
that  apparently  there  is  a  very  marked  difference  in  the  response  of  rabbit 
and  human  erythrocytes  to  Hg++. 

The  question  as  to  the  relation  of  glucose  metabolism  to  hemolysis  is 
still  unanswered.  There  is  one  observation  which  suggests  a  relation,  the 
finding  by  Moore  (1959)  that  10-100  roM  glucose  inhibits  the  hemolysis 
induced  by  p-MB,  this  being  manifest  mainly  in  a  lengthening  of  the  lag 
period.  No  reaction  between  j^-MB  and  glucose  can  be  detected  spectrosco- 
pically  and  it  is  not  an  osmotic  effect.  Certain  other  sugars,  e.g.  fructose 
and  sorbose,  are  also  effective.  It  was  postulated  that  glucose  may  combine 
with  some  component  of  the  rat  erythrocyte  and  protect  it  from  the  mer- 
curial; if  so,  this  would  probably  be  the  transport  system  for  glucose,  which 
is  inhibited  readily  by  p-MB,  the  situation  being  similar  to  the  protection 
of  enzymes  by  substrates.  Sheets  et  al.  (1956  a)  had  found  that  glucose  exerts 
no  protection  against  hemolysis  of  human  erythrocytes  by  p-MB,  but  the 
glucose  was  added  at  various  times  after  the  p-MB  and  was  only  3.3  mM. 

Membrane  or  transport  ATPase  is  inhibited  by  the  mercurials  but  the 
Hg++  or  chlormerodrin  which  is  initially  bound  is  without  effect  (Roth- 
stein,  1964).  Chlormerodrin  can  bind  to  about  3%  of  the  total  SH  groups 
without  inhibition  of  ATPase,  but  by  the  time  of  maximal  binding  with 
25%  of  the  SH  groups  the  ATPase  is  inactivated,  possibly  leading  to  the 
loss  of  K+,  although  some  increase  in  permeability  may  also  play  a  role. 

The  possibility  that  mercurial  hemolysis  is  related  to  the  reaction  of 
erythrocytic  glutathione  with  SH  reagents  was  considered  by  Tsen  and 


906  7.  MERCURIALS 

Collier  (1960).  However,  Hg++  and  p-MB  can  produce  hemolysis  without 
significant  loss  of  glutathione,  whereas  iodoacetate  and  iV-ethylmaleimide 
reduce  the  glutathione  completely  without  lysis.  Jacob  and  Jandl  (1962) 
also  showed  that  p-M3  reacts  readily  with  glutathione  in  solution,  but  does 
not  attack  erythrocyte  glutathione,  and  concluded  that  p-MB  does  not 
penetrate  into  the  cells.  However,  Weed  et  at.  (1962)  found  that  of  the  three 
major  sources  of  SH  groups  in  the  erythrocyte  —  stroma,  hemoglobin,  and 
glutathione  —  the  last  has  the  lowest  affinity  for  the  mercurial  and  consti- 
tutes only  5%  of  the  total  SH  groups.  It  is  thus  possible  that  glutathione 
would  be  reacted  only  when  aU  the  other  SH  groups  are  saturated.  In  any 
event,  it  is  evident  that  glutathione  does  not  play  a  significant  role  in 
hemolysis. 

Hg++  is  able  to  produce  structural  changes  in  the  erythrocytic  membrane 
which  are  detectable  by  electron  microscopy  (Jung,  1947).  Isolated  hemo- 
globin-free membranes  treated  with  high  concentrations  of  Hg++  (37  ulM) 
show  gross  changes  in  structure  —  a  crumpling  with  increased  density  and 
apparent  thickness  —  but  with  lower  concentrations  (0.37  vaM)  the  pic- 
ture is  different,  a  network  of  holes  appearing  in  the  otherwise  unaltered 
membrane.  Intact  erythrocytes  treated  with  1.85  ruM  Hg++  for  several 
hours  no  longer  lyse  in  distilled  water,  and  the  membrane  is  seen  to  have 
been  replaced  by  a  thick  mass  of  coagulated  protein.  Certain  changes  in 
the  over-all  erythrocyte  configuration  were  also  observed  by  Vincent  (1958) 
in  preparations  allowed  to  bind  Hg++  for  5  min,  especially  sphering  and 
crenation.  Possibly  a  more  detailed  study  of  structural  changes  induced  by 
low  prohemolytic  concentrations  of  the  mercurials  would  be  useful  in  clari- 
fying the  mechanism  of  hemolysis. 

We  have  assumed  with  others  in  this  discussion  of  hemolysis  and  permea- 
bility changes  brought  about  by  the  mercurials  that  SH  groups  only  are 
attacked.  Certain  nonelectrolytes,  such  as  glucose  and  glycerol,  enter  the 
erythrocyte  by  facilitated  diffusion  and,  since  the  transport  is  usually  ef- 
fectively blocked  by  SH  reagents,  it  has  been  thought  that  SH  groups  are 
involved  in  some  manner.  We  have  seen  that  Wilbrandt  (1941)  claimed  a 
marked -reduction  in  glycerol  permeability  with  0.05  vaM  Hg++.  Further- 
more, it  was  believed  that  inhibition  of  glycerol  penetration  occurs  only 
while  the  Hg++  is  entering  the  cells,  i.e.,  when  the  Hg++  is  bound  to  the 
membrane.  When  the  Hg++  has  been  picked  up  by  the  hemoglobin,  there 
may  be  little  left  in  the  membrane  and  the  permeability  to  glycerol  is  re- 
stored. LeFevre  (1948)  established  that  p-MB  likewise  blocks  glycerol  entry 
into  the  human  erythrocyte.  However,  Barnard  and  Stein  (1958)  have  sug- 
gested that  an  imidazole  group  is  involved  in  this  transport.  The  fact  that 
histidine  as  well  as  cysteine  can  reverse  the  inhibition  (it  requires  a  5-  to 
10-fold  excess  of  histidine)  is  not  valid  evidence;  it  simply  shows  that  mer- 
curials are  bound  to  histidine.  It  was  also  claimed  that  mercurial  action 


EFFECTS    ON    PERMEABILITY    AND    ACTIVE    TRANSPORT  907 

was  characterized  by  a  lag  period,  this  being  due  to  the  preferential  binding 
to  SH  groups  before  the  imidazole  groups  are  attacked;  however,  there  are 
other  possible  reasons  for  such  a  lag  period,  and  indeed  Wilbrandt  (1941) 
claimed  the  inhibition  occurs  before  the  mercurial  is  bound  intracellularly. 
They  also  point  out  that  Cu++  is  much  more  potent  than  p-MB  in  depressing 
glycerol  entry  and  that  this  favors  an  imidazole  group;  p-MB  is,  however, 
rather  ineffective  relative  to  Hg++,  which  exerts  an  effect  at  0.025  mM,  due 
possibly  to  steric  factors.  It  does  not  appear  that  the  evidence  is  sufficient 
to  establish  an  imidazole  group  as  involved  in  the  glycerol  transport,  but 
one  cannot  argue  against  this  theory,  and  it  is  quite  possible  that  in  the 
complex  mechanisms  of  penetration  there  are  both  SH  and  imidazole  groups. 
In  either  case,  one  cannot  attribute  an  active  role  to  these  groups  in  the 
transport  on  the  basis  of  the  evidence  available. 


EFFECTS   ON    PERMEABILITY  AND  ACTIVE   TRANSPORT 

The  general  discussion  of  the  mechanisms  by  which  transport  systems 
in  the  membrane  may  be  affected  by  SH  reagents  (see  III-1-171,  180)  is 
applicable  to  the  mercurials.  We  shall  confine  our  attention  to  certain 
important  problems  and  interesting  results,  as  far  as  possible,  and  only 
summarize  most  of  the  studies  in  Table  7-18.  The  effects  of  the  mer- 
curials on  renal  transport  wiU  be  taken  up  in  the  following  section.  It 
is  clear  from  the  results  in  the  table  that  the  mercurials  often  cause  a  loss 
of  intracellular  substances,  e.g.,  K+,  carbohydrate,  and  amino  acids.  It  is 
likely  that  a  good  many  substances  leak  out  of  cells  treated  with  the  mer- 
curials as  a  result  of  not  only  interference  with  active  transport  but  direct 
distorting  effects  on  the  membrane  structure  leading  to  increases  in  per- 
meability. Possibly  coenzymes,  enzymes,  and  other  large  molecules  may  be 
lost.  Ohr  (1960)  observed  the  release  of  some  ultraviolet-absorbing  material 
from  diaphragm  exposed  to  Hg++,  and  Weed  et  al.  (1962)  detected  the  early 
release  of  some  Hg++-complexing  material  from  human  erythrocytes,  this 
altering  the  binding  kinetics  at  low  concentrations  of  Hg++.  It  is  not  easy 
to  determine  if  the  action  is  primarily  on  active  uptake  or  on  outward  dif- 
fusion, even  with  labeled  substances.  For  example,  if  one  is  studying  K^^ 
efflux,  an  inhibition  of  a  pump  involved  in  maintaining  a  high  intracellular 
K+  level  might  alter  this  efflux,  either  by  changing  the  membrane  potential 
or  directly  if  part  of  the  K+  efflux  is  mediated  by  the  pump,  since  a  Na+ 
pump  might  not  be  completely  specific  for  Na+  and  might  carry  some  K+ 
out  of  the  cell.  If  K*^  influx  is  measured,  an  alteration  of  the  permeability 
could  easily  change  the  rate  at  which  active  transport  occurs,  particularly 
if  K+  loss  accelerates  the  pump.  In  most  cases  there  seems  to  be  a  decrease 
in  the  intracellular  K+/Na+  ratio,  but  the  mechanism  is  not  clear.  Further- 
more, a  decrease  in  transport  is  occasionally  not  accompanied  by  a  signifi- 


908 


7.  MERCURIALS 


H 


03 


ii 

o 

^ 

O 

<B 

o  -^ 

c 
o 
o 


'd 

iV 

o 

u 

fi 

O 

cS 

P< 

-tJ 

CO 

03 

C 

-fl 

go       N 
O   CO       c 


Ph 


M 


pq 


(M    CD 

d  d 


p. 


^     O 


OS 

s 

05 

1— 1 

^ 

_ 

S 

-3 
c 

o3 

s 

cS 

73 

'-s 

05 

T3 

C! 

>> 

00 

S 

a> 

C 

IB 

»o 

-Q 

"& 

a 

c 

02 

O 

o 

W 

O 

> 

>-» 

^.   o 


+ 

+ 

pq 

pq 

+ 

S 

S 

W 

W 

?i. 

a 

W 


pq 


a-    Q, 


(^    P5 


w 


o 

gj 

o 

CQ 

r^ 

s 

0) 

c 

« 

o 

O 

"*- 

■~- 

•<* 

's 

lO 

s 

d 

(n' 

pq 


W 


W 


M 
S 


w 


EFFECTS    ON    PERMEABILITY    AND    ACTIVE    TRANSPORT 


909 


O 


— 

c 

u 
o 

c« 

-1-2 

-tj 

o 

-ti 

o 

o 

p- 

o 

CC 

03 

M 

«,•    ''^ 

<M    rNi 

^   o 

1 

1       ' 

w  ^ 

(M 

o 

o  o 

05 

ir^ 

^  »o 

p. 


& 


+  + 


+ 


o 
o 

T3 
c3 

0) 

a 

es 

OJ 

H 

in 

o 
CO 

O 

02 

■T! 

"c 

T3 

C 

ci 

>» 

cS 

kl 

Id 

c 

OJ 

J= 

rft 

.^ 

M   (M 

>o 

C   S 

o  -=  C5    -s 


»  -c  Xi 


■*    M       l>    t— 


I        + 


C<1 

o 

t- 

■* 

^H 

i>-l 

o 

1— 1 

lO 

00 

^-1 

o 

^^ 

^^ 

o 

i-H 

f—* 

o 

o 

o 

o 

o 

'^ 

»c 

o 

o 

o 

o 

o 

o 

-- 

o 

-^ 

d 

d 

d 

d 

d 

-H 

d 

d 

d 

-H 

d 

d 

d 

d 

d 

d 

d 

d 

+  m 

CQ 

« 

pa 

tc.Sg 

S 

S 

S 

ffi    A  CM 

a 

?5, 

a, 

M 

w 


ffl 


o 

V. 

o 
a 

2 

cS 

•s. 

^ 

t-  'E 

-e  "E 

v. 
^ 

'3 

4)   -o 

S    T3 

?;, 

c 

>    c 

«     C 

u 

—    o 

-C     O 

e 

■S  -2 

1 

a 

f0  s 

s   p 

o 

tf 

05    - 

a. 

S      c 


ftS     §     05 


05 


05 


o 


910 


7.  MERCURIALS 


P5 


O 


Ti 

a> 

11 

o 

ft 

c 

;=; 

x> 

3 

.s    s 


m 


w 


p^    P5 


tH 

^ 

C5 

a 

£ 

;;i 

o3    -—. 

« 

" 

Si    CD 

ft 

73 

S 

-2 

M    i-H 

0 

_C 

"o 

C    ■ — • 

M 

7T1 

,lH 

Pm 


_w)  ill 


o 


00     (N 


o 

d      -H 


m 


w 


P5 


c   o 


O 

PM 


M 


— 

T3 

O! 

03 

^ 

^ 

eg 

o      -S  -'. 


00 


^     o   o     o   o     o 


CO 

m 

PL, 

+ 

+ 

PP 

g 

§ 

h  s 

S 

a 

?i, 

W      PlH 

ffi 

i, 

P5 


o   o 

7  2 


o    -^ 

d   d 


W 


EFFECTS    ON    PERMEABILITY    AND    ACTIVE    TRANSPOET 


911 


?s   -^ 


pq 


^-~  o 


W 


<0  « 


fa 

^4 


>  ^ 

CO  lO 

ft,  05 

h-) 


05 

o 
c 
o 

►-5 

^_^ 

.^ 

^-^ 

O 

<» 

S 

-* 

CO 

CO 

§ 

d 

1 

1 

1 

o   o   o 
-^   o 


\    ft 


o   -^  o 

o   cS   ~^     d 


O        -H 

6  <~>     '^ 


ffi 


m 


m 


pq 


m 


s  .s 

o  -o 
IS    o 


a;    « 


I        I 


P5 


(^ 


ap 


tf 


P5 


P5 


o    fl 


X 


+ 


912  7.  MEECURIALS 

cant  depression  of  respiration,  as  in  the  uptake  of  I"  by  Fucus,  where  0.05 
YoM  p-MB  inhibits  transport  50%  but  0.2  mM  does  not  affect  respiration 
(Klemperer,  1957),  or  the  accumulation  of  K+  by  Porphyra,  where  p-MB 
decreases  the  number  of  ions  pumped  per  Og  consumed  (Eppley,  1960),  or 
the  active  transport  of  Na+  through  frog  skin,  which  is  blocked  by  Hg++ 
at  a  concentration  not  altering  respiration  (Linderholm,  1952).  In  such 
cases  it  has  generally  been  assumed  that  the  action  is  on  the  transport 
system  itself,  but  this  is  not  necessarily  true.  It  is  rather  surprising  that  so 
few  have  reported  instances  of  decreased  permeability  brought  about  by 
the  mercurials,  particularly  the  organic  ones,  inasmuch  as  they  might  be 
expected  to  react  with  SH  groups  in  or  around  the  membrane  pores  to 
impede  the  passage  of  substances  across  the  membrane;  perhaps  this  would 
be  observed  more  often  if  lower  concentrations  were  examined.  The  per- 
meability of  frog  skin  to  Cl~  is  decreased  by  p-MB,  and  Janacek  (1962) 
postulated  that  the  mercurial  sterically  hinders  the  movement  of  anions 
through  the  pores. 

Certain  results  occasionally  point  to  an  effect  of  mercurials  on  the  end- 
ergonic  phase  of  transport  rather  than  a  simple  depression  of  ATP  forma- 
tion. The  fact  that  Hg+^  at  0.5  mM  inhibits  the  20-sec  uptake  of  acetate 
by  diaphragm  without  a  lag  period  (Foulkes  and  Paine,  1961),  taken  with 
the  rather  slow  depression  of  metabolism,  is  indicative  of  an  action  directly 
on  the  membrane.  We  have  also  seen  that  p-MB  lyses  erythrocytes  without 
reacting  with  intracellular  glutathione  or  inhibiting  glycolysis,  in  contrast 
to  A^-ethylmaleimide,  and  that  this  has  been  attributed  to  a  failure  to  pene- 
trate into  the  cells,  so  that  the  effects  observed  must  involve  an  attack  on 
the  membrane  (Jacob  and  Jandl,  1962).  Hg++  very  potently  inhibits  stro- 
mal ATPase  —  50%  inhibition  at  around  0.00125  mM,  and  plots  of  log 
{vjvi  —  1)  against  log  (Hg++)  suggest  that  3  Hg++  ions  are  required  for 
each  ATPase  molecule  (Laris  et  al.,  1962).  The  same  type  of  behavior  was 
observed  for  the  inhibition  of  glucose  uptake,  and  the  concentrations  of 
Hg++  required  to  inhibit  are  comparable  (LeFevre,  1954).  Laris  et  al.  re- 
plotted  LeFevre's  data  and  found  that  roughly  6  ions  of  Hg++  are  necessary 
for  the  inhibition  of  the  transport  of  each  glucose  molecule.  The  similar 
sensitivities  and  kinetics  allowed  them  to  postulate  that  the  two  inhibi- 
tions may  be  closely  related.  The  stimulation  of  the  uptake  of  certain  sugars 
(e.g.,  D-xylose  and  L-arabinose)  into  diaphragm  by  p-MB,  and  the  inhibition 
of  the  stimulation  produced  by  insulin,  may  well  be  on  the  muscle  mem- 
branes (Kono  and  Colowick,  1961).  There  is  certainly  no  correlation  with 
the  level  of  ATP,  and  p-MB  actually  seems  to  increase  ATP  somewhat.  A 
block  between  ATP  and  the  transport  system  was  considered  a  possibility. 

The  reaction  of  mercurials  with  a  membrane  carrier  was  adduced  to  ex- 
plain the  inhibition  of  phosphate  transport  in  Micrococcus  pyogenes  by  Hg++ 
and  PM  (P.  Mitchell,  1953).  An  inhibition  of  50%  is  given  by  2.2  //moles 


EFFECTS    ON    PERMEABILITY    AND    ACTIVE    TRANSPORT  913 

PM/g  cells  and  by  4.7  //moles  Hg++/g  cells.  The  inhibition-concentration 
curves  are  said  to  conform  to  the  equation  ^  =  (I)  (1  —  i)ji  and  hence  to 
suggest  reaction  of  the  mercurial  with  a  phosphate  carrier  X,  according  to 
I  +  X  :±5:  IX.*  This  equation  is  simply  that  for  noncompetitive  inhibition 
and  it  is  difficult  to  understand  how  it  would  serve  to  indicate  any  partic- 
ular mechanism  by  which  transport  is  depressed.  Mitchell  then  proceeds 
to  calculate  the  number  of  carrier  molecules  for  100  molecules  of  intra- 
cellular phosphate;  since  the  cells  contained  147  //moles  P,/g  cells,  1.5  and 
3.2  molecules  of  PM  and  Hg++,  respectively,  are  required  for  50%  inhibition 
per  100  molecules  of  internal  P^.  It  was  apparently  assumed  that  if  50% 
inhibition  is  given  by  these  numbers  of  molecules,  100%  inhibition  would 
be  given  by  twice  these,  namely,  3.0  and  6.4  molecules  of  PM  and  Hg++, 
respectively.  If  this  were  true  titration  or  zone  C  inhibition,  this  would  be 
correct,  but  inspection  of  the  curves  shows  that  it  is  not;  indeed,  it  is  evi- 
dent that  approximately  20  and  40  //moles/g  cells  of  PM  and  Hg++  are 
needed  for  90%  inhibition  (curves  do  not  reach  complete  inhibition,  which 
would  require  appreciably  more  of  the  mercurials).  Therefore,  his  conclusion 
that  the  number  of  carriers  is  not  more  than  3%  of  the  internal  P,  is  not 
valid.  In  addition,  the  mercurials  must  be  bound  to  cell  components  other 
than  a  hypothetical  carrier,  so  that  under  any  circumstances  it  would  be 
difficult  to  estimate  the  relative  amount  of  carrier  present,  just  as  it  is 
impossible  to  calculate  the  amount  of  an  enzyme  present  in  a  complex 
mixture  by  the  quantity  of  mercurial  required  for  50%  inhibition. 

Transmembrane  and  Transcellular  Transports 

The  uptake  of  a  substance  into  a  cell  is  often  a  process  different  from 
the  transport  across  a  layer  of  the  cells.  If  a  substance  is  moved  against 
a  concentration  gradient  from  one  medium  into  a  similar  medium,  it  is 
an  active  transport,  whereas  accumulation  of  a  substance  within  a  cell  can 
be  the  result  of  binding.  A  good  example  of  this  is  the  transport  of  triiodo- 
thyroacetate  by  rat  intestine  (Herz  et  al.,  1961).  The  mucosal  -^  serosal 
transport  is  inhibited  93%  by  1  raM  Hg++,  but  the  accumulation  in  the 
tissue  is  actually  accelerated  16%.  The  cellular  uptake  was  postulated  to 
be  due  mainly  to  binding.  The  accumulation  of  Fe+++  (Saltman  et  al,  1955) 
and  Cu++  (Saltman  et  al.,  1959)  by  rat  liver  slices  is  slightly  stimulated  by 
p-MB,  and  it  is  very  likely  that  these  are  instances  of  binding  to  intracel- 
lular ligands.  There  is  sometimes  not  so  clear  a  separation  of  transmembrane 
and  transcellular  transports.  Rat  intestinal  slices  accumulate  Ca++  to  a  tis- 

*  Mitchell  gives  the  equation  as  ^  =  (I)i7(l  —  i),  changing  his  symbols  to  those 
used  in  the  present  work,  which  is  obviously  incorrect,  since  it  would  mean  that  the 
inhibition  would  vary  inversely  with  the  inhibitor  concentration,  so  I  have  taken  the 
liberty  of  rewriting  it. 


914  7.  MERCURIALS 

sue/medium  ratio  of  5.8,  and  1  mM  Hg++  reduces  this  to  1.5  (Schachter 
et  al.,  1960).  The  transport  across  the  intestinal  wall  leads  to  an  inside- 
outside  ratio  of  4.6,  and  1  mM  Hg++  drops  this  to  1.1  (Schachter  and  Rosen, 
1959).  Thus  in  this  instance  there  is  no  significant  difference  in  the  mercurial 
inhibition,  but  certain  other  inhibitors  affect  the  transintestinal  process 
more  strongly.  The  results  wiU  often  depend  on  the  location  of  the  active 
transport  mechanism.  We  shall  see  that  this  is  an  important  point  in  con- 
sidering the  effects  of  the  mercurials  on  the  kidney. 

Mitochondrial    K+ 

Mitochondria  isolated  from  rat  liver  in  0.25  M  sucrose  contain  620-640 
-//moles  K+/g  N  and  this  can  be  lost  if  the  mitochondria  are  placed  in  hypo- 
tonic media  or  treated  with  saponin,  2,4-dinitrophenol,  or  Hg+"'"  (Spector, 
1953).  Most  of  the  K+  appears  to  be  free,  but  a  fraction  may  be  bound;  the 
retention  of  the  free  K+  is  dependent  on  oxidative  phosphorylation.  How- 
ever, Gamble  (1957)  found  that  mitochondrial  fragments  catalyzing  oxida- 
tive phosphorylation  can  bind  K+.  This  binding  is  not  dependent  on  ATP 
but  is  abolished  by  2.4-dinitrophenol,  and  a  relation  between  the  K+  bind- 
ing and  the  sites  for  oxidative  phosphorylation  was  postulated.  Hg++  at 
0.01  mM  and  p-MB  at  0.03  mM  produce  a  5-fold  increase  in  the  K+  ex- 
change rate.  This  was  later  investigated  in  detail  (Scott  and  Gamble,  1961), 
and  Hg++  in  concentrations  around  0.01  mM  was  found  to  increase  the  ex- 
change rate,  reduce  the  bound  K+  by  50%,  and  inhibit  phosphorylation 
50%.  The  organic  mercurials  are  less  effective.  These  three  actions  are  pre- 
sumably not  mediated  through  the  same  mechanism,  since  EDTA  prevents 
the  effects  of  Hg++  and  p-MB  on  K+  binding,  has  no  effect  on  the  stimulation 
of  the  exchange  rate,  and  protects  oxidative  phosphorylation  from  Hg++ 
but  not  from  p-MB.  These  complex  relationships  are  not  completely  under- 
stood at  the  present  time,  but  obviously  are  of  importance  in  certain  cases 
of  K+  accumulation  and  transport. 

Gastric  Acid  Secretion 

Reduction  of  gastric  acidity  by  0.25  mM  Hg++  introduced  into  the  stom- 
ach was  shown  by  Mann  and  Mann  (1939),  and  the  mechanisms  involved 
were  studied  by  Davenport  and  his  group  at  Utah.  Gastric  secretion  of 
HCl  is  depressed  to  a  basic  level  by  1  mM  p-MB;  if  the  secretion  is  stimul- 
ated by  carbachol  or  histamine,  the  inhibition  appears  to  be  greater  but 
the  rate  is  reduced  to  the  same  level  (Fig.  7-42),  i.e.,  p-MB  effectively  abol- 
ishes the  secretion  brought  about  by  these  drugs  (Davenport,  1954,  Da- 
venport et  al.,  1954).  There  is  thus  a  basal  level  of  secretion  (around  30% 
of  maximal)  resistant  to  the  mercurials.  Graphical  analysis  indicated  that 
2  SH  groups  are  involved  in  the  inhibition.  Lactate  formation  when  glu- 


EFFECTS    ON    PERMEABILITY    AND    ACTIVE    TRANSPORT 


915 


cose  is  the  substrate  is  inhibited  by  p-MB,  one  SH  group  being  involved 
here,  but  it  is  unlikely  that  glycolytic  inhibition  is  the  mechanism  by  which 
acid  secretion  is  depressed,  inasmuch  as  inhibition  occurs  when  pyruvate  or 
acetoacetate  is  the  substrate.  Respiration  associated  with  secretion  is  also 
inhibited  by  p-MB;  it  was  felt  that  this  is  not  a  generalized  effect  on  oxi- 


0  72 


0  63 


0.54 


0.45 


0  27 


0  09 


CARBACHOL    (0  1    MG  X) 


0.2      0.4     0.6      0.8      1.0       1.2       1.4       18 


Fig.  7-42.  Inhibition  of  gastric  acid  secretion  by 

p-MB,  in  the  presence  of  20  raM  glucose  and  in 

the  absence  and  presence  of  carbachol.    (From 

Davenport  et  al.,  1954.) 


dative  reactions,  but  that  the  site  of  attack  is  some  unknown  system  inti- 
mately concerned  with  the  secretory  process.  The  relation  between  the  inhi- 
bitions of  respiration  and  secretion  by  p-MB  is  reasonably  linear  (Fig.  7-43), 
in  contrast  to  the  results  with  antimycin  and  2,4-dinitrophenol  (Davenport 
and  Chavre,  1956).  Possibly  the  primary  inhibition  by  j^-MB  is  on  the  trans- 
port system  itself,  the  respiration  being  reduced  secondarily.  Other  SH 
reagents  inhibit  secretion  but  apparently  act  at  somewhat  different  sites 
than  p-MB  (Davenport  et  al.,  1955),  so  that  it  is  difficult  to  correlate  the 
blockade  of  SH  groups  with  the  secretory  suppression  or  to  locate  exactly 
these  SH  groups. 


916 


7.   MERCURIALS 


Intestinal  Transport 

The  transports  of  Na+,  water,  and  glucose  across  the  rat  intestine  are 
inhibited  by  Hg++  (Clarkson  and  Cross,  1961).  The  transintestinal  electric 
potential  is  dependent  on  the  Na+  transport  and  the  ionic  permeabilities 
of  the  lumenal  membranes.  Hg++  0.01-1  mM  causes  a  rapid  brief  elevation 
of  the  potential  which  is  followed  by  a  fall,  the  rapidity  of  which  is  deter- 
mined by  the  Hg++  concentration.  There  are  two  phases  in  the  response: 
(1)  an  immediate  loss  of  K+  and  phosphate  from  the  intestinal  wall  and  a 
marked  inhibition  of  glucose  uptake,  and  (2)  a  delayed  (occurring  after 
20  min  or  longer)  inhibition  of  transintestinal  transport  of  Na+,  water,  and 

0  4 


0.3 


0.2 


0.1 


\ 

-^/ 

ANTIMYCIN        / 

/p-MB 

) 

"                // 

' 

^/dnp 

'    ^  -^ 

.^ 

0.1 


0.2 


03  0.4  0.5  0.6  0.7  0.8 


Fig.  7-43.  Effects  of  inhibitors  on  the  respiration  and  acid  secretion 

of  mouse  gastric  mucosa  stimulated  by  carbachol.  (From  Davenport 

and  Chavre,  1956.) 


glucose,  with  a  suppression  of  lactate  formation.  The  uptake  kinetics  of 
Hg++  show  two  phases,  a  fast  component  dominant  during  the  initial  20- 
30  min  of  exposure  (k^  =  0.0032  min'^)  and  a  slower  component  (^2  — 
0.0017  min-^).  It  was  pointed  out  that  the  system  is  so  complex  that  it  is 
difficult  to  interpret  the  uptake  data,  but  possibly  there  is  some  correlation 
with  the  initial  and  delayed  responses  discussed  above.  Analysis  showed 
that  the  potential  changes  are  produced  when  certain  quantities  of  Hg++ 
are  bound  to  the  intestine;  i.e.,  when  different  concentrations  of  Hg++  are 
applied  to  the  intestine,  the  potential  changes  occur  at  different  rates  which 
are  related  to  the  uptake  rates.  Since  the  transports  and  potential  are  de- 


EFFECTS  ON  THE  KIDNEY  917 

pendent  on  glucose,  it  was  suggested  that  the  inhibitions  of  transintestinal 
transport  are  perhaps  all  the  result  of  interference  with  glucose  uptake  into 
the  cells.  This  does  not  explain  the  immediate  responses,  which  may  be 
due  to  the  direct  effect  of  Hg++  on  the  cell  membranes.  Indeed,  all  the 
changes  observed  may  arise  from  modifications  in  the  permeability  prop- 
erties of  the  lumenal  membranes. 


EFFECTS  ON   THE   KIDNEY 

The  clinical  effectiveness  of  the  mercurials  in  certain  edemas  has  stimulat- 
ed much  investigation  directed  at  discovering  the  mechanism  by  which  diu- 
resis is  produced.  It  is  now  clear  that  the  action  is  on  tubular  transport 
processes.  Since  these  transports  are  mainly  active  and  depend  on  tubular 
cell  metabolism,  as  well  as  on  certain  specific  carrier  and  enzyme  systems, 
we  shall  be  primarily  concerned  with  the  possible  effects  of  the  mercurials 
on  renal  metabolism  as  a  basis  for  their  diuretic  activity.  The  pharmacology 
textbooks  and  recent  reviews  (Beyer  and  Baer,  1960;  Farah  and  Miller, 
1962;  Kessler,  1960;  Mudge  and  Weiner,  1958;  Orloff  and  Berliner,  1961; 
Pitts,  1958,  1959)  cover  the  general  renal  actions  of  the  mercurials  and  also 
discuss  many  of  the  controversial  points.  We  shall  confine  ourselves  here  to 
a  summary  of  these  actions  and  then  proceed  to  the  mechanisms  which  may 
be  involved  in  the  alteration  of  the  transport  systems.  The  structures  of  the 
four  most  common  mercurial  diuretics  used  experimentally  are  shown  in 
their  ionic  forms.  The  preparations  provided  for  clinical  use  are  complexed 
with  different  ligands  (OH^,  Cl^,  thioacetate,  or  theophylline),  but  once 
introduced  into  the  body  or  experimental  media,  the  mercurials  usually 
establish  new  equilibria  with  the  available  ligands,  as  discussed  previously 


OCHjCOO"  H3C  CH3 

r,:^  OCH,  bOC^N (-CH,  OCH3 

L        ))— CONH-CH2— CH— CHj— Hg  V— CONH— CH— CH— CH2— Hg" 

Mersalyl  (Salyrgan)  Mercurin  (mercaptomerin,   Thiomerin) 

OCH3 
I 
HjN— CONH— CHj— CH— CH2 — Hg 

Chlormerodrin  (Neohydrin) 

OCH3 
I 
OOC— CH,CH,— CONH— CONH— CH,— CH— CHj— Hg 

Meralluride  (Mercuhydrin) 


918  7.  MERCURIALS 

(page  742).  Thus  mercurin  (Mercuzan)  is  the  un-ionized  acid,  the  Na+  salt 
is  Novurit,  the  complex  with  thioacetate  is  mercaptomerin  (Thiomerin), 
and  the  complex  with  theophylline  is  Mercurophylline  (Mercuzanthin),  but 
the  fundamental  active  structure  is  the  same  in  all  cases.  The  complexers 
alter  the  solubility,  local  actions  on  tissues,  and  rates  of  absorption,  but 
probably  do  not  significantly  affect  the  basic  effects  on  the  tubular  trans- 
port systems. 

Summary   of  the    General    Renal    Actions   of  the    Mercurials 

(A)  The  diuretic  action  is  entirely  renal.  Since  the  classic  transplantation 
and  unilateral  injection  studies  of  Govaerts  in  1928  and  Bartram  in  1932  it 
has  been  clear  that  any  extrarenal  actions  of  the  mercurials  are  unimportant 
for  diuresis.  Effects  on  tissues  other  than  the  kidneys  may  of  course  occur, 
especially  at  high  doses,  but  do  not  contribute  significantly  nor  are  they 
necessary  for  diuresis. 

(B)  The  primary  action  is  on  tubular  transport  rather  than  glomerular  fil- 
tration. Clearance  studies  have  demonstrated  that  glomerular  filtration  rates 
are  not  altered  much  or  at  all  during  marked  clinical  diuresis.  Analyses  of 
the  changes  in  composition  of  the  filtrate  as  it  passes  down  the  nephron 
show  that  the  increased  flow  of  urine  can  be  accounted  for  entirely  by  the 
depression  of  certain  tubular  transport  processes.  Mercurials  given  intramus- 
cularly or  orally  to  human  subjects  do  not  affect  renal  blood  flow  or  glom- 
erular filtration  (Brun  et  al.,  1947),  but  injected  intravenously  in  animals 
they  produce  definite  effects  which  may  possibly  modify  the  primary  diu- 
retic action.  Vasoconstriction  and  a  reduction  in  blood  flow  are  usually  ob- 
served. Jackson  (1926  b)  reported  that  intravenous  mersalyl  causes  a  rise 
in  blood  pressure  and  a  profound  constriction  of  the  kidney  (measured  on- 
cometrically),  while  Farah  (1952)  observed  the  renal  blood  flow  to  fall  50% 
from  mersalyl  at  10  mg/kg  intravenously  and  70%  from  4  mg/kg  intraarte- 
rially,  these  changes  occurring  before  the  onset  of  diuresis.  Kessler  et  al. 
(1957  b)  invariably  found  a  significant  decrease  in  glomerular  filtration  rate 
following  various  mercurial  diuretics  given  intravenously,  and  such  often 
persists  long  after  the  maximal  diuretic  effect  occurs.  Occasionally  a  trans- 
ient antidiuresis  is  noted  after  intravenous  diuretics  and  this  could  be  the 
result  of  renal  vasoconstriction  and  a  reduced  glomerular  filtration  rate 
(Vargas  and  Cafruny,  1959;  Cafruny  and  Palmer,  1961).  It  is  interesting  to 
note  that  the  nondiuretic  p-MB  likewise  produces  these  vascular  changes. 
Thus  the  effects  on  glomerular  filtration  when  observed  experimentally 
would  be  antidiuretic  and  might  reduce  the  over-all  diuretic  response  in- 
stead of  favoring  it. 

(C)  The  diuretic  action  is  mainly  due  to  an  inhibition  of  active  iVa+  resorp- 
tion in  the  proximal  tubules.  It  is  now  generally  agreed  that  the  mercurials 


EFFECTS  ON  THE  KIDNEY  919 

are  primarily  natriuretics,  and  that  the  resorption  of  Cl~  and  water  by  the 
proximal  tubules  follows  the  movement  of  Na+  for  electrostatic  and  osmotic 
reasons.  The  evidence  for  the  proximal  site  of  action  of  the  mercurials  will 
be  presented  later  (page  920).  The  fundamental  mechanism  of  action  must 
therefore  be  sought  in  the  modifications  of  active  Na+  transport  by  the 
mercurials. 

(D)  The  mercurials  can  also  act  elsewhere  on  the  nephron  to  modify  urine 
composition  and  flow  rate.  Sufficient  evidence  has  been  accumulated  to  show 
that  the  mercurials  can  exert  minor  effects  throughout  the  nephron  —  on 
the  loop  of  Henle,  the  distal  tubule,  and  the  collecting  duct  —  to  further 
alter  the  urine  volume,  and  that  transport  processes  and  exchange  reac- 
tions of  various  types  can  be  inhibited.  That  is,  the  primary  site  of  action 
may  be  on  active  Na+  transport  in  the  proximal  tubules,  but  this  is  by  no 
means  the  sole  site  of  action.  The  transports  of  a  variety  of  substances,  in- 
cluding K+,  Ca++,  urate,  p-aminohippurate,  amino  acids,  and  various  dyes, 
are  depressed  by  the  mercurials. 

(E)  Only  a  relatively  small  effect  on  Na+  resorption  need  be  exerted  to  pro- 
duce marked  diuresis.  Inasmuch  as  98-99%  of  the  filtered  Na+,  CI",  and 
water  is  resorbed,  it  is  evident  that  a  reduction  of  this  to  90-95%  would 
cause  up  to  a  10  fold  increase  in  excretion  rate.  Consequently  one  might 
predict  that  only  a  small  metabolic  disturbance  by  the  mercurials  would  be 
necessary  for  diuretic  action,  and  that  so  small  an  effect  might  be  difficult 
to  detect  under  the  usual  conditions.  Furthermore,  only  15-30%  of  the 
total  NaCl  resorption  can  be  inhibited  by  the  mercurials,  the  remainder 
presumably  being  mediated  through  mercurial-resistant  systems  (Pitts, 
1958). 

(F)  The  selective  action  of  the  mercurials  on  the  kidney  is  mainly  a  conse- 
quence of  the  accumulation  of  mercurial.  Mercurials  exert  demonstrable  ef- 
fects only  on  the  kidneys  over  a  dosage  range,  and  this  is  primarily  due  to 
the  relatively  high  concentrations  of  mercurial  reached  in  the  renal  tissue, 
whether  this  is  achieved  by  tubular  secretion  or  filtrate  resorption  (page 
928).  The  transport  systems  are  probably  no  more  sensitive  than  in  other 
tissues  to  the  mercurials  (at  least  there  seems  to  be  no  clear  evidence  for 
this).  However,  the  point  mentioned  in  the  previous  paragraph  that  only 
small  effects  on  renal  transport  need  be  exerted  may  be  a  factor  in  increas- 
ing the  apparent  sensitivity  of  the  kidney. 

(G)  The  reported  renal  responses  to  the  mercurials  are  quite  variable.  One 
cannot  fail  to  be  impressed  by  the  general  lack  of  agreement  on  certain 
basic  actions  of  the  mercurials  despite  the  great  amount  of  work  done  over 
many  years,  and  it  is  disturbing  that  almost  every  hypothesis  can  be  re- 
futed by  evidence  of  apparent  validity.  It  may  be  helpful  to  list  some  of  the 
reasons  for  these  disagreements.  (1)  Work  done  with  different  species  can 


920  7.  MERCURIALS 

frequently  not  be  compared.  For  example,  mercurials  in  diuretic  dosage 
inhibit  glucose  resorption  and  p-aminohippurate  secretion  in  man,  but  have 
no  effect  on  these  transports  in  the  dog;  also  p-MB  is  not  diuretic  in  the 
dog,  but  increases  urine  flow  in  the  rat  (Cafruny  and  Palmer,  1961).  (2) 
Animals  in  different  states  of  water  load,  ion  load,  or  pH  wiU  respond  dif- 
ferently to  the  mercurials.  (3)  The  use  of  theophylline-containing  mercurials 
has  often  confused  interpretation,  since  theophylline  itself  is  a  diuretic  act- 
ing by  a  mechanism  quite  different  than  the  mercurials.  Thus  Goldstein  et 
al.  (1961)  found  Mercuhydrin  (meralluride  complexed  with  theophylline) 
to  produce  two  phases  of  diuresis,  the  first  due  to  the  theophylline.  Certainly 
some  of  the  results  attributed  to  the  mercurials  have  had  their  origin  in  the 
theophylline  present,  and  for  this  reason  it  is  always  advisable  to  use  mer- 
curials complexed  with  inactive  substances  if  a  pure  mercurial  action  is  to 
be  investigated.  (4)  Much  of  the  work  on  distribution  of  the  mercurials  and 
their  actions  on  enzymes  in  the  kidney  has  been  done  with  toxic  or  lethal 
doses  or  concentrations.  If  a  mechanism  for  the  normal  diuretic  effect  is  to 
be  found,  one  must  use  mercurial  concentrations  which  do  not  deviate  ap- 
preciably from  those  producing  maximal  diuresis.  (5)  Different  routes  of  ad- 
ministration often  lead  to  different  results.  We  have  seen  that  intravenous 
injection  causes  changes  in  blood  flow  and  glomerular  filtration  not  seen 
with  the  usual  routes  of  administration. 

Sites  of  Action  in  the  Nephron 

Several  types  of  evidence  have  been  used  to  locate  the  major  sites  of 
mercurial  action  on  renal  transport  processes;  these  will  be  discussed  briefly, 
since  they  also  provide  interesting  information  on  the  mechanisms  involved. 

(A)  Inhibition  of  transport  processes  located  in  different  regions  of  the 
nephron.  The  mercurials  interfere  with  the  transport  of  a  variety  of  sub- 
stances by  the  proximal  segment  of  the  nephron.  This  includes  the  resorp- 
tion of  glucose,  amino  acids,  urate,  phosphate,  bicarbonate,  Na+,  K+,  and 
Ca++,  and  the  active  secretion  of  p-aminohippurate,  iodopyracet  (Diodrast), 
tetraethylammonium  ion,  phenol  red,  and  various  dyes.  Izar  (1909)  noted 
an  increase  in  urinary  urate  in  dogs  given  HgClg  intravenously,  and  Dale 
and  Sanderson  (1954)  demonstrated  that  urate  excretion  in  man  rises  rap- 
idly following  administration  of  mersalyl.  However,  if  oliguria  is  produced 
by  lethal  doses  of  HgCla,  urate  excretion  is  impaired  and  the  tissue  concen- 
tration will  rise  (Wells,  1916).  Mild  poisoning  by  mercurials  leads  to  an 
aminoaciduria  in  man  (Clarkson  and  Kench,  1956).  There  has  been  dis- 
agreement with  respect  to  the  effects  on  glucose  resorption,  and  perhaps  in 
man  there  is  little  reduction  at  diuretic  doses,  but  Vander  (1963)  has  shown 
a  very  definite  inhibition  in  the  dog,  A  Tm  being  around  — 100  during  max- 
imal diuresis.  Mercaptomerin  in  dogs  lowers  the  bicarbonate  threshold  of 
the  proximal  tubules  by  35%  without  significant  effect  on  the  distal  tubules 


EFFECTS  ON  THE  KIDNEY  921 

(Gardier  and  Woodbury,  1955).  Since  the  resorption  of  bicarbonate  is  about 
equal  in  the  proximal  and  distal  segments,  this  indicates  an  exclusively 
proximal  action.  Increased  excretion  of  Ca++  and  Mg++  in  both  man  and 
dog  treated  with  mercurials  has  been  reported,  and  it  is  likely  that  the  site 
is  proximal  (Wesson,  1962).  The  inhibition  of  the  secretion  of  p-aminohip- 
purate,  tetraethylammonium  ion,  and  phenol  red  has  been  shown  not  only 
in  intact  animals  but  with  low  concentrations  of  the  mercurials  in  isolated 
tubules  or  slices  (Forster  and  Taggart,  1950;  Farah  and  Rennick,  1956; 
Koishi,  1959  b).  Thus  0.01  mil/  Hg++  completely  blocks  phenol  red  trans- 
port in  the  flounder  tubule.  These  and  other  observations  all  point  definitely 
to  a  major  site  of  action  in  the  proximal  segment  of  the  nephron.  However, 
there  is  also  evidence  that  more  distal  transports  can  be  affected.  For  exam- 
ple, the  secretion  of  K+  and  the  H+  and  NH4+  exchanges  in  the  distal  seg- 
ment (Dale  and  Sanderson,  1954),  and  the  resorption  of  solute-free  water 
by  the  loop  of  Henle  (Lambie  and  Robson,  1960)  and  the  distal  segment 
(Goldstein  et  al.,  1961),  are  depressed  by  the  mercurials.  It  is  difficult  to 
compare  the  actions  on  proximal  and  distal  portions  of  the  nephron  be- 
cause of  the  different  magnitudes  of  the  transport  processes;  i.e.,  effects 
on  proximal  transport  would  be  much  more  marked  because  of  the  major 
role  of  this  segment  in  resorption. 

(B)  Disappearance  of  renal  SH  grovps.  Histochemical  determination  of 
the  free  SH  groups  in  different  regions  of  the  kidney  in  normal  and  mer- 
curial-treated animals  might  provide  some  information  on  the  site  of  ac- 
tion if  clear-cut  differences  are  observed.  Cafruny  et  al.  (1955  b)  determined 
the  free  protein  SH  groups  in  rat  kidney  sections  by  treatment  with  the  SH 
reagent  DDD  (2,2'-dihydroxy-6,6'-dinaphthyldisulfide),  coupling  of  the 
naphthol  moiety  with  the  azo  dye  Fast  Blue  RR,  and  photometric  analysis. 
Following  injection  of  a  large  dose  of  mersalyl  (20  mg/kg),  reduction  of  SH 
groups  was  observed  in  all  portions  of  the  nephron  except  the  proximal  and 
distal  convoluted  portions  (see  accompanying  tabulation).  Even  at  the 
markedly  nephrotoxic  dose  of  40  mg/kg  there  is  no  decrease  in  SH  groups 


Extinction 

values 

%  Change 

Control 

Mersalyl 

Proximal  tubules  (convoluted) 

0.619 

0.613 

-   1 

Proximal  tubules  (terminal) 

0.415 

0.227 

-45 

Brush  borders  (terminal) 

0.701 

0.545 

-22 

Loop  of  Henle  (descending) 

0.355 

0.238 

-33 

Loop  of  Henle  (ascending) 

0.373 

0.229 

-39 

Distal  tubules  (convoluted) 

0.592 

0.596 

+   1 

Collecting  duct  (medullary) 

0.279 

0.136 

-51 

922  7.    MERCURIALS 

in  the  convoluted  segments.  With  low  doses  (2.5  mg/kg),  disappearance  of 
SH  groups  was  observed  only  in  the  terminal  portion  of  the  proximal  tu- 
bules and  the  ascending  loops,  the  latter  being  the  most  sensitive  region 
of  the  nephron.  Time  studies  showed  that  the  terminal  proximal  tubules 
are  affected  first  and  up  to  1  hr  show  more  reduction  than  the  loops  of 
Henle.  Incubation  of  kidney  sections  with  20  mM  mersalyl  (Cafruny  et  al., 
1955  b)  or  saturated  HgClg  (Cafruny  et  al.,  1955  a)  produces  marked  non- 
specific reduction  in  free  SH  groups,  indicating  that  the  pattern  seen  in  the 
whole  animal  is  due  in  part  to  the  factors  involved  in  the  resorption  and 
secretion  of  the  mercurials.  Cafnmy  and  Farah  (1956)  later  used  dogs  so 
that  correlation  between  diuresis  and  SH  group  disappearance  might  be 
made.  Kidneys  were  removed  at  the  peak  of  diuresis  (around  90  min)  from 
10  mg/kg  of  mersalyl,  urine  flow  and  Na+  being  increased  5-  to  6-fold,  and 
the  changes  given  in  the  accompanying  tabulation  were  observed,  indicat- 


Extinction 

.  values 

0/ 

/o 

Change 

Control 

Mersalyl 

Proximal  tubules  (convoluted) 

0.537 

0.541 

+   1 

Proximal  tubules  (terminal) 

0.410 

0.267 

-35 

Loop  of  Henle  (ascending) 
Distal  tubules 

0.360 
0.495 

0.261 
0.510 

-28 
+  3 

Collecting  ducts 

0.242 

0.192 

-21 

ing  that  selective  reaction  with  SH  groups  in  certain  regions  of  the  kidney 
does  occur.  It  should  be  pointed  out  that  the  nature  of  these  SH  groups  is 
not  known;  they  may  be  on  enzymes,  carriers,  or  nonfunctional  proteins. 
Farah  and  Kruse  (1960)  used  seven  mercurials  at  4  mg  Hg/kg  in  rats 
and  found  moderate  reduction  of  the  protein  SH  groups  (around  20-30%) 
in  the  terminal  proximal  tubules,  the  loops  of  Henle,  and  the  collecting 
ducts,  and  it  was  concluded  that  maximal  diuresis  occurs  when  20%  of  the 
protein  SH  groups  of  the  proximal  tubule  are  reacted,  and  thus  that  no 
more  than  this  can  be  related  to  the  diuresis.  However,  there  is  no  correla- 
tion between  diuresis  and  decrease  in  the  SH  groups,  since  p-MB  and  MM, 
both  nondiuretic,  produce  similar  changes  in  these  groups.  HgClg  and  mer- 
salyl at  equimolar  doses  cause  comparable  decreases  in  renal  SH  groups  in 
rats,  and  this  was  noted  particularly  at  the  bases  of  the  proximal  tubular 
cells  (Gayer  and  Partowi,  1962).  Renal  S — S  groups  do  not  change  for  sev- 
eral hours  after  injections  of  HgClg  or  chlormerodrin,  but  from  6  to  24  hr 
there  is  a,n  increase  in  S — S  groups  at  the  expense  of  SH  groups  (Shore  and 
Shore,  1962).  This  may  be  related  to  the  potent  inhibition  of  protein  disul- 
fide reductase,  but  is  probably  not  correlated  with  diuresis  since  it  occurs 


EFFECTS  ON  THE  KIDNEY  923 

some  time  after  maximal  urine  flow.  Renal  damage  and  conversion  of  SH 
to  S — S  groups  could  be  related  in  some  as  yet  unexplained  way. 

These  results  all  demonstrate  that  mercurials  react  with  renal  SH  groups, 
and  that  some  selectivity  on  certain  regions  may  be  exerted,  but  do  not 
necessarily  have  any  bearing  on  the  site  of  transport  inhibition,  since  the 
SH  groups  involved  in  the  transport  (assuming  they  are)  may  be  only  a 
very  small  fraction  of  the  total  in  the  tissue;  indeed,  it  is  quite  possible  that 
only  1-2%  of  the  total  SH  groups  need  be  reacted  to  produce  maximal 
diuresis. 

(C)  Reduction  of  electrical  potentials  of  tubular  cells.  There  are  two  elec- 
trical potentials  of  the  proximal  tubular  cells  of  the  isolated  Necturus  neph- 
ron, a  transmembrane  potential  of  —72  mv  and  a  transtubular  potential 
of  —20  mv  (lumen  negative)  (Giebisch,  1958,  1960,  1961).  Chlormerodrin 
in  a  concentration  around  220  //g  Hg/g  tissue  reduces  both  potentials;  in 
the  perfused  nephrons  the  transmembrane  potential  is  decreased  62%  and 
the  transtubular  potential  63%.  Since  these  potentials  are  dependent  on 
active  ion  transports,  quite  possibly  they  relate  to  renal  function.  These 
results  show  that  mercurials  can  affect  the  proximal  tubules,  but  whether 
this  is  related  to  the  diuretic  effect  is  impossible  to  say. 

(D)  Pattern  of  accumulation  of  mercurials  in  the  kidney.  The  kidneys  of 
rats  poisoned  with  HgClj  (3  mg/kg  intraperitoneally)  were  examined  from 
5  min  to  48  hr  afterward  by  the  silver  sulfide  method,  and  mercury  was 
found  to  be  deposited  first  in  the  endothelial  cells  of  the  interstitial  capilla- 
ries, then  in  the  glomerular  tufts,  and  eventually  in  the  epithelium  of  the 
proximal  tubules,  beginning  apically  and  progressing  toward  the  bases  of 
the  cells  (Wockel  et  al.,  1961).  The  mercury  in  the  proximal  ceUs  is  partic- 
ularly associated  with  the  basally  situated  mitochondria.  It  was  conclud- 
ed that  Hg++  is  filtered  through  the  glomerulus  and  picked  up  by  the  tu- 
bular cells  during  resorption,  which  is  the  most  obvious  route  for  Hg++ 
and  one  which  explains  the  early  and  marked  effects  on  the  proximal  tu- 
bule. However,  it  has  recently  been  claimed  that  another  route  is  more 
important.  Brun  et  al.  (1947)  suggested  that  mersalyl  is  secreted  by  the  tu- 
bular cells,  and  that  this  accounts  for  the  high  concentration  of  mercury  in 
the  tubules  and  the  selective  effects  on  proximal  transport.  It  was  claimed 
by  Borghgraef  et  al.  (1956)  that  the  excretory  rate  of  chlormerodrin  is 
too  fast  for  glomerular  filtration,  especially  considering  that  a  large  frac- 
tion of  the  plasma  mercurial  is  bound  and  not  filtered,  and  that  tubular 
secretion  is  responsible  for  essentially  all  the  mercury  in  the  tubules.  This 
theory  has  also  been  proposed  by  Weiner  et  al.  (1956),  Kessler  et  al.  (1957 
a,  b),  and  Campbell  (1959).  Greif  (1960)  held  that  the  uptake  of  Hg^o^ 
by  Phascolosoma  nephridia  is  an  active  transport,  presumably  because  it  is 
inhibited  by  cyanide;  however,  no  inhibition  by  '2,4-dinitrophenol,  azide, 
or  iodoacetate  was  noted.  Despite  the  evidence  for  the  tubular  secretion  of 


924  7.    MERCURIALS 

mercurials,  I  am  not  convinced  that  it  is  more  important  than  resorption 
from  the  glomerular  filtrate.  If  1-5%  of  the  total  plasma  mercurial  is  in  a 
freely  diffusible  form,  this  fraction  will  certainly  be  filtered  readily  and 
concentrated  in  the  lumen;  renal  accumulation  occurs  with  many  drugs 
which  are  bound  to  the  plasma  proteins.  Second,  only  the  free  mercurial  is 
available  to  the  tubular  cells  from  the  peritubular  fluid,  and  it  is  difiicult 
to  envision  a  mechanism  by  which  the  cells  can  pick  up  or  actively  secrete 
protein-bound  mercurial.  The  tubular  cells  undoubtedly  accumulate  mer- 
curials from  the  plasma  as  do  other  tissues,  and  may  secrete  them  to  some 
extent,  but  the  rates  of  excretion  are  not  such  as  to  imply  secretion  as  the 
major  pathway.  It  is  also  strange  that  mecurials  of  so  many  different  struc- 
tures would  be  actively  secreted;  those  with  carboxylate  groups  might  be 
carried  by  the  transport  system  for  acids,  as  Campbell  (1959)  postulated 
for  mercaptomerin,  but  neither  probenecid  nor  p-aminohippurate  interferes 
with  the  excretion  of  mercurials  in  the  dog  (Kessler  et  al.,  1957  b).  In  any 
event,  there  is  little  likelihood  that  the  pattern  of  mercurial  distribution  in 
the  kidney  can  be  directly  correlated  with  the  site  of  action.  Weiner  et  al. 
(1959)  emphasized  that  there  is  no  obvious  relationship  between  diuresis 
and  the  total  amount  of  mercurial  in  the  kidney  or  its  parts,  and  stated, 
"Diuresis  is  a  consequence  of  the  action  on  a  specific  renal  receptor  by  a 
very  small  amount  of  mercury." 

(E)  Stop-flow  technique.  Serial  sampling  of  urine  following  ureteral  clamp- 
ing allows  an  analysis  of  the  composition  changes  throughout  the  nephron, 
and  such  studies  have  uniformly  pointed  to  a  proximal  rather  than  a  distal 
site  of  mercurial  action  (Kessler  et  al.,  1958;  Vander  et  al.,  1958).  This  ap- 
plies exclusively  to  the  site  of  inhibition  of  Na+  resorption  and  diuresis. 

(F)  Differential  damage  to  renal  cells.  It  has  been  thought  that  those  por- 
tions of  the  nephron  most  readily  damaged  by  toxic  doses  of  the  mercurials 
might  be  the  same  portions  primarily  affected  to  produce  diuresis.  Suzuoki 
(1912)  was  the  first  to  show  by  adequate  methods  that  mercurials  can  dam- 
age rather  selectively  the  more  terminal  portions  of  the  proximal  tubules. 
Edwards  (1942),  on  the  other  hand,  claimed  that  Hg++  exerts  selective 
damage  on  the  central  region  of  the  proximal  convoluted  tubule,  injury 
to  the  distal  convolution  being  rarely  seen.  The  loops  of  Henle  are  too  thin 
and  squamous  to  permit  satisfactory  examination.  Simonds  and  Hepler 
(1945)  confirmed  the  observations  of  Suzuoki  in  finding  selective  damage 
to  terminal  proximal  tubules.  More  recent  work  has  not  greatly  extended 
our  knowledge.  Relatively  little  damage  to  the  glomeruli  has  been  confirm- 
ed (Staemmler,  1956)  even  when  a  severe  nephrosis  is  produced,  although 
Schorcher  and  Loblich  (1960)  found  some  changes  in  the  glomerular  filtra- 
tion membrane  by  electron  microscopy.  Tubular  cells  show  apical  edema 
and  vacuolization,  these  occurring  primarily  in  the  proximal  segment  in 
rats  injected  with  meralluride  (Sanabria,  1963).  The  brush  border  may  show 


EFFECTS  ON  THE  KIDNEY  925 

a  separation  of  the  villi,  and  mitochondrial  disintegration  occurs.  With 
large  doses,  damage  may  be  observed  with  the  electron  microscope  within 
10  min.  The  maximal  diuretic  dose  of  mersalyl  (6  mg/kg)  in  the  rabbit 
produces  nuclear  pycnosis,  mitochondrial  changes,  and  vacuolization  in  the 
convoluted  tubule  (Dejung,  1963).  Again,  these  results  show  selective  ef- 
fects, but  may  be  the  result  of  differential  distribution  and  may  not  relate 
to  the  diuretic  site  of  action. 

The  evidence  taken  together  suggests  that  various  portions  of  the  nephron 
are  affected  in  one  way  or  another  by  the  mercurials,  and  what  portion  may 
be  involved  will  depend  on  the  particular  transport  considered.  The  diuretic 
action,  i.e.,  the  inhibition  of  Na+  resorption,  appears  to  be  limited  mainly 
to  the  proximal  tubule;  whether  this  is  primarily  in  the  convoluted  or  ter- 
minal segments  is  not  known.  Analyses  of  th6  filtrate  composition  through- 
out the  nephron  in  the  presence  of  mercurials  are  very  difficult  to  interpret, 
but  most  of  the  data  are  compatible  with  a  proximal  action  (Welt  et  al., 
1953).  More  indirect  evidence  will  be  provided  by  studies  on  enzyme  inhi- 
bition in  the  following  section,  but  it  is  clear  that  the  basic  mechanism  of 
mercurial  action  must  be  sought  in  the  Na+  transport  system  of  the  prox- 
imal tubules. 

Effects  on  Enzyme  Activity  in  the  Kidneys 

Much  of  the  work  has  unfortunately  been  on  succinate  dehydrogenase, 
presumably  because  the  activity  is  easily  measured,  but  this  enzyme  is 
probably  not  directly  involved  in  renal  transport  and,  in  fact,  is  much  less 
sensitive  to  the  mercurials  than  are  many  other  enzymes.  The  results,  how- 
ever, may  be  taken  as  a  rough  indication  that  the  mercurials  can  in  diuretic 
doses  inhibit  various  renal  enzymes,  and  to  some  extent  provide  evidence 
for  the  primary  site  of  action.  Handley  and  Lavik  (1950)  found  that  mer- 
alluride  injected  intravenously  at  a  dosage  of  8  mg  Hg/kg  in  dogs  and  rats 
reduces  the  succinate  dehydrogenase  activity  around  45%  in  the  kidney 
at  the  peak  of  diuresis,  whereas  no  inhibitions  were  observed  in  the  liver 
or  heart.  Fawaz  and  Fawaz  (1951),  on  the  other  hand,  could  detect  no 
changes  in  succinate  oxidation  by  renal  homogenates  from  rats  given  the 
diuretic  dose  (4  mg  Hg/kg)  of  mersalyl,  and  concluded  that  if  the  mercurial 
is  acting  on  an  SH  enzyme,  succinate  dehydrogenase  is  not  involved.  Mus- 
takallio  and  Telkka  (1953)  reported  that  high  doses  (10-60  mg  Hg/kg)  of 
mercurophylline  lead  to  loss  of  succinate  dehydrogenase  activity  in  the  distal 
tubules,  using  a  tetrazolium  staining  technique,  and  later  (Telkka  and  Mus- 
takallio,  1954)  found  some  inhibition  in  the  proximal  tubules  and  loops  of 
Henle,  their  conclusion  being  that  there  is  little  correlation  between  such  in- 
hibition and  transport  inhibition.  Somewhat  different  results  were  obtained 
by  Rennels  and  Ruskin  (1954),  who  found  marked  inhibition  of  succinate 
dehydrogenase  in  the  proximal  tubules  of  the  rat  following  40  mg  Hg/kg 
of  meralluride,  little  or  no  effect  being  exerted  on  the  distal  tubules  or  loops 


926  7.    MERCURIALS 

of  Henle.  However,  this  is  a  very  high  toxic  dose  and  the  maximal  inhibi- 
tion occurred  at  24-48  hr,  although  some  effect  could  be  observed  at  3  hr. 
In  doses  under  10  mg  Hg/kg,  no  inhibition  could  be  demonstrated,  so  that 
again  there  is  no  reason  to  relate  the  inhibition  to  diuresis.  Wachstein  and 
Meisel  (1954)  also  observed  succinate  dehydrogenase  inhibition  in  the  prox- 
imal tubules  of  the  rat  following  large  nephrotoxic  doses  of  meralluride. 
Bahn  and  Longley  (1956)  confirmed  these  results  in  that  diuretic  doses  of 
meralluride  (4  mg  Hg/kg)  produce  insignificant  inhibition  of  succinate  de- 
hydrogenase while  toxic  doses  (12  mg  Hg/kg)  inhibit  moderately,  especially 
in  the  terminal  proximal  tubules.  Finally,  Bickers  et  al.  (1960)  found  that 
meralluride  at  3-4  mg  Hg/kg  does  not  inhibit  succinate  dehydrogenase, 
whereas  5  mg  Hg/kg  does  to  some  extent  after  10  hr.  Later  changes  in  en- 
zyme activity,  at  a  period  of  tubular  damage,  may  be  the  result  of  secondary 
changes  and  necrosis,  so  have  little  bearing  on  the  mechanism  of  the  diuretic 
action. 

a-Ketoglutarate  oxidase  is  more  sensitive  to  mercurials  than  is  succinate 
dehydrogenase,  and  is  inhibited  64%  at  3-4  hr  and  91%  at  6  hr  in  rats 
given  HgClg  at  3  mg  Hg/kg  (Shore  and  Shore,  1960).  In  animals  fed  sucrose, 
the  inhibitions  are  less,  and  sucrose  also  protects  somewhat  against  the 
nephrotoxic  effects  of  Hg++,  the  mechanism  being  unknown.  But  again  it 
is  impossible  to  correlate  this  action  with  the  diuretic  effect,  since  mersalyl 
in  diuretic  dose  (4  mg  Hg/kg)  does  not  alter  the  renal  concentration  of 
a-ketoglutarate,  although  toxic  doses  (10  mg  Hg/kg)  produce  an  early  and 
prolonged  rise  in  a-ketoglutarate  (Dziirik  and  Krajci-Lazary,  1962).  No 
changes  in  sorbitol  dehydrogenase  activity  in  the  kidneys  of  rats  given 
nephrotoxic  doses  of  mercurin  are  detectable,  and  it  would  probably  not 
be  very  significant  if  they  were  (Pietschmann  et  al.,  1962).  Moderate  inhibi- 
tions of  NAD  and  NADP  diaphorases  are  found  after  toxic  doses  of  the 
mercurials  (Bickers  et  al.,  1960;  Wachstein  and  Meisel,  1957),  but  in  all 
cases  where  enzyme  inhibition  is  seen,  there  is  histologically  demonstrable 
damage  to  the  tubules.  Protein  disulfide  reductase  is  apparently  rather 
potently  inhibited  in  the  kidney  during  mercurial  action,  but  this  is  un- 
doubtedly unrelated  to  the  diuretic  effect  (Shore  and  Shore,  1962). 

Various  phosphatases  have  been  the  subject  of  investigation  although 
there  is  little  reason  to  expect  a  relation  to  ion  transport.  Nephrotoxic 
doses  of  HgClg  inhibit  renal  phosphatase  slightly  (10-20%)  without  affect- 
ing serum  phosphatase,  but  subtoxic  doses  somewhat  increase  the  phospha- 
tase activity  (Hepler  et  al.,  1945).  There  is  no  correlation  between  glucosuria 
and  phosphatase  inhibition  (Hepler  and  Simonds,  1946).  Very  high  doses 
(175-260  mg/kg)  of  HgClg  decrease  tubular  phosphatase,  although  not  in 
capillaries  or  glomeruli,  but  low  doses  (6-10  mg/kg),  which  are  lethal  in  3 
weeks,  do  not  alter  the  phosphatase  (Sachs  and  Dulskas,  1956).  Diuretic 
doses  of  various  mercurials  do  not  inhibit  alkaline  phosphatase,  |5-glycero- 


EFFECTS  ON  THE  KIDNEY  927 

phosphatase,  or  glucose-6-phosphatase  (Shore  et  al.,  1959;  Bickers  et  al., 
1960),  but  toxic  doses  lead  to  a  diffuse  distribution  of  various  phosphatass 
in  the  tubular  cells,  and  some  inactivation  (Wachstein  and  Meisel,  1957). 
Very  little  work  has  been  done  on  ATPase,  but  there  is  shght  evidence  that 
it  can  be  inhibited  in  the  kidney  by  both  mersalyl  (DeGroot  et  al.,  1955) 
and  p-MB  (Padykula  and  Herman,  1955).  It  is  probably  fair  to  say  that 
none  of  these  investigations  of  phosphatases  has  established  a  relation  to 
their  role  in  transport  or  mercurial  diuresis. 

Effects   on    Renal    Metabolism 

Mercurials  can  certainly  depress  the  respiration  of  kidney  at  fiufficiently 
high  doses,  and  this  may  possibly  be  a  factor  in  the  renal  damage  produced, 
but  there  is  little  evidence  that  nontoxic  diuretic  doses  of  the  mercurials 
reduce  Og  consumption.  The  early  work  of  Gremels  (1929)  indeed  showed 
that,  in  the  heart-lung-kidney  preparation,  mersalyl  actuaUy  increases  renal 
respiration  during  diuresis.  Cohen  (1953  a,b)  showed  particularly  well  the 
difference  between  diuretic  and  toxic  doses.  Mercaptomerin  at  a  dosage  of 
10.7  mg  Hg/kg  in  rats  produces  a  marked  diuresis.  After  1  hr  the  animals 
were  sacrificed  and  a  mitochondrial  suspension  prepared  from  the  kidneys; 
no  change  in  the  Og  uptake  was  noted,  and  the  P  :  0  ratio  may  actually 
have  been  increased.  However,  with  a  toxic  dose  of  26.7  mg  Hg/kg,  the  O2 
uptake  was  depressed  35%  and  the  P:0  ratio  dropped  from  0.55  to  0.092. 
Very  similar  effects  were  observed  on  kidney  slices,  only  toxic  doses  reduc- 
ing the  respiration.  Even  toxic  doses  do  not  affect  the  O2  uptake  or  P:0 
ratio  of  liver  mitochondria  obtained  from  poisoned  animals.  That  mercurials 
exert  very  little  effect  on  the  cycle  when  given  in  diuretic  doses  was  demon- 
strated by  Fawaz  and  Fawaz  (1954).  Rats  were  poisoned  with  fluoroacetate, 
and  citrate  accumulation  was  determined  in  the  heart  and  kidney  of  control 
animals  and  those  given  mersalyl  at  4-5  mg  Hg/kg;  no  significant  change 
in  citrate  formed  was  noted,  so  that  the  operation  of  the  cycle  would  seem 
to  be  unaltered.  Although  these  results  indicate  no  appreciable  interference 
with  coenzyme  A,  Leuschner  et  al.  (1957)  claimed  that  the  coenzyme  A 
induced  acetylation  of  sulfanilamide  is  reduced  by  HgCla,  mersalyl,  and 
mercaptomerin  in  the  same  order  of  potency  as  for  diuresis.  Toxic  doses 
reduce  the  coenzyme  A  reaction,  but  it  is  not  certain  if  diuretic  doses  are 
able  to  do  this.  The  mercurial  diuresis  is  less  in  coenzyme  A-deficient  rats, 
but  the  significance  of  this  is  unknown.  Eesults  on  oxidative  phosphoryla- 
tion are  contradictory.  Greif  and  Jacobs  (1958)  found  that  even  large  doses 
of  chlormerodrin  (up  to  20  mg  Hg/kg,  which  is  20  times  the  diuretic  dose) 
do  not  alter  the  P :  0  ratio  of  kidney  mitochondria  with  glutamate  as  the 
substrate,  but  Shore  and  Shore  (1960)  reported  a  marked  fall  in  P:0  ratio 
with  a-ketoglutarate  as  the  substrate  following  toxic  doses  (3  mg  Hg/kg) 
of  HgCl^. 


928  7.    MERCURIALS 

The  results  obtained  on  kidney  slices  are  generally  in  agreement  that, 
although  respiration  may  be  reduced  by  reasonably  high  concentrations 
of  mercurials  (0.25-1  mM),  the  inhibition  of  various  transports  is  much 
greater.  Thus  Cross  and  Taggart  (1950)  found  that  1  mM  Hg++  depresses  Og 
uptake  of  rabbit  kidney  slices  35%  while  inhibiting  p-aminohippurate  ac- 
cumulation 89%,  and  Mudge  (1951)  showed  that  respiration  is  scarcely  af- 
fected by  Hg++  at  concentrations  markedly  altering  Na+  and  K+  transport. 
Mendelsohn  (1955)  confirmed  that  Hg++  can  reduce  p-aminohippurate  ac- 
cumulation as  much  as  70%  without  affecting  respiration  in  kidney  slices. 
Kobinson  (1956)  believed  that  the  inhibition  of  respiration  by  mercapto- 
merin  might  be  the  basis  for  the  swelling  of  rat  kidney  slices  and  the  inter- 
ference with  water  transport,  but  there  was  no  direct  evidence  for  a  relation- 
ship. Maizels  and  Remington  (1958)  also  observed  moderate  respiratory 
inhibition  with  mercaptomerin  and  meralluride,  but  did  not  feel  that  this 
was  the  chief  factor  in  the  increase  in  water  and  Na+  of  the  slices.  Further- 
more, the  lowest  concentration  of  the  mercurials  which  exerts  an  effect  in 
vitro  is  much  greater  than  the  maximal  tolerated  plasma  concentration  in 
rats  in  vivo,  so  it  is  doubtful  if  these  inhibitions  of  respiration  are  relevant 
to  the  diuretic  action. 

Summarizing  all  the  data  obtained  on  enzyme  and  metabolic  inhibition 
in  the  kidney,  it  is  disappointing  that  no  system  particularly  sensitive  to 
the  mercurials  has  been  found,  and  that  no  correlation  between  inhition 
and  transport  processes  has  been  demonstrated.  If  the  basis  of  mercurial 
diuresis  is  metabolic,  no  clear  evidence  for  this  has  yet  been  provided. 

Accumulation  and  Excretion  of  Mercurials  by  the  Kidneys 

All  mercurials  are  rather  slowly  accumulated  by  the  kidneys  and  the  high 
levels  are  sustained  for  periods  of  several  days.  The  kidney/plasma  ratio  is 
maximally  around  1000  in  the  rat  and  dog  when  chlormerodrin  is  given 
(Borghgraef  and  Pitts,  1956;  Giebisch  and  Dorman,  1958),  but  these  ratios 
are  reached  only  after  many  hours,  and  are  in  part  due  to  the  retention  by 
the  kidney  with  falling  plasma  levels.  The  correlation  between  distribution 
and  excretion  of  a  mercurial  and  the  diuresis  is  well  shown  in  the  results  of 
Borghgraef  et  al.  (1956)  (Fig.  7-44).  The  loss  of  mercurial  from  the  plasma 
is  divided  into  three  components:  that  excreted  in  the  urine,  that  entering 
the  various  tissues,  and  that  accumulated  by  the  kidneys.  Some  mercurials 
are  excreted  fairly  rapidly  and  others  slowly;  meralluride  administered  in- 
tramuscularly in  man  is  50%  excreted  in  3  hr  (Burch  et  al.,  1950)  but 
chlormerodrin  given  by  the  same  route  to  rats  is  only  21%  excreted  after 
24  hr,  67%  remaining  in  the  kidneys  (Borghgraef  and  Pitts,  1956).  We  have 
discussed  the  theories  of  the  excretory  mechanisms  (page  923)  and  the 
possible  role  of  tubular  secretion.  The  mercurials  are  not  excreted  entirely 
in  the  form  administered.  Some  may  be  split  into  inorganic  Hg++  but  most 


EFFECTS  ON  THE  KIDNEY 


929 


is  excreted  as  a  complex  with  cysteine,  or  other  thiols  (Weiner  and  Miiller, 
1955).  The  origin  of  this  cysteine  complex  is  not  known,  and  it  may  be  in 
the  kidney  or  the  blood.  The  various  mercurials  are  distributed  differently 
throughout  the  tissues,  as  one  might  expect  from  the  different  properties  of 
their  molecules,  and  this  must  play  some  role  in  the  effects  they  produce, 


lOOfc   -w 


Fig.   7-44.   Distribution  of  Hg^"^  after  intravenous   injection  of 

chlormerodrin   in    dogs   at   1   mg   Hg/kg.   The    curve  shows   the 

change  in  urine  flow  (diuresis)  estimated  from  the  figures  given. 

(From  Borghgraef  et  al.,   1956.) 


not  only  on  the  kidney  but  on  other  tissues  in  higher  doses.  The  most 
thorough  study  has  been  made  by  Kessler  et  al.  (1957  a)  and  some  of  their 
results  are  summarized  in  Table  7-19.  Hg++  behaves  quite  differently  than 
the  organic  mercurials;  it  does  not  enter  the  kidney  rapidly  but  eventually 
reaches  very  high  levels  after  several  hours.  It  may  be  noted  that  the  dis- 
tribution of  p-MB  is  not  markedly  different  from  the  diuretic  mercurials. 
Little  is  known  about  the  cellular  fractions  of  the  kidney  which  accumulate 
the  mercurials,  but  it  is  somewhat  surprising  that  Greif  et  al.  (1956)  found 
that  by  far  the  most  mercury  after  injection  of  chlormerodrin  to  rats,  fol- 
lowed by  fractionation  of  kidney  homogenates  in  sucrose  solutions,  is  in 
the  soluble  fraction,  about  one  third  the  amount  in  granules,  and  much 
less  in  the  nuclei.  Although  all  of  these  distribution  studies  are  important 
in  understanding  many  facets  of  mercurial  action,  they  do  not  appreciably 
contribute  to  our  knowledge  of  where  or  how  the  mercurials  produce  dis- 
turbances in  the  renal  function. 


930 


7.    MERCURIALS 


Table  7-19 

Concentration  of  Mercury  in  Tissues  of  the  Dog  After  Intravenous 

Administration  of  Mercurials  " 


Tissue 

concentration 

(//g/g  wet 

weight) 

Tissue 

Chlormerodrin 

Meralluride 

Mersalyl 

HgCl, 

p-MB 

Renal  cortex 

155 

19 

7.7 

113 

36 

Renal  medulla'' 

125 

13 

4.0 

79 

— 

Renal  papilla 

4.5 

1.4 

0.4 

2.3 

— 

Liver 

3.9 

0.8 

0.8 

6.6 

2.8 

Heart 

0.3 

0.2 

0.4 

0.5 

0.7 

Spleen 

1.4 

0.6 

1.1 

74 

2.1 

Lung 

1.2 

0.9 

0.9 

2.0 

2.8 

Diaphragm 

0.2 

0.2 

0.1 

0.3 

0.8 

Intestine 

0.7 

0.5 

0.7 

1.1 

2.5 

Skin 

0.4 

1.0 

0.7 

0.8 

1.0 

Plasma 

1.1 

1.3 

1.2 

2.9 

— 

"  At  2  mg  Hg/kg  and  sacrifice  of  the  animals  at  160  niin.  (From  Kessler  et  nl., 
1957  a;  p-MB  data  from  Kessler  et  ah,   1957  b.) 

''  Only  outermost  sections  of  medulla  considered  here. 


Active  Form  of  the  Mercurials  and  Relation  of  Action  to  Structure 

The  concept  that  the  organic  mercurials  in  order  to  be  active  diuretics 
must  dissociate  into  inorganic  mercury  is  an  old  one  and  has  been  revived 
recently  to  explain  some  of  the  differences  between  mercurials  and  the  ef- 
fects of  pH  on  the  activity.  Most  diuretic  mercurials  have  the  structure: 

OCH3 

I 
R— C  H2— CH— CH2— Hg+ 

The  methoxy  group  arises  because  these  mercurials  are  synthesized  by  the 
oxymercuration  of  alkenes  in  methanol;  the  nature  of  this  group  is  not  par- 
ticularly important  for  the  activity.  It  is  possible  that  the  reverse  of  this 
reaction 

OCH3 

1 
R— CH2— CH— CH2— Hg+  +  H+  -►  R— CH2— CH  =  CHj  +  CH3OH  +  Hg++ 

might  occur  under  physiological  conditions,  as  suggested  by  Hughes  (1957). 


EFFECTS  ON  THE  KIDNEY 


931 


Such  a  splitting  would  occur  more  rapidly  the  lower  the  pH.  Certainly  this 
release  of  Hg++  is  not  generally  responsible  for  the  actions  and  toxicity  of 
the  organic  mercurials,  and  most  of  the  mercurials,  such  as  p-MB,  PM, 
and  MM,  are  quite  stable.  Hepp  (1887)  emphasized  long  ago  that  alkyl 
mercurials  do  not  release  Hg++  in  the  body  and  exert  a  toxic  action  much 
different  than  Hg++.  However,  the  diuretic  mercurials  present  a  different 
situation  and  the  theory  of  Hg++  release  must  be  given  serious  consideration. 
The  possibility  that  the  above  reaction  might  be  catalyzed  or  accelerated  by 
thiols  through  the  formation  of  mercaptides  was  presented  by  Benesch  and 
Benesch  (1952)  as  a  result  of  their  polarographic  investigations  of  the  reac- 
tion between  mersalyl  and  dimercaprol.  In  this  scheme,  free  Hg++  may  not 
be  produced  directly;  instead,  a  cyclic  mercaptide  is  formed,  which  could 

OCH, 

I 
-S-Hg— CHg— CH— R 

hS-Hg— CH2—  CH— R 
OCH, 


-s. 


Hg 


hS 


/ 


CH,=  CH  — R 


CH,OH 


OCH3 

I  + 

R— CH— CHj— Hg 


conceivably  be  the  inhibiting  complex  in  renal  transport,  although  it  is  also 
possible  that  monothiols  can  act  similarly: 


OCH3 
— S— Hg— CH2— CH— R  +  H^ 


-S— Hg+  +  CH2=CH— R  +  CH3OH 


since  Mudge  and  Weiner  (1958)  showed  that  cysteine  increases  the  split- 
ting of  mersalyl  in  acid  medium.  Other  acid-stimulated  types  of  splitting 
would  be  the  simple  reactions: 


\\     // 


Hg 


v\     // 


Hg 


R— CH,— Hg      +     H 


R-CH,    -    Hg* 


but  these  usually  occur  fairly  slowly,  especially  for  the  alkyl  mercurials. 


932  7.    MERCURIALS 

The  hypothesis  that  organic  mercurial  diuretics  to  be  active  must  release 
inorganic  Hg++  in  the  kidney  was  proposed  by  Mudge  and  Weiner  (1958) 
and  the  evidence  was  presented  by  Weiner  et  al.  (1962).  This  had  been 
suggested  occasionally  ever  since  the  first  use  of  the  organic  mercurials  but 
very  little  evidence  either  for  or  against  was  reported,  and  the  idea  generally 
was  not  taken  seriously  because  all  the  other  metal  compounds  used  clini- 
cally had  been  shown  to  act  directly  without  splitting  off  the  metal  ion.  The 
evidence  now  accumulated  impels  one  to  consider  this  possibility.  If  such 
splitting  occurs,  it  is  important  not  only  for  the  diuretic  action  but  for  many 
other  effects  of  the  mercurials,  even  in  vitro.  The  evidence  is  mainly  of  two 
sorts:  (1)  a  correlation  between  acid  lability  of  organic  mercurials  and  their 
diuretic  activity,  and  (2)  the  potentiation  of  diuretic  activity  by  the  acid- 
ifying NH^Cl. 

Mudge  and  Weiner  (1958)  pointed  out  that  mersalyl  does  not  split*  in 
acid  medium  over  3  hr,  but  in  the  presence  of  cysteine  the  half-time  for 
splitting  is  105  min  and  with  dimercaprol  5  min.  The  acid  lability  of  32 
mercurials  was  tested  by  Weiner  et  al.  (1962)  by  incubating  the  mercurial 
at  1  WlM  with  cysteine  at  2  mM  in  an  Og-free  medium  at  pH  4  and  37° 
for  3  hr.  The  diuretic  activity  was  expressed  as  A  CI  (//moles/min/kg).  All 
of  the  22  mercurials  which  are  diuretic  are  acid-labile,  while  of  the  9  non- 
diuretic  mercurials  6  are  stable  and  3  labile  (1  mercurial  is  indeterminate 
in  diuretic  activity).  There  is  thus  a  reasonably  good  correlation  between 
lability  and  diuretic  activity.  The  3  labile  nondiuretic  mercurials  are  all 
of  the  ether  series  with  structures  of  the  type  R — CHg — 0 — CH2CH2 — Hg+, 
and  possibly  their  distribution  is  such  that  Hg++  is  not  released  in  the 
proper  region.  The  pH  dependence  of  the  splitting,  according  to  the  reac- 
tions of  Benesch  and  Benesch  (1952),  indicate  the  rate  of  splitting  to  be 
approximately  one  thousandth  as  fast  at  pH  7  as  at  pH  4.  If  one  assumes 
that  X  =  XqC'^',  where  X  is  the  amount  of  organic  mercurial,  it  may  be 
calculated  that  at  pH  7  around  0.14%  splitting  would  occur  in  3  hr,  since 
the  mean  splitting  of  the  labile  mercurials  is  about  75%.  Since  maximal 
diuresis  occurs  in  1-2  hr,  approximately  0.1%  would  be  split  in  this  time. 
Now,  this  calculation  is  not  very  valid  because  one  does  not  know  the  con- 
ditions for  splitting  in  the  kidney;  e.g.,  dimercaptides  may  be  formed  there 
and  split  more  rapidly  than  the  cysteine  complexes.  Diuretic  activity  was 
examined  by  injecting  the  mercurials  with  a  10-fold  excess  of  cysteine,  but 
presumably  in  the  kidney  there  would  be  a  transfer  of  the  mercurial  from 
cysteine  to  other  thiols.  Thus  it  is  difficult  to  obtain  an  idea  of  the  amount 
of  inorganic  mercury  which  is  released  in  the  kidney.  If  much  splitting  oc- 
curs, one  might  expect  to  find  Hg++  excreted  in  the  urine  in  some  form. 

*  The  term  "split"  will  be  used  to  designate  the  dissociation  of  the  mercurial  into 
inorganic  mercury  so  that  there  will  be  no  confusion  with  the  term  "dissociation" 
which  is  used  to  indicate  the  reaction  R — Hg — X  ±^  R — Hg+  +  X~. 


EFFECTS  ON  THE  KIDNEY  933 

Moyer  et  al.  (1957)  and  Handley  and  Seibert  (1956)  could  detect  no  inor- 
ganic mercury  after  administration  of  meralluride,  but  Weiner  et  al.  (1962) 
pointed  out  that  Hg++  would  be  excreted  mainly  as  the  cysteine  complex 
and  this  would  be  included  with  the  organic  mercurial  in  their  chromato- 
graphic fractionations.  It  is  also  possible  that  very  tight  binding  of  the  ac- 
tive Hg++  in  the  kidney  would  occur,  so  that  the  excretion  would  be  slow. 
Weiner  et  al.  (1962)  could  detect  Hg-cysteine  in  the  urine  following  injec- 
tion of  3  particularly  labile  mercurials,  but  of  course  this  is  not  valid  evi- 
dence that  the  Hg++  is  the  active  form. 

Clinical  diuretic  refractoriness  to  the  mercurials  has  been  known  for  years 
and  it  is  often  possible  to  restore  the  diuretic  response  by  giving  NH4CI. 
The  potentiating  action  of  NH4CI  has  been  the  subject  of  much  work  and 
speculation,  but  the  mechanism  is  still  unknown.  One  hypothesis  is  that 
the  urinary  acidification  is  the  major  factor.  Weiner  et  al.  (1962)  assume 
that  this  acidification  increases  the  splitting  of  the  labile  mercurials.  Ad- 
ministration of  NH4CI  drops  the  pH  of  the  urine  below  5  and  an  optimal 
effect  is  usually  seen  around  4.5;  thus  the  splitting  of  the  mercurial  in  the 
urine  will  be  significantly  accelerated.  However,  the  Hg++  in  the  urine  wiU 
presumably  be  complexed  with  cysteine  or  other  simple  thiols  and  the  rate 
of  splitting  will  not  be  very  great.  It  has  also  generally  been  assumed  that 
the  splitting  occurs  in  the  tubular  cells.  Although  the  intracellular  pH  un- 
doubtedly falls  after  NH4CI,  the  decrease  is  certainly  not  as  great  as  in 
the  urine.  Pending  determinations  of  intracellular  changes,  one  cannot  esti- 
mate the  effect  this  would  have  on  mercurial  splitting.  The  mercury  con- 
tents of  the  proximal  tubules  in  the  dog  were  determined  histochemically 
by  Cafruny  (1962),  using  di-/?-naphthylthiocarbazone,  and  acidosis  was 
shown  to  increase  the  levels  for  chlormerodrin  and  Hg++,  although  a  de- 
crease occurs  with  p-MB.  He  felt  that  acidosis  either  increases  the  available 
receptors  for  mercurials  or  somehow  alters  the  affinity  of  the  mercurial  for 
the  receptors.  Change  in  the  acid-base  balance  does  not  alter  the  excretion 
of  the  mercurials,  so  it  is  presumably  not  a  matter  of  the  tubular  concen- 
tration of  mercurial.  If  the  fall  in  pH  is  responsible  for  increased  splitting 
of  the  mercurials  and  thus  a  greater  diuretic  effect,  NH4CI  administration 
or  other  acid-base  changes  should  not  affect  the  diuresis  produced  by  Hg++ 
complexes.  Mudge  and  Weiner  (1958)  and  Levy  et  al.  (1958)  reported  that 
the  action  of  meralluride  is  altered  more  than  Hg-cysteine  by  variations  in 
the  urinary  pH.  However,  Hg-cysteine  diuresis  is  increased  2.4-fold  in  go- 
ing from  alkalosis  to  acidosis,  so  that  the  results  are  not  as  clear-cut  as  one 
might  wish.  If  acidosis  is  responsible  for  greater  mercurial  action,  one 
would  also  expect  that  any  type  of  acidosis  would  be  effective.  However, 
Kessler  (1960)  points  out  that  inhalation  of  12%  CO.,  actually  decreases 
mercurial  diuresis,  although  not  as  much  as  alkalosis  produced  by  HCOg" 
infusion.  But  inhalation  of  12%  CO2,  although  it  produces  an  acidification 


934 


7.    MERCURIALS 


of  the  plasma  (pH  7.4  to  7.14),  does  not  alter  urinary  pH  significantly,  as 
does  administration  of  NH4CI;  no  one  knows  what  happens  to  intracellular 
pH  in  the  tubules.  It  may  also  be  pointed  out  that  alkalinization  of  the 
urine  with  acetazolamide  or  K+  does  not  alter  the  diuretic  response  to  mer- 
curials (Pitts,  1958). 

It  is  not  immediately  apparent  why  Hg++  must  be  formed  from  organic 
mercurials  to  inhibit  renal  transport,  since  in  most  cases  the  organic  mer- 
curials react  readily  with  SH  groups  which  may  be  involved.  If  mercaptides 
or  dimercaptides  participate  in  the  splitting  of  the  mercurials,  the  Hg++ 
formed  must  dissociate  from  these  SH  groups  and  attach  to  others,  because 
the  cell  component  originally  binding  the  mercurial  must  be  blocked  and 
there  would  be  no  necessity  for  splitting.  If  Hg++  is  necessary  for  diuresis, 
it  must  be  that  either  (1)  a  cyclic  mercaptide  is  required,  or  (2)  the  impor- 
tant SH  groups  are  not  sterically  accessible  to  the  larger  organic  mercur- 
ials. Weiner  et  al.  (1962)  assume  that  the  specific  receptor  for  the  diuretic 
action  contains  two  groups,  one  being  an  SH  group  and  the  other  either  an 
SH  group  or  some  other  ligand  complexing  with  Hg++  (e.g.,  an  amino 
group).  The  complete  scheme  as  outlined  by  Weiner  et  al.  (1962)  is  shown 


NONSPECIFIC 
PROTEIN 


R-Hg-S-PROTEIN 


4= 


R-Hg-S-CYST  ■ 


CYST-S-Hg-S-CYSTj 


R-Hg-S-CYST 
iCYST-S-Hg-S-CYST 


SPECIFIC 
RECEPTOR 


Fig.   7-45.    Scheme  of   mercurial   re- 
actions in  the  kidney.  (From  Weiner 
et  al.,  1962.) 


in  Fig.  7-45.  It  is  strange  that  this  transport  component  would  not  be  inhib- 
ited by  organic  mercurials  bound  to  one  SH  group,  or  that  certain  potent 
SH  reagents,  such  as  MM  or  p-MB,  would  not  inactivate  it.  The  nondiuretic 
p-MB  prevents  and  reverses  the  diuretic  effects  of  the  mercurials,  while 
MM  does  not  do  this  (Miller  and  Farah,  1962  a).  A  competition  between 
p-MB  and  Hg++  for  the  receptor  SH  groups  was  suggested.  Miller  and  Farah 


EFFECTS  ON  THE  KIDNEY  935 

(1962  b)  also  postulate  a  two-group  receptor,  one  group  being  SH;  mer- 
curials which  are  diuretics  attach  to  the  SH  group,  split,  and  the  resulting 
Hg++  makes  a  two-point  attachment.  The  block  by  p-MB  is  due  to  its 
binding  to  the  SH  group;  being  stable  it  does  not  split.  If  this  is  so,  p-MB 
might  be  expected  to  displace  mercurials  in  the  kidney,  and  this  was  dem- 
onstrated using  Hg^^^-labeled  chlormerodrin.  The  decrease  in  radioactivity 
of  the  kidneys  parallels  the  antagonism  of  the  diuresis  by  p-MB. 

Another  hypothesis  for  diuretic  mechanism  was  outlined  by  Kessler  et 
al.  (1957  b),  who  assumed  that  the  organic  mercurials  act  as  intact  molecules 
by  a  two-point  attachment  to  a  receptor.  The  basic  structure  for  diuretic 
activity  was  given  as  R — C — C — C — Hg+,  i.e.,  a  hydrophilic  group  sepa- 
rated from  the  Hg  by  three  carbon  atoms,  the  R  group  interacting  in  some 
manner  with  a  group  spaced  appropriately  in  relation  to  an  SH  group.  This 
hypothesis  in  its  simple  form  has  had  to  be  abandoned  in  the  light  of  fur- 
ther work  showing  that  various  mercurials  not  conforming  to  this  structure 
are  diuretic,  e.g.,  some  of  the  aryl  mercurials  (Weiner  et  al.,  1962).  How- 
ever, the  idea  that  there  is  some  relationship  between  structure  and  diuretic 
action  should  not  be  given  up,  inasmuch  as  the  situation  may  be  more  com- 
plex than  originally  assumed.  If  one  considers  the  three  simple  alkyl  mer- 
curials (see  accompanying  tabulation),  one  sees  that  diuretic  activity  is 

Alky]  mercurial  Diuretic  activity  Lability 


CH3— CH2— Hg+  -  - 

HO— CH2— CH2— Hg+  +  + 

HO— CH2— CH2— CHj— Hg+  -  - 


correlated  with  acid  lability,  and  that  HO— CH2CH2CH2— Hg+,  which 
should  be  diuretic  in  the  scheme  of  Kessler  et  al.,  is  not.  It  is  intriguing 
that  HO— CH2CH2— Hg+  is  95%  split  under  the  conditions  of  the  lability 
test,  whereas  the  other  two  compounds  are  completely  stable;  it  is  also 
surprismg  that  HO^— Hg+  and  HgN— <^— Hg+  are  73%  and  88%  split, 
respectively,  while  "OOC — (f — Hg+  and  (f — Hg+  are  not  split  at  all.  The 
structural  requirements  of  lability  seem  to  be  very  rigid.  Could  it  be  that 
the  structural  requirements  for  splitting  are  the  same  as  for  combination 
with  a  receptor  to  produce  diuresis,  splitting  not  being  a  necessary  prelude 
to  an  effect?  It  is  clear  that  a  final  decision  as  to  these  important  matters 
cannot  be  reached  at  this  time  and  that  more  data  must  be  accumulated; 
one  would  like  to  have  some  information  on  the  rates  of  splitting  of  mer- 
curials in  homogenates  or  extracts  of  kidney,  and  the  effects  of  pH  on  this. 


936  7.    MERCURIALS 

Mechanism  of  Transport  Inhibition 

If  one  knows  essentially  nothing  of  the  cellular  site  of  action  of  the  mer- 
curials, and  is  completely  ignorant  of  the  molecular  nature  of  ion  transport, 
it  is  difficult  to  discuss  possible  mechanisms  of  inhibition  without  becoming 
ethereal.  All  the  evidence  points  to  a  lack  of  significant  depression  of  the 
exergonic  phases  of  renal  metabolism  at  concentrations  markedly  affecting 
transport,  so  it  is  likely  that  the  action  is  on  some  component  of  the  func- 
tional system.  If  the  postulated  specific  receptor  for  mercurials  is  a  carrier, 
then  there  is  the  problem  of  accounting  for  the  inhibition  of  many  types 
of  transport;  furthermore,  it  is  not  at  all  certain  that  a  carrier  is  involved 
in  ion  transport.  On  the  basis  of  what  was  said  in  the  previous  section  on 
the  effects  of  mercurials  on  membranes  and  permeabilities,  it  is  most  likely 
that  the  site  of  action  is  the  tubular  cell  membrane.  The  Na+  pump  is  prob- 
ably located  in  the  peritubular  membrane  and  the  diffusion  of  Na+  across 
the  lumenal  membrane  is  passive  along  concentration  and  electrical  gra- 
dients. The  fact  that  the  transmembrane  potential  is  around  —  43  mv  at 
the  lumenal  surface  and  —  64  mv  at  the  peritubular  surface  was  interpreted 
by  Giebisch  (1960)  as  indicating  the  greater  Na+  permeability  of  the  lu- 
menal membrane.  Mercurials  could  thus  depress  Na+  resorption  by  acting 
in  three  ways:  (1)  inhibition  of  the  Na+  pump,  (2)  decrease  of  the  permeabi- 
hty  to  Na+  of  the  lumenal  membrane,  or  (3)  increase  of  the  permeability 
to  Na+  of  the  peritubular  membrane.  It  has  frequently  been  assumed  that 
the  mercurials  inhibit  active  transport  directly  but,  as  has  been  discussed 
for  other  systems,  it  is  possible  that  the  primary  effect  is  on  the  membrane 
structure  controlling  permeability.  The  evidence  from  the  changes  in  elec- 
trical potentials  brought  about  by  chlormerodrin  shows  that  both  mem- 
branes are  affected  (Giebisch,  1961).  The  potential  across  the  peritubular 
membrane  is  decreased  to  —25  mv  and,  since  the  transtubular  potential 
is  simultaneously  decreased  to  —  7  mv,  it  would  seem  that  the  lumenal 
membrane  potential  is  decreased  to  —18  mv.  These  changes  can  be  in- 
terpreted as  due  to  increases  in  the  Na+  permeability  of  both  membranes, 
but  it  is  also  possible  to  conclude  that  there  is  a  general  decrease  in  the 
ionic  permeability.  Until  information  on  the  effects  of  mercurials  on  ion 
fluxes  is  available,  one  cannot  distinguish  between  these  possibilities.  The 
results  of  Mudge  (1951)  on  rabbit  kidney  slices  point  either  to  an  increase 
in  permeability  to  Na+  and  K+,  or  to  an  inhibition  of  active  transport. 
However,  the  relationship  of  these  in  vitro  results  to  mercurial  diuresis  is 
obscure,  especially  as  Auditore  and  Holland  (1956)  found  that  diuresis  can 
be  produced  without  appreciable  loss  of  cell  K+,  although  the  latter  can 
occur  with  minimal  toxic  doses.  That  mercurials  can  alter  ion  permeabilities 
without  appreciably  depressing  active  transport  is  demonstrated  by  studies 
on  other  tissues,  such  as  atria  (page  945).  Mercurials  could  conceivably 
alter  pore  sizes  by  distorting  membrane  structure,  or  actually  clog  ion- 


EFFECTS    ON   TISSUE    FUNCTIONS  937 

transporting  pores  by  reacting  with  SH  groups  on  the  walls,  or  interfere 
with  the  open-closed  transition  postulated  to  occur  in  the  membrane  by 
Kavanau  (1963).  White  et  al.  (1961)  have  provided  the  only  direct  evidence 
that  mercurials  increase  the  permeability  of  the  proximal  tubule  cells  to 
Na+.  They  infused  Na^'  into  the  renal  artery  during  mannitol  diuresis  and 
found  the  Na+  flux  to  be  increased  by  meralluride  at  diuretic  doses.  It  was 
thus  concluded  that  the  net  resorption  of  Na+  is  decreased  because  of  the 
augmented  backward  leak. 


EFFECTS   ON   TISSUE    FUNCTIONS 

Surprisingly  few  thorough  or  quantitative  investigations  of  the  effects  of 
mercurials  on  tissue  function  have  been  made,  especially  considering  the 
long-known  toxic  and  therapeutic  actions  of  these  substances,  and  most  of 
them  are  on  the  heart.  Since  in  essentially  no  case  have  functional  and 
metabolic  disturbances  been  correlated,  most  of  the  results  will  be  treated 
cursorily  and  presented  principally  to  point  out  some  fields  in  which  inter- 
esting work  may  be  done.  There  is  great  need  for  the  study  of  the  metabolic 
changes  produced  in  tissues  isolated  from  animals  treated  with  mercurials, 
since  only  in  this  way  can  one  be  certain  that  the  effects  observed  are  re- 
lated to  the  in  vivo  interference  with  function.  This  has  been  done  with 
the  kidney  but  is  notably  absent  with  other  tissues.  Much  of  the  in  vitro 
work  with  mercurials  has  been  done  with  relatively  high  concentrations 
(around  1  mM  or  higher)  so  that  it  is  impossible  to  determine  if  the 
results  are  applicable  to  the  effects  seen  in  the  whole  animal.  Indeed, 
studies  of  the  simultaneous  changes  in  fuction  and  metabolism  under 
any  conditions   are  very  rare. 

Skeletal    Muscle 

Resting  rat  diaphragm  treated  with  1-2  niM  mersalyl  soon  exhibits  fibril- 
latory  twitches  accompanied  by  rapid  small  (1  mv)  variations  of  the  mem- 
brane potential,  which  persist  for  several  minutes,  followed  by  a  rise  in 
the  resting  tension  during  the  next  10  min,  and  finally  by  a  further  slowly 
developing  irreversible  contracture  and  loss  of  excitability  (Kuschinsky  et 
al.,  1953).  Stimulation  during  the  early  action  of  mersalyl  produces  a  nor- 
mal contractile  tension  but  there  is  a  marked  retardation  of  relaxation, 
the  duration  of  contraction  increasing  5-  to  10-fold.  Hg++,  on  the  other 
hand,  causes  only  the  slow  contracture  and  loss  of  excitability.  Decame- 
thonium  and  curare  abolish  the  fibrillation  due  to  mersalyl,  indicating  that 
the  action  of  the  mercurial  is  on  the  end-plate.  Furthermore,  chronically 
denervated  muscle  does  not  show  mersalyl  fibrillation.  Inasmuch  as  physo- 
stigmine  and  neostigmine  cause  a  similar  fibrillation,  and  since  mersalyl 


938  7.    MERCURIALS 

inhibits  muscle  cholinesterase,  it  was  concluded  that  the  fibrillation  results 
from  inhibition  of  cholinesterase,  allowing  acetylcholine  to  accumulate 
(Kuschinsky  and  LiiUmann,  1954).  None  of  the  other  actions  of  mersalyl 
appears  to  be  related  to  this  inhibition.  The  delayed  relaxation  was  claimed 
to  be  similar  to  that  produced  by  veratrine  but,  if  so,  it  does  not  provide 
much  information  on  the  mechanism,  since  one  is  ignorant  of  how  vera- 
trine acts. 

Contracture  of  frog  muscle  by  Hg++  had  been  noted  by  Bacq  (1942), 
Beck  and  Bein  (1948),  and  Krueger  (1950).  Bacq  assumed  this  to  be  an 
effect  such  as  that  given  by  iodoacetate,  i.e.,  a  Lundsgaard  effect  due  to 
glycolytic  inhibition,  but  Krueger  showed  that  Hg++,  in  contrast  to  iodo- 
acetate, does  not  bring  about  a  reduction  in  lactate  concentration  during 
rigor.  Kuschinsky  and  Liillmann  (1954)  found  that  mersalyl  causes  a  rapid 
loss  of  muscle  K+  and  attributed  the  initial  rapid  contracture  to  a  depola- 
rization of  the  fibers,  a  conclusion  shared  by  Muscholl  (1958),  who  demon- 
strated a  fall  in  the  membrane  potential  coincident  with  contracture.  How- 
ever, the  delayed  contracture  must  have  another  origin  and,  although  Kus- 
chinsky and  Liillmann  postulated  a  Lundsgaard  mechanism,  there  is  no 
evidence  one  way  or  the  other.  The  action  potential  traces  presented  by 
Muscholl  (1958)  show  that  mersalyl  reduces  the  magnitude  so  that  the  over- 
shoot is  lost,  slows  both  depolarization  and  repolarization,  and  hence  pro- 
longs the  duration  of  the  action  potential,  effects  quite  different  than  those 
seen  in  heart  (page  945).  Frog  muscle  after-potentials  seem  to  be  unaffect- 
ed by  mercurials  between  0.1  and  2  mM  (Macfarlane  and  Meares,  1958). 
Contracture  by  mersalyl  is  dependent  on  Ca++  in  the  medium,  but  this 
may  be  true  for  most  types  of  contracture  (Kutscha,  1961). 

The  contraction  of  glycerinated  muscle  fibers  by  ATP  is  inhibited  by 
Hg++  at  concentrations  around  0.01-0.1  mM  (Godeaux,  1944,  Korey, 
1950;  Hasselbach,  1953;  Edman,  1958)  and  by  mersalyl  at  similar  con- 
centrations (Portzehl,  1952;  Edman,  1959).  A  preparation  contracted  by 
ATP  is  relaxed  upon  addition  of  the  mercurial.  These  effects  are  irrever- 
sible by  washing  or  treatment  with  cysteine.  Weber  and  Portzehl  (1954) 
suggested  that  the  inhibition  of  the  ATP  effect  is  due  to  a  block  of  ATPase 
so  that  ATP  can  act  only  as  a  plasticizer,  but  there  is  also,  according  to 
Edman  (1958),  a  direct  effect  since  there  is  some  relaxation  in  the  absence 
of  ATP. 

A  great  deal  of  work  has  been  done  on  the  behavior  of  muscle  contractile 
proteins  exposed  to  mercurials,  and  the  importance  of  SH  groups  has  been 
conclusively  demonstrated.  The  effects  of  the  mercurials  are  summarized  in 
Table  7-20.  Both  actin  and  myosin  possess  SH  groups  of  differing  degrees  of 
reactivity  and  function.  In  the  complexing  of  actin  and  myosin  to  form 
actomyosin,  it  is  the  SH  groups  of  myosin  which  are  important  (Bailey 
and  Perry,  1947;  Kuschinsky  and  Turba,  1951).  Bailey  and  Perry  felt  that 


EFFECTS    ON   TISSUE    FUNCTIONS 


939 


zo  ^^        --^        o 


tf 


M 

^ ^ 

o 

^H 

, — . 

Oi 

OJ 

o 

_i 

Ol 

IC 

cS 

lO 

^ — ■' 

(M 

IC 

CO 

<— < 

Ci 

05 

O 

* " 

— 

05 

o 

>> 

C5 
05 

Oi 

3 
S 

05 
CO 

cS 

c3 

> 

c 
2 

CO 

>> 

^— ^ 

k4 

lO 

05 

2 

t-( 

<u 

0) 

Eh 

^-^ 

o 

^•^ 

fi 

a> 

s 

o> 

CO 

0) 

O 

O 

3 

Cvl 

3 

a 

«<1 

c 

lO 

H 

i-H 

> — 

O 

eS 

H 

O 

O 

O 

o 
tn 

i4 

05 

1 

C5 

in 

"e 

>> 

C 

S 
g 
O 

c 

cS 

IS 

c 

c« 
>> 

o 

c 

"S 

CO 

fl 

ts 

2 

-0 

CO 

X 

CD 

"S 

'm 

3 
cS 
g 

"S 

c3 

>> 

c 

>> 

c 

>> 

_C 

i 

cS 

3 

S 

3 

o 

o 

J 

c 

O 
C 
O 

3 

3 

-o3 

Si 

CO 

o 

O 

c 
o 
H 

c« 

3 

Drab 
Kusc 
Bara 

c« 

J 

pa 

-«  -c  — 

C     fi     Si 


3     o     O 

^  ;z:  tsi 


o 


CO       I      ^ 

ffi  3 

CO  c 


Ed  ^^ 


^  (6  ci  <~> 


— I       ^       -H       O 


;§        CD    '^ 

to  S 

i,  ffi  §  g    ?».  § 


PQ 


§    g     ?^ 


k- (k_,fc_ji_<+     2;    (2^+     ii  ^  '■^  ^    f^    <^    <^ 


m 


3   Oj 
•2  H 

1^  < 


m  ta 


,-1    "^ 


0*0 


ft 

«>  .s 

>.    3 

2  '-5 


'  %  6 


I -I 

3 


o 

o 

^      W) 

'^  s 

o  ^ 
c  .S 
.2  ^ 


>.   3 


O    42 


+ 

g      M 

a 

3 

O 

o 


o    «« 


Q  fiH        Q  Ph 


«    ?.  'S 


•3    S 


hS  Q  P5 


940  7.    MERCURIALS 

the  binding  might  be  through  SH  groups  but,  as  Gergely  et  al.  (1959)  point- 
ed out,  it  only  indicates  that  SH  groups  are  in  the  vicinity  of  the  binding 
groups.  The  myosin  SH  groups  concerned  with  the  binding  of  actin  react 
with  mercurials  more  readily  than  the  SH  groups  upon  which  ATPase  ac- 
tivity depends  (Fig.  7-22)  (Barany,  1959).  However,  MM  presents  an  ex- 
ception, in  that  it  inhibits  ATPase  and  actomyosin  formation  in  a  parallel 
fashion  (Barany  and  Barany,  1959  a),  possibly  indicating  that  the  size 
of  the  group  on  the  mercurial  is  important.  G-actin  is  more  reactive  than 
F-actin,  due  perhaps  to  shielding  of  the  SH  groups  in  the  polymerized  form. 
G-actin  cannot  polymerize  unless  ATP  is  present  and,  since  mercurials  re- 
lease ATP  from  actin,  the  possibility  of  the  effect  on  the  G-actin  ±^  F-actin 
transformation  being  due  to  an  interference  with  ATP  binding  was  examin- 
ed, but  most  mercurials  were  found  to  cause  a  rapid  loss  of  polymerizability 
without  appreciable  loss  of  ATP  (Drabikowski  and  Gergely,  1963).  When 
the  ATP  is  finally  lost,  the  actin  has  been  changed  irreversibly,  and  there 
is  further  evidence  from  optical  rotation  that  structural  changes  are  pro- 
duced (Tonomura  and  Yoshimura,  1962).  Katz  and  Mommaerts  (1962)  con- 
sider the  six  SH  groups  of  G-actin  to  fall  into  three  categories:  two  rapidly 
reacting,  two  intermediately  reacting,  and  two  slowly  reacting,  only  the 
last  two  being  necessary  for  polymerization.  It  is  interesting  that  the  SH 
groups  of  G-actin  are  made  more  reactive  to  p-MB  by  Mg++  and  less  reac- 
tive by  Ca++  (Katz,  1963).  It  was  postulated  that  Mg++  brings  about  an 
open  configuration  whereas  Ca++  tends  to  produce  a  closed  configuration, 
the  SH  group  being  in  a  crevice. 

The  effects  of  the  mercurials  on  extracted  muscle  proteins  are  certainly 
interesting  and  often  obtained  at  low  concentrations,  but  there  is  at  pres- 
ent essentially  no  way  of  determining  if  they  are  at  all  responsible  for  any 
of  the  changes  observed  in  intact  muscle.  It  would  be  particularly  important 
to  know  if  rigor  is  related  to  any  of  the  actions  on  actomyosin,  but  actually 
most  of  the  actions  described  above  could  not  very  well  explain  why  a 
muscle  goes  into  contracture.  The  mechanisms  by  which  mercurials  alter 
muscle  function  are  thus  obscure,  but  it  is  not  unlikely  that  the  earliest 
effects  are  on  the  permeability  and  transport  systems  in  the  membrane. 
More  information  will  be  provided  in  the  following  section  in  which  the 
effects  of  the  mercurials  on  cardiac  muscle  will  be  discussed. 

Heart 

The  detrimental  effects  of  the  mercurials  on  the  heart  have  long  been 
recognized  and  many  cases  of  clinical  deaths  from  intravenous  injections 
of  mercurial  diuretics  have  been  reported.  It  is  generally  agreed  that  death 
is  attributable  to  the  direct  action  on  the  heart  during  temporary  high 
plasma  concentrations  of  the  mercurials,  whereas  at  the  usual  low  concen- 
trations required  for  diuresis  there  are  no  detectable  cardiac  effects.  The 


EFFECTS    ON   TISSUE    FUNCTIONS  941 

actions  of  the  mercurials  on  the  heart  have  been  well  studied  but  the 
mechanisms  involved,  and  whether  these  actions  involve  metabolic  distur- 
bances, are  not  known.  Hepp  (1887)  observed  that  the  toxic  effects  of  the 
organic  mercurials  differ  from  those  produced  by  Hg++,  and  that  ethyl- 
Hg++  stops  the  frog  heart  in  diastole.  Dreser  (1893)  studied  the  cardiac 
effects  of  several  complexes  of  Hg++,  both  in  vivo  and  on  perfused  isolated 
frog  hearts.  Cardiac  depression  was  noted  after  injection  of  around  2.5  mg 
of  the  rhodanate,  succinimide,  and  cyanide  complexes  of  Hg++,  and  stand- 
still of  the  isolated  heart  was  brought  about  by  0.45  mM  of  the  thiocyanate 
complex.  These  and  other  early  investigations  showed  only  that  the  heart 
can  be  depressed  by  mercurials  and  that  the  relative  potencies  depend  on  the 
substances  with  which  the  Hg++  is  complexed.  For  example,  Miiller  et  al. 
(1911)  showed  that  compounds  of  the  type  R — Hg — OH  or  R — Hg — CN  are 
around  10  times  as  cardiotoxic  in  cats  as  compounds  of  the  type  R — C — 
Hg — C — R  in  which  the  Hg  is  bonded  to  two  C  atoms,  and  that  the  toxicity 
is  related  to  the  rate  at  which  these  compounds  react  with  sulfide.  The  first 
serious  study  of  the  effects  of  the  mercurials  on  the  heart  was  undertaken 
by  Salant  and  his  co-workers  in  Georgia  from  1921  to  1931,  the  results  of 
which  will  be  discussed  throughout  the  following  sections. 

(A)  Isolated  heart  preparations.  The  effects  of  the  mercurials  depend  on 
the  species,  the  type  of  preparation,  the  mercurial  used  and  its  concentra- 
tion (Table  7-21).  As  would  be  expected,  there  is  generally  depression  of 
contractile  amplitude,  rate  of  beating,  and  rate  of  conduction,  the  last  lead- 
ing to  varying  degrees  of  a-v  block  and  dissociation  of  the  atria  and  ven- 
tricles. Ventricular  standstill  often  occurs  before  the  atria  cease  to  beat. 
Tachycardia  and  fibrillation  are  frequently  seen  in  animals  poisoned  acutely 
with  the  mercurials,  but  are  not  noted  in  isolated  preparations,  although 
occasionally  with  low  concentrations,  or  initially,  some  increase  in  rate  and 
contractile  amplitude  may  be  observed.  The  ventricular  dysrhythmias  in 
vivo  may  be  in  part  the  result  of  altered  a-v  and  ventricular  conduction, 
but  in  isolated  preparations  there  is  little  evidence  for  the  appearance  of 
rapidly  discharging  ectopic  foci.  Most  of  these  effects  are  irreversible,  or 
only  partially  reversible  at  the  lowest  concentrations  and  with  short  expo- 
sures, but  dimercaprol  or  cysteine  is  occasionally  able  to  reverse  rather  ad- 
vanced degrees  of  depression  (Ruskin  and  Johnson,  1949).  The  selective  de- 
pression of  the  rat  atrial  rate  is  marked;  at  0.0013  vaM  Hg++  the  rate  is 
reduced  35%  in  15-20  min  while  the  amplitude  is  unaffected,  and  at  0.0025 
raM  the  rate  may  be  inhibited  90%  and  the  amplitude  some  15%  greater 
than  normal  (Berman,  1951).  Glutathione  and  dimercaprol  effectively  pro- 
tect both  the  atria  and  ventricles. 

The  development  of  contracture  is  not  nearly  as  common  with  the  mer- 
curials as  with  iodoacetate,  and  in  fact  has  been  noted  only  in  frog  and  turtle 
hearts.  Rat  atria  treated  with  p-MB  are  slowly  and  markedly  depressed 


942 


7.  MERCURIALS 


W 


tf 


ta 


o 

-»^ 

VI 

2 

o 

c 

o 

>H 

c 

tf 

o 

<! 

o 

s 

s 

& 

w. 

&H 


^   &H 


V 

V 

> 

o 
55 

m 


"S  '^ 


s       '^ 


.S  .Z  /5   0) 


"     "5  Ij   >>  O 


Q  «  P 


o 


A  "J^  TJ 

,,   !h   (I> 


A  "Jl  T3 


O  O  O  O 


O  O 


cq  m 


a  ?i. 


fq 


w 


C5 


EFFECTS   ON   TISSUE   FUNCTIONS  943 


W       §       5         -S  "^  "^  "^ 


5  "« 


pq  ^      P5      Pi        Pi 


•r    c   -r    c 
-^  .2    -§   .2 


Cod 

c3     o      e5 


o 

>, 

'C 

P< 

j= 

>. 

iS 

o 

o 

y, 

o 

<i) 

n1 

i 

V 

Q) 

tf 

1 

A 

cS 

^  4<  -^ 
S   §   S 


VV  VV          VVV  <dV 

M  "m    ^  § 

lO     so     O                                 ,,.  ^„     ,.„  ^     'M     lO      CO     C5     >o  O 

O     ^     (M                                  Sf  M      M  o     O     O      O     O     ^1      Tt<  I 

OOOiO          -H           2  PP  OOOOOOTtHi-HiC  0-1               t^eCCJ 

OOOO          O  --  OOOOOOO-H-HTi*  o               C<lfOfC 

o  com         •             ... 

OOOO        O        O  fOI>  oooooooooo  o            ooo 


o        •=<  -So  

HTpeeT  I*  o  01 


944  7.  MERCURIALS 

• 

but  no  elevation  of  resting  tension  is  seen  even  after  2  hr  (Webb  and  Hollan- 
der, 1959).  It  may  well  be  that  the  mercurials  so  readily  depress  membrane 
functions  —  pacemaker  discharge,  conduction,  etc.  —  that  the  hearts  cease 
beating  before  they  are  depleted  of  ATP.  Mendez  (1946)  noted  that  frog 
heart  treated  with  0.56  mM  p-MB  stops  beating  quite  soon  and  before  full 
contracture  has  developed;  however,  if  the  heart  is  electrically  stimulated 
it  can  be  put  into  complete  rigor.  It  has  been  observed  many  times  that  the 
heart  stops  in  diastole  in  vivo  following  intravenous  injections  of  the  mer- 
curials. 

(B)  Cardiac  effects  in  whole  animals.  Sudden  death  during  or  following 
the  intravenous  injection  of  diuretic  mercurials  clinically  has  usually  been 
attributed  to  ventricular  fibrillation.  In  animals  (usually  cats  and  dogs) 
the  mercurials  produce  the  following  cardiac  effects:  initial  cardiac  depres- 
sion, disturbances  in  a-v  conduction  leading  occasionally  to  a  temporary 
ventricular  bradycardia,  atrial  flutter  or  fibrillation  (rarely),  slowing  of 
ventricular  conduction,  and  often  ventricular  tachycardia  before  the  ter- 
minal fibrillation  (Jackson,  1926  a,  b;  Salant  and  Nadler,  1927;  Macht, 
1931  b;  DeGraff  and  Lehman,  1942).  The  effects  may  be  quite  complex  and 
are  undoubtedly  due  to  a  variety  of  actions.  McCrea  and  Meek  (1929)  felt 
that  one  of  the  major  actions  is  a  descending  stimulation  followed  by  a  de- 
pression of  the  cardiac  conducting  tissue.  The  innervation  of  the  heart 
probably  plays  a  role  in  the  initiation  of  the  dysrhythmias,  since  atropini- 
zation  or  cutting  the  vagi  in  dogs  prevents  ventricular  fibrillation  due  to 
mersalyl  (Jackson,  1926  a).  It  is  also  known  that  epinephrine  potentiates 
the  fibrillatory  action  of  the  mercurials.  The  fall  in  blood  pressure  invariably 
observed  during  intravenous  infusion  of  the  mercurials  must  induce  sym- 
pathetic activity  and  a  rise  in  plasma  catecholamines.  Various  salts  of  Hg++ 
apparently  are  not  so  likely  to  induce  dysrhythmias  as  the  organic  mer- 
curials and  are  more  directly  depressant  (Salant  and  Kleitman,  1922). 
These  effects  are  not  dependent  on  the  vagi  since  they  occur  after  atropini- 
zation.  However,  Hg++  can  produce  conduction  disturbances  and  dysrhyth- 
mias in  dogs,  and  occasionally  ventricular  fibrillation  (McCrea  and  Meek, 
1929).  The  intravenous  lethal  dose  depends  on  the  mercurial,  the  species 
used,  and  the  rate  of  injection;  in  most  cases  it  falls  between  10  and  50  mg 
Hg/kg  for  the  common  diuretic  mercurials  (DeGraff  and  Lehman,  1942; 
Chapman  and  Shaffer,  1947;  Lehman  et  al.,  1950;  Farah  et  al.,  1951).  Inor- 
ganic Hg++  is  somewhat  more  toxic,  the  lethal  dose  usually  being  around 
one  third  to  one  half  that  for  the  organic  mercurials.  The  toxicity  of  Hg++ 
is  dependent  on  the  blood  pH,  being  least  between  7.4  and  7.6,  and  increas- 
ing on  either  side,  especially  between  7.14  and  7.35  (Salant  and  Nadler, 
1927).  Since  comparable  experiments  have  not  been  run  on  isolated 
hearts,  it  is  impossible  to  understand  the  mechanism  for  this  sensitive  pH 
dependence;  it  is  difficult   to   accept  that  a   direct   effect  of  pH   could 


EFFECTS    ON   TISSUE    FUNCTIONS  945 

alter  2-  or  3-fold  the  sensitivity  of  the  heart  to  the  Hg++  ion,  and  it  is 
more  likely  that  secondary  changes  due  to  the  alteration  of  the  pH  are 
responsible. 

The  electrocardiographic  changes  in  dogs  are  similar  for  all  the  diuretic 
mercurials  tested  and  for  HgCl,,  and  are  primarily  the  result  of  conduction 
disturbances.  They  may  be  summarized  briefly  as  follows:  depression  and 
change  of  configuration  of  the  st  segment,  increase  in  height  of  the  t  wave, 
widening  and  notching  of  the  qrs  complex,  widening  of  the  p  wave,  and 
increase  of  the  p-r  interval  (McCrea  and  Meek,  1929;  Farah  et  al.,  1951). 
In  the  rat  there  is  an  initial  flattening  of  the  t  wave,  and  eventually  the 
p  wave  may  disappear  (Gessler  and  Kuner.  1960).  Most  of  these  changes 
are,  of  course,  simply  due  to  the  slowing  of  conduction  throughout  the 
myocardium.  The  t  wave  changes  are  different  from  those  seen  with  most 
metabolic  inhibitors  and  are  perhaps  related  more  to  a  membrane  effect 
than  a  metabolic  disturbance.  It  is  interesting  that  p-MB  acts  differently 
than  the  mercurial  diuretics  in  that  no  qrs  changes  are  seen,  even  at  doses 
2-3  times  the  lethal  doses  of  the  other  mercurials,  and  death  is  not  due  to 
fibrillation  but  to  ventricular  asystole  (Farah  et  al.,  1951).  The  lethal  dose 
of  p-MB  is  also  about  4  times  as  great  as  for  the  other  mercurials. 

(C)  Transmembrane  potentials  and  ionic  shifts.  The  membrane  charac- 
teristics of  rat  atria  are  changed  markedly  by  0.05  mM  p-MB,  although 
the  rate  of  action  is  rather  slow  (this  is  probably  not  due  to  slow  penetra- 
tion into  the  atria  since  the  potentials  are  recorded  from  cells  at  the  sur- 
face) (Webb  and  Hollander,  1959).  During  the  first  20-30  minutes  there  is 
no  significant  alteration  of  the  contractile  behavior,  but  there  is  a  progres- 
sive reduction  in  the  magnitude  of  the  action  potential,  an  acceleration  of 
the  repolarization  rate,  and  a  slowing  of  conduction.  It  is  possible  that  these 
early  effects  arise  from  selective  action  on  the  cell  membranes.  During  the 
next  hour  these  changes  continue  but,  in  addition,  contraction  becomes  im- 
paired. At  1  hr  the  changes  may  be  summarized  as  follows:  no  significant 
change  in  resting  potential  (  +  2.1%),  a  severe  depression  of  the  action 
potential  magnitude  (  —  29%),  a  faster  repolarization  (  +  51%)  leading  to 
a  shorter  action  potential  (—60%),  a  decrease  of  the  developed  tension 
(  —  48%),  a  slowing  of  conduction  (—38%),  and  a  prolongation  of  the 
latent  period  (+61%).  Even  during  this  later  period  it  appears  that  the 
contractile  depression  is  due  mainly  to  the  shortening  of  the  action  potential, 
and  to  some  extent  to  its  reduced  magnitude,  and  there  is  little  evidence 
for  direct  effects  on  the  contractile  systems.  It  may  well  be  that  p-MB 
penetrates  into  the  cells  rather  poorly  and  that  some  of  the  other  mercurials 
would  not  have  so  selective  an  action  on  the  membrane.  Stein  et  al.  (1960) 
reported  that  mersalyl  (0.2  mM)  causes  a  faster  repolarization  and  contrac- 
tile depression  in  guinea  pig  atria,  but  no  changes  in  either  the  resting  or 
action  potential  magnitudes  were  observed.  The  failure  of  the  mercurials 


946  7.   MERCURIALS 

to  affect  the  resting  potential  in  either  rat  or  guinea  pig  atria  indicates 
that  no  marked  changes  in  intracellular  K+  occur  during  the  duration  of  the 
experiments,  so  that  appreciable  depression  of  ion  pumps  or  increase  in  ion 
permeabilities  seem  not  to  be  a  characteristic  of  the  action  on  the  heart. 
It  is  difficult  to  interpret  the  cardiac  ionic  changes  noted  by  Gessler  and 
Bass  (1960)  in  rats  poisoned  with  HgClg,  because  the  electrolyte  changes 
resulting  from  the  renal  effects  (either  polyuria  or  anuria)  probably  com- 
plicate the  picture.  However,  with  a  dose  of  HgClg  sufficient  to  produce  a 
long-lasting  polyuria  there  is  only  a  minor  fall  in  the  myocardial  K+/Na+ 
ratio  (1.78  to  1.60),  and,  although  plasma  K+  rises,  the  tissue/plasma  ratio 
for  K+  certainly  does  not  drop  very  much,  although  Gessler  and  Kuner 
(1960)  felt  that  the  qt  changes  are  perhaps  correlated  with  alterations  of 
this  ratio.  The  results  on  isolated  atria  support  the  concept  that  the  major 
effect  is  on  the  ionic  flux  rates  during  membrane  activation. 

The  various  regions  of  the  heart  respond  differently  to  the  mercurials 
as  they  do  to  other  inhibitors  and  drugs.  Isolated  pig  ventricle  fibers  are 
not  appreciably  affected  by  0.26  milf  p-MB  but  the  Purkinje  fibers  are 
more  sensitive  (Kleinfeld  et  al,  1964).  The  action  potential  magnitude  in 
the  ventricle  may  fall  around  10%  within  15  min  but  there  is  little  further 
change,  while  the  resting  potential  and  action  potential  duration  are  not 
significantly  modified.  In  the  Purkinje  fibers,  on  the  other  hand,  the  magni- 
tude of  the  action  potential  is  rapidly  depressed,  falling  approximately  25% 
within  10  min,  after  which  another  rapid  fall  occurs  between  20  and  30  min. 
Since  the  resting  potential  is  unchanged  for  20  min,  there  is  initially  a 
marked  decrease  of  the  overshoot;  the  resting  potential  later  falls  gradually. 
The  duration  of  the  action  potential  is  surprisingly  not  altered  in  contrast 
to  the  results  in  atria.  Although  no  evident  explanation  for  these  differences 
is  at  hand,  it  was  considered  that  the  greater  glycolytic  activity  of  the 
Purkinje  fibers  might  predispose  them  to  inhibition.  We  have  seen,  however, 
that  the  glycolytic  pathway  is  probably  less  sensitive  than  the  cycle  to  the 
mercurials. 

(D)  Cardiac  innervation  and  responses  to  acetylcJwline  and  epinephrine. 
Salant  and  Kleitman  (1922)  noted  that  Hg++  exerts  some  vagal  blocking 
action  in  the  cat  heart,  but  Jackson  (1926  b)  could  find  no  acceleration  of 
the  heart  and  no  evidence  of  vagal  block  by  mersalyl  in  dogs.  Salant  and 
Brodman  (1929  a)  reinvestigated  this  question  and  established  that  Hg++ 
first  sensitizes  the  heart  to  the  vagus  and  later  blocks  the  vagal  endings. 
It  is  during  the  first  sensitization  phase  that  dysrhythmias  are  apt  to  occur, 
which  is  reasonable  since  acetylcholine  is  profibrillatory  as  a  result  of  its 
marked  shortening  of  the  action  potential  duration.  Hg++  and  p-MB  at 
0.01-0.1  niM  antagonize  the  effects  of  acetylcholine  on  the  frog  heart,  the 
mercurials  being  allowed  to  act  for  3-40  sec  and  then  washed  out  (Pohle 
and  Matthias,  1959).  It  was  concluded  that  the  acetylcholine  receptors  may 


EFFECTS    ON   TISSUE    FUNCTIONS  947 

contain  SH  groups,  a  conclusion  previously  reached  by  Turpaev  (1955). 
Nistratova  and  Turpaev  (1959)  titrated  the  SH  groups  in  a  frog  ventricle 
homogenate  and  found  that  the  presence  of  acetylcholine  alters  the  shape 
of  the  titration  curve,  but  not  the  total  number  of  SH  groups  titrated  — 
part  of  the  SH  groups  becomes  less  reactive  in  the  presence  of  acetylcholine. 
Inasmuch  as  cholinesterase  is  inhibited  to  some  extent  by  mercurials  (Ta- 
ble 7-13).  it  is  possible  that  this  can  account  for  the  vagal  sensitization,  a 
secondary  blocking  of  the  receptors  for  acetylcholine  reversing  this  effect. 
There  is  no  evidence  for  specific  interference  with  the  action  of  .the  cate- 
cholamines on  the  heart,  but  epinephrine  potentiates  the  profibrillatory 
action  of  the  mercurials  (Jackson,  1026  a).  Yet  Salant  and  Brodman  (1929 
c)  claimed  that  the  cat  heart  is  most  sensitive  to  Hg++  when  the  sympa- 
thetics  are  blocked  by  ergotamine,  and  that  high  concentrations  of  epineph- 
rine actually  protect  the  heart  against  the  mercurials.  In  any  event,  it  is 
likely  that  the  over-all  effects  of  the  mercurials  on  the  heart,  especially  in 
the  whole  animal,  must  to  some  extent  involve  the  sympathetic  and  para- 
sympathetic innervation.  Mercurials  can  release  catecholamines  from  ad- 
renal medulla  granules  (D'lorio,  1957)  but  it  is  not  known  if  such  a  release 
can  occur  in  the  heart  or  other  tissue. 

(E)  Consideration  of  some  mechanisms  of  cardiac  action.  There  is  little 
justification  for  discussing  mechanisms  by  which  the  mercurials  affect  the 
heart  because  essentially  no  basic  work  to  elucidate  the  cellular  actions  has 
been  done.  The  interesting  observations  of  Salant  and  Nagler  (1930,  1931) 
on  the  relation  of  the  response  to  Hg++  of  the  frog  heart  and  the  level  of 
Ca++  in  the  medium  may  provide  some  clue.  If  the  Ca++  is  reduced  to  around 
one  half  normal,  the  heart  is  depressed  much  more  readily  by  Hg++,  but  if 
the  Ca++  is  reduced  further  (this  in  itself  suppressing  contractions),  Hg++ 
may  then  actually  stimulate  the  amplitude.  High  Ca++  somewhat  antago- 
nizes the  action  of  Hg++.  Increasing  the  K+  slows  the  rate  and  then  Hg++ 
accelerates  the  heart  and  seems  to  have  less  effect  on  the  contraction.  The 
authors  suggested  that  the  alterations  in  response  to  Hg++  might  be  due 
to  permeability  changes  brought  about  by  Ca"^+,  but  we  now  know  that 
Ca++  has  other,  perhaps  more  important,  effects  on  the  heart.  It  would 
be  interesting  to  know  how  mercurials  affect  the  positive  inotropic  action 
of  Ca++.  It  would  also  be  worthwhile  to  determine  if  mercurials  inhibit  the 
various  ATPases  of  the  heart.  Padykula  and  Herman  (1955)  showed  histo- 
chemically  that  p-MB  strongly  inhibits  cardiac  ATPase,  but  it  is  not  known 
if  this  occurs  in  vivo. 

For  the  purpose  of  this  volume  it  would  be  of  some  importance  if  the 
cardiac  effects  could  be  correlated  with  any  of  the  well-known  enzymic  or 
metabolic  inhibitions  exerted  by  the  mercurials,  but  this  cannot  be  done 
because  there  are  no  investigations  of  metabolic  changes  during  mercurial 
action.  Even  the  results  reported  on  respiratory  inhibition,  for  example  by 


948  7.   MERCURIALS 

Ruskin  and  Ruskin  (1953),  where  5.8  mM  meralluride  depresses  rat  heart 
slices  57%,  are  scarcely  pertinent  to  understanding  how  the  mercurials  act. 
We  have  postulated  previously  that  the  site  of  mercurial  action  on  the  heart 
is  mainly  at  the  membrane  to  alter  ionic  fluxes.  This  does  not  imply  that 
the  action  is  nonmetabolic,  since  enzymes  in  the  membrane  may  be  the 
ultimate  vulnerable  points  of  attack. 

Smooth  Muscle 

It  is  necessary  to  consider  briefly  the  effects  of  the  mercurials  on  smooth 
muscle,  if  only  because  calomel  was  used  for  centuries  as  a  purgative.  We 
have  not  discussed  Hg+  because  little  is  known  about  its  actions  on  meta- 
bolic systems.  Hg++  salts  are  also  capable  of  causing  diarrhea,  and  it  is 
likely  that  Hg+  is  active  after  being  oxidized  to  Hg++.  Hand  et  at.  (1943) 
developed  histochemical  tests  for  Hg,  Hg+,  and  Hg++,  and  found  in  various 
animals  that  within  a  few  minutes  of  the  intravenous  injection  of  mercurous 
acetate  they  could  detect  both  Hg+  and  Hg++  in  the  parenchymal  and  en- 
dothelial cells  of  the  kidney,  the  latter  predominating.  Whole  blood  oxi- 
dizes Hg+  to  Hg^^  quite  rapidly.  Many  theories  of  the  mechanism  of  the 
purgative  action,  several  rather  fanciful,  have  been  advanced,  but  very 
few  justify  even  serious  criticism.  HgClg  did  not  achieve  its  name  of  "cor- 
rosive sublimate"  in  vain;  it  is  a  direct  irritant  of  tissues,  by  which  is  meant 
that  it  induces  cellular  damage  of  both  metabolic  and  nonmetabolic  origin, 
this  initiating  an  inflammatory  sequence,  which  in  the  intestine  causes  in- 
creased activity,  depression  of  the  ability  to  absorb  water  and  various  sub- 
stances, and  consequently  colitis  and  diarrhea.  In  severe  mercury  poison- 
ing there  is  a  hyperemic  and  hemorrhagic  appearance  of  the  intestine,  with 
erosion  and  necrosis.  One  is  thus  tempted  to  attribute  the  diarrhea  to  such 
a  nonspecific  action,  but  there  is  some  evidence  against  this.  Isolated  in- 
testine is  stimulated  in  a  characteristic  way  by  Hg++  in  low  concentrations 
(0.004-0.02  mM),  tonic  contractions  being  markedly  increased  with  sup- 
pression of  rhythmic  activity  (Salant  and  Brodman,  1929  b).  Organic  mer- 
curials apparently  can  stimulate  similarly;  e.g.,  merbaphen  augments  peri- 
staltic activity  of  isolated  cat  intestine  (Govorov,  1936),  and  p-MB  at 
0.0057  mM  stimulates  the  rat  intestine  around  10%  (Goodman  and  Hiatt, 
1964),  although  p-MB  was  reported  to  inhibit  rabbit  intestine  (Haley,  1945), 
perhaps  because  of  too  high  a  concentration  (not  given,  but  around  0.02 
mM).  The  stimulations  produced  by  both  Hg++  and  merbaphen  are  readily 
blocked  by  atropine,  indicating  that  the  action  is  to  some  extent  mediated 
through  the  vagal  nerves  in  the  intestine.  Govorov  believed  merbaphen  to 
be  a  parasympathomimetic  substance.  There  is  also  some  increase  in  the 
sensitivity  of  the  intestine  to  vagal  stimulation  when  the  tissue  is  treated 
with  Hg++,  recalling  similar  actions  on  the  heart,  and  it  is  possible  that 
inhibition  of  cholinesterase  is  involved.  However,  it  seems  unlikely  that 


EFFECTS   ON   TISSUE    FUNCTIONS  949 

mercurials  in  vivo  can  produce  intestinal  stimulation  by  such  a  selective 
inhibition,  and  the  subject  needs  further  investigation. 

Another  mechanism  which  must  be  given  serious  consideration  is  hista- 
mine release.  Bachmann  (1938)  showed  that  the  isolated  cat  intestine  ex- 
posed to  Hg++  releases  a  substance  which  behaves  like  histamine  pharma- 
cologically, and  felt  that  at  least  some  of  the  action  on  the  intestine  can  be 
explained  by  this  release.  It  may  be  mentioned  that  Hg++  has  been  report- 
ed to  release  histamine  from  perfused  dog  Kver  (Feldberg  and  Kellaway, 
1938)  and  p-MB  to  release  histamine  from  rat  mast  cells  (Bray  and  Van- 
Arsdel,  1961),  but  in  both  cases  the  concentrations  used  were  too  high  to 
enable  correlation  with  in  vivo  effects;  it  is  quite  likely  that  any  substance 
at  high  enough  concentration  or  any  irritant  histotoxic  agent  will  release 
histamine. 

Nervous  System 

Neurological  dysfunction  is  common  in  mercury  poisoning  (page  951 )  but 
it  is  not  known  if  the  action  is  axonal  or  synaptic.  One  usually  assumes  that 
metabolic  disturbances  affect  primarily  junctional  transmission.  Halasz  et 
al.  (1960)  have  shown  that  transmission  in  the  cat  superior  cervical  gan- 
glion is  rapidly  and  reversibly  depressed  by  p-MB  at  0.0056-0.02  mM,  while 
simultaneously  the  effects  of  injected  acetylcholine  are  potentiated.  If  the 
concentration  is  increased  toward  0.028  mM,  this  potentiation  of  acetyl- 
choline is  lost.  The  stimulatory  action  of  K+  is  unaffected  by  lower  and 
depressed  by  higher  concentrations.  Inhibition  of  acetylcholine  synthesis  is 
apparently  not  involved  since  there  is  a  store  of  acetylcholine  and  the  de- 
pression of  transmission  is  immediate,  so  the  authors  postulate  a  reduction 
of  the  response  of  the  postganglionic  cells  to  acetylcholine.  However,  at  the 
time  of  the  initial  suppression  of  transmission  there  is  actually  a  potentia- 
tion of  the  acetylcholine  response,  which  is  difficult  to  explain,  particularly 
since  cholinesterase  inhibition  is  not  a  likely  hypothesis  for  ganglia.  It  is 
possible  that  SH  groups  of  the  acetylcholine  receptors  are  reacted  at  higher 
concentrations  of  the  mercurial,  as  has  been  suggested  for  cardiac  receptors. 
Recordings  of  the  postganglionic  membrane  potential  changes  are  needed 
to  interpret  these  results. 

Axonal  conduction  is  also  depressed  by  p-MB  at  low  concentrations  (H. 
M.  Smith,  1958).  Conduction  in  the  frog  sciatic  nerve  is  blocked  in  4  min 
by  0.002-0.02  mM  p-MB  and  in  lobster  giant  axon  in  3  min  by  0.045-0.07 
mM  p-MB.  There  is  a  gradual  depolarization  of  the  axon  but  block  occurs 
long  before  the  potential  is  lost.  The  post-tetanic  hyperpolarization  of  sym- 
pathetic C  fibers  is  more  sensitive  to  metabolic  inhibitors  than  the  magnitude 
of  the  action  potential,  and  is  decreased  by  mersalyl  at  0.34  mM  (Greengard 
and  Straub,  1962).  However,  the  nature  of  such  a  hyperpolarization  and  its 
significance  for  conduction  are  not  understood.  The  injection  into  the  squid 
axon  of  6  X  10~^  ml/mm  of  7.5  mM  p-MB  is  without  effect  on  the  action 


950  7.   MERCURIALS 

potential  (Brady  et  al.,  1958),  which  may  indicate  that  the  mercurial  must 
act  on  the  external  surface  of  the  membrane  to  block  conduction.  The 
changes  in  the  structure  of  the  myelin  sheath  of  nerves  brought  about  by 
Hg++  at  high  concentrations  (around  10  mM)  are  certainly  not  relevant  to 
acute  experiments  with  low  concentrations,  but  in  chronic  mercury  poison- 
ing it  is  possible  that  sufficient  Hg++  is  incorporated  in  the  myelin  to  disturb 
nerve  function  (Millington  and  Finean,  1958,  1961).  Thus  at  the  present 
time  we  cannot  decide  whether  the  primary  action  of  the  mercurials  is  on  the 
axon  or  on  the  synaptic  regions,  or  on  both,  especially  in  chronic  poisoning. 

Skin 

Various  types  of  skin  reaction  to  the  mercurials  administered  both  sys- 
temically  and  topically  have  been  recognized  for  years.  Some  of  these  are 
undoubtedly  of  the  allergic  or  sensitivity  category  and  need  not  concern  us. 
Mercurial  diuretics,  like  mersalyl,  when  injected  in  small  amounts  into  the 
skin  cause  blisters,  and  Hahn  and  Taeger  (1931)  concluded  that  there  is  a 
relationship  between  diuretic  activity  and  vesication.  Almkvist  (1922)  had 
claimed  that  mercurials  cause  vascular  dilatation  in  the  skin,  with  result- 
ing edema,  by  a  paralysis  of  the  sympathetic  nerves,  but  there  is  little  evi- 
dence that  this  is  a  significant  factor.  Hellerman  and  Newman  (1932)  noted 
that  alkyl  mercurials  are  powerful  vesicants  and  can  cause  a  severe  der- 
matitis. These  early  observations  are  of  interest  in  the  light  of  the  relation- 
ship between  SH  group  reaction  in  the  skin  and  vesication  established  by 
work  on  the  arsenical  war  gases,  and  one  might  postulate  that  the  mercur- 
ials have  a  metabolic  basis  for  their  effects  on  skin,  perhaps  an  inhibition 
of  the  cycle.  Hg++  reduces  frog  skin  potentials  across  both  borders  and  in- 
creases the  outer  membrane  resistance  (Lodin  et  al.,  1963). 


EFFECTS    OBSERVED    IN    THE   WHOLE   ANIMAL 

It  is  difficult  to  present  the  toxicology  of  the  mercurials  concisely  be- 
cause the  effects  depend  on  the  type  of  mercurial,  the  species  considered, 
whether  the  poisoning  is  acute  or  chromic,  the  route  by  which  the  mercurial 
is  taken  into  the  body,  and  many  other  factors.  The  symptoms  of  chronic 
mercury  poisoning  (mercurialism)  in  man  are  quite  variable  and  usually 
not  correlated  with  the  blood  or  urinary  levels  of  mercury.  Frequently 
urinary  mercury  may  be  considerably  higher  than  the  normal  range  and 
yet  no  symptoms  occur;  however,  definite  symptoms  may  sometimes  be  ob- 
served in  those  whose  level  is  in  the  normal  range.  This  lack  of  correlation 
with  urinary  levels  and  the  protean  nature  of  the  poisoning  not  only  make 
diagnosis  frequently  difficult  but  indicate  that  the  individual  pattern  of 
response  must  relate  to  a  number  of  obscure  factors,  such  as  hereditary 


EFFECTS   OBSERVED   IN   THE    WHOLE   ANIMAL  951 

constitution,  vitamin  intake,  electrolyte  balance,  protein  nutrition,  and 
other  imponderables.  The  concentrations  of  mercury  in  the  blood  or  urine 
are,  of  course,  not  the  critical  determinants  in  poisoning  when  the  mercury 
has  been  slowly  taken  into  the  body  over  a  period  of  months  or  years. 
Mercury  is  picked  up  by  the  various  tissues  at  different  rates  and  to  dif- 
ferent degrees,  and  it  is  the  eventual  levels  of  mercury  in  these  tissues  which 
determine  the  toxic  response.  Such  accumulation  may  occur  over  a  long 
time  and  several  weeks  be  required  before  a  balance  between  intake  and 
excretion  is  achieved.  One  factor  which  must  be  of  importance,  but  about 
which  little  is  known,  is  the  concentrations  of  the  various  thiols  in  the  blood, 
since  this  will  not  only  alter  the  over-all  tissue  uptake  but  will  modify  the 
pattern  of  distribution  in  the  body.  Most  mercurialism  in  adults  is  industrial 
in  origin  and  due  to  the  inhalation  or  ingestion  of  small  amounts  of  metallic 
mercury  or  mercury  compounds  daily  over  a  prolonged  period. 

General  Symptoms  of  Mercury  Poisoning 

We  have  discussed  the  most  im])ortant  aspects  of  acute  poisoning  by  the 
inorganic  and  organic  mercurials,  namely,  the  effects  on  the  cardiovascular 
and  renal  systems,  and  little  more  need  be  added.  Slow  inhalation  of  mer- 
cury vapor  produces  typical  poisoning  of  the  kind  commonly  seen  with  the 
inorganic  mercury  salts,  because  the  metallic  mercury  is  oxidized  during 
and  after  absorption.  However,  when  the  concentration  of  mercury  vapor 
is  high,  absorption  is  faster  than  oxidation  and  unique  symptoms  are  ex- 
hibited, e.g.,  hyperthermia,  tachypnea,  cough,  nausea,  dizziness,  and  weak- 
ness (Carpenter  and  Benedict,  1909).  These  may  be  due  primarily  to  the 
greater  uptake  of  mercury  by  the  central  nervous  system  under  these  con- 
ditions. It  is  important  to  emphasize  that  the  character  of  mercury  poison- 
ing depends  greatly  on  the  tissue  distribution,  and  hence  on  the  physico- 
chemical  properties  of  the  mercurial.  Thus  the  more  or  less  volatile,  lipid- 
soluble,  alkyl  mercurials  produce  quite  a  different  picture  from  the  inor- 
ganic or  diuretic  mercurials  (Miiller  et  al.,  1911;  Hunter  et  al.,  1940).  Alkyl 
mercurials  act  rather  selectively  on  the  central  nervous  system  to  produce 
ataxia,  paralysis,  and  depression  —  in  higher  concentrations  they  act  much 
like  certain  anesthetics  —  and  acutely  these  effects  are  possibly  unrelated 
to  mercury  or  reactions  with  SH  groups. 

The  acute  effects  of  mercurials  on  the  central  nervous  system  are  often 
marked  but  have  not  been  analyzed  in  detail.  When  HgCL,  is  injected  sub- 
cutaneously  into  rats  at  the  high  dose  of  17  mg/kg,  there  is  progressive  loss 
of  the  reflexes  and  all  have  disappeared  after  54  hr,  this  being  reversible 
upon  administration  of  a  Hg++-binding  thiol  (Galoyan  and  Turpaev,  1958). 
Conditioned  reflexes  are  suppressed  partially  at  the  much  lower  dose  of 
3.7  mg/kg  (Galoyan,  1957).  The  respiration  is  usually  affected  and  may  be 
taken  as  an  index  of  certain  central  actions.  Respiratory  stimulation  by 


952  7.  MERCURIALS 

Hg++  (Hanzlik,  1923  c)  and  mersalyl  (Jackson,  1926  b)  has  been  noted, 
this  being  attributed  to  a  direct  medullary  effect  at  low  mercurial  concen- 
tration, but  lethal  doses  of  both  HgClg  and  PM  cause  dyspnea  and  depres- 
sion of  the  respiration  (Wien,  1939).  The  injection  of  certain  mercurials  can 
produce  a  very  rapid  fulminant  type  of  reaction  characterized  by  dyspnea, 
convulsions,  and  death,  and  this  was  also  attributed  to  a  central  effect  on 
the  respiratory  centers  (Fourneau  and  Melville,  1931).  Direct  effects  on  the 
central  nervous  system  were  observed  by  Pentschew  and  Kassowitz  (1932) 
following  suboccipital  injections  of  HgClj  at  the  minimal  lethal  dose  (around 
0.2  mg);  tremors,  convulsions,  and  other  motor  disturbances  occur  after 
16  hr  and  last  for  several  days.  It  would  be  interesting  to  know  the  form, 
or  forms,  in  which  Hg++  penetrates  into  the  nervous  system,  whether  mainly 
as  the  uncharged  HgClg  or  in  combination  with  thiols  and  other  substances 
in  the  blood. 

The  most  characteristic  symptoms  of  mercurialism,  regardless  of  the  type 
of  mercurial  responsible,  may  be  summarized  as  follows  (Hunter  et  al.,  1940; 
Cumings,  1959,  p.  78;  Noe,  1960;  Kantarjian,  1961).  (1)  A  fine  intention 
tremor,  starting  in  the  fingers  and  hands,  and  progressing  to  the  feet,  eye- 
lids, cheeks,  tongue,  and  neck.  The  motor  activity  is  primarily  affected  and 
usually  there  is  little  if  any  disturbance  in  sensation.  (2)  Insomnia,  anorexia, 
and  various  emotional  alterations,  such  as  mood  depression  and  timidity. 
There  is  generally  little  effect  on  intelligence  or  memory.  (3)  Erethism,  or 
blushing,  is  often  common,  but  whether  it  is  due  to  emotional  disturbances 
or  alteration  of  the  autonomic  vascular  control  is  unknown.  Sometimes, 
especially  in  infants,  the  skin  may  become  red;  such  erythema  is  most  likely 
vascular  in  origin.  (4)  Stomatitis,  salivation,  and  gingival  swelling  are  fre- 
quent and  possibly  due  to  the  secretion  of  mercurials  in  the  saliva.  Of  these 
symptoms,  and  others  less  common,  only  the  nervous  system  changes  lend 
themselves  to  an  analysis  of  the  mechanisms  which  may  be  involved,  but  we 
shall  see  that  regrettably  little  can  be  concluded. 

Urinary  mercury  excretion  in  normal  individuals  is  usually  between  1 
and  15  //g/day,  but  may  be  so  low  as  to  be  imdetectable  or  considerably 
higher  without  obvious  symptoms.  In  patients  with  evident  mercurialism, 
the  urinary  mercury  may  vary  widely  —  excretions  between  3  and  8000 
//g/day  have  been  reported  —  but  it  is  generally  above  250  //g/day.  The 
level  will  depend  on  the  daily  uptake  and  whether  the  individual  is  in  com- 
plete balance  or  not.  The  fact  that  the  same  degree  of  severity  of  symp- 
toms may  be  observed  in  patients  with  very  different  urinary  levels  sug- 
gests that  the  susceptibility  to  mercury  varies  widely,  but  possibly  the 
tissue  concentrations  of  mercury  are  much  more  uniform  than  the  urinary 
concentrations. 

In  view  of  the  selective  effect  of  iodoacetate  on  the  retina  and  visual 
function,  it  is  interesting  to  inquire  as  to  whether  other  SH  reagents,  such 


EFFECTS   OBSERVED  IN  THE   WHOLE   ANIMAL  953 

as  the  mercurials,  possess  this  action.  Since  the  mercurials  seem  to  lack 
marked  effects  on  glycolysis  in  intact  tissues,  one  would  not  expect  visual 
disturbances  if  these  are  indeed  due  to  glycolytic  inhibition.  Sorsby  et  al. 
(1957)  could  detect  no  retinal  degeneration  in  rats  or  rabbits  given  HgClg 
or  p-MB  under  conditions  where  lesions  are  produced  by  iodoacetate.  How- 
ever, in  poisoning  with  certain  organic  mercurials  there  may  be  a  marked 
constriction  of  the  visual  field  (Hunter  et  al.,  1940).  It  is  not  known 
whether  this  is  a  retinal  effect  or  due  to  nerve  degeneration.  The  sjaithesis 
of  rhodopsin  from  opsin  and  retinenci  is  blocked  by  0.1  mM  p-MB  though 
an  action  on  opsin  (Wald  and  Brown,  1951,  1952),  and  the  synthesis  of 
iodopsin  is  likewise  suppressed  (Wald  et  al.,  1955).  The  mercurial  does  not 
alter  rhodopsin  directly  but  readily  bleaches  iodopsin.  If  this  effect  on  the 
regeneration  of  visual  pigments  occurs  in  vivo  it  has  not  been  reported. 

For  many  years  a  rather  uncommon  disease  of  infants  has  been  recog- 
nized and  called  pink  disease,  infantile  acrodynia,  or  erythredema,  and  is 
characterized  by  a  redness  and  swelling  of  the  extremities  and  certain  other 
skin  areas,  along  with  photophobia,  irritability,  loss  of  reflexes,  and  muscular 
hypotonia.  Dr.  Warkany  of  the  Children's  Hospital  in  Cincinnati  in  1945 
examined  an  infant  suffering  from  this  disease  and  found  a  urinary  mercury 
concentration  of  360  ^g/liter.  A  summary  of  20  cases  showed  that  most 
infants  with  pink  disease  had  definitely  elevated  mercury  levels  —  75% 
had  more  than  50  //g/liter  and  10%  more  than  400  //g/liter  ■ —  whereas 
most  control  infants  showed  undetectable  levels  (Warkany  and  Hubbard, 
1948).  It  is  now  generally  agreed  that  the  majority  of  cases  of  pink  disease 
are  due  to  mercury  poisoning,  which  manifests  itself  somewhat  differently 
in  infants  than  in  adults  although  some  of  the  neurological  changes  are 
similar.  The  source  of  the  mercury  is  usually  calomel  or  mercury  ointments. 
Of  the  group  of  54  studied  by  Zellweger  and  Wehrli  (1951),  around  80%  had 
been  given  such  drugs,  but  in  the  remainder  there  was  no  obvious  source 
of  mercury.  The  situation  is  probably  more  complex  than  believed  originally. 
Some  of  the  symptoms  may  not  be  directly  due  to  the  mercury  but  are 
predisposing  conditions  to  mercury  poisoning  (Barrett,  1957).  Thus  a  high 
intestinal  alkalinity,  due  in  infants  to  faulty  acid  secretion  in  the  stomach, 
may  accelerate  the  oxidation  of  calomel  and  increase  its  toxicity,  and  simul- 
taneously reduce  absorption  of  certain  fatty  acids,  this  latter  possibly  be- 
ing responsible  in  part  for  the  acrodynia.  Since  acrodynia  in  animals  may 
be  induced  by  pyridoxine  deficiency,  there  is  also  some  possibility  that  the 
mercury  either  inhibits  some  phase  of  pyridoxine  metabolism  or  blocks  a 
pyridoxal-P  enzyme  to  produce  symptoms  similar  to  deficiency.  Since  typi- 
cal pink  disease  can  occur  in  the  absence  of  excess  mercury  intake  or  signifi- 
cant urinary  levels,  one  must  assume  that  some  basic  metabolic  disturbance 
is  the  basis  of  this  malady  and  that  it  can  be  brought  about  in  various  ways. 
That  aU  the  symptoms  are  not  immediately  due  to  mercury  seems  to  be 


954  7.   MERCURIALS 

indicated  by  the  fact  that  administration  of  dimercaprol  is  not  remarkably 
successful,  although  certain  clinical  improvement  has  been  noted  (Bivings 
and  Lewis,  1948). 

Histological   Changes 

The  neurological  picture  of  fasciculations,  hyperreflexia,  tremor,  and  mo- 
tor weakness,  followed  by  muscular  atrophy,  often  seen  in  chronic  poison- 
ing with  the  organic  mercurials  —  as  reported,  for  example,  by  Kantarjian 
(1961)  in  individuals  eating  bread  treated  with  the  fungicide  Granosan  M 
(ethylmercuri-p-toluene  sulfonanilide)  —  accompanied  by  occasional  numb- 
ness or  paresthesias,  may  clinically  resemble  amyotrophic  lateral  sclerosis 
(Brown,  1954).  Rats  and  monkeys  chronically  poisoned  with  MM  show 
Wallerian  degeneration  of  the  peripheral  nerves  (Hunter  et  al.,  1940).  The 
peripheral  nerves  and  the  posterior  spinal  roots  are  affected  first,  and  later 
the  posterior  columns  and  granular  layer  of  the  middle  lobe  of  the  cerebel- 
lum. The  ataxia  and  tremor  could  result  from  the  cerebellar  lesions.  Bila- 
teral cortical  atrophy  in  the  area  striata  was  associated  with  the  reduction 
of  the  visual  field.  A  good  review  of  the  neurological  changes  has  been  pro- 
vided by  Noe  (1960).  The  renal  and  intestinal  changes  have  been  described, 
and  we  shall  only  note  that  degeneration  of  the  liver  has  also  been  observed 
(MacNider,  1918  b).  Dogs  given  HgClg  orally  (15  mg/kg)  exhibit  a  deposi- 
tion of  fat  in  the  cells  surrounding  the  central  vein,  followed  by  cloudy 
swelling  and  necrosis,  with  eventual  extension  to  the  periphery  of  the  lobule. 
Such  hepatic  changes  could  well  be  responsible  for  some  of  the  over-all 
metabolic  disturbances  observed  in  animals. 

Foulerton  (1921)  believed  that  Hg++  has  an  affinity  for  lipids,  not  only 
because  of  the  solubility  of  HgClg  in  fat  but  also  due  to  the  formation  of 
oleates,  and  is  transported  to  the  liver  in  the  circulating  blood  fat.  The  liver 
damage  then  results  in  a  defective  lipid  metabolism.  There  are  certainly 
definite  disturbances  in  lipid  metabolism  —  for  example,  Ogilvie  (1932) 
found  an  immediate  and  considerable  rise  in  blood  lipid  following  adminis- 
tration of  HgCla;  subsequently  there  is  a  fall  and  a  second  rise  —  but  no 
evidence  to  indicate  the  mechanism  involved.  Toxic  doses  of  mersalyl  in 
rats  produce  hypoglycemia  and  reduce  the  liver  glycogen  to  essentially  zero 
(Dzurik  et  al.,  1963).  It  was  stated  that  these  effects  are  secondary  to  renal 
dysfunction  and  not  a  manifestation  of  a  direct  action  of  the  mercurial  on 
the  tissues,  but  this  seems  unlikely  and  there  are  possibly  several  factors 
of  importance,  including  epinephrine  release  as  a  result  of  a  nonspecific 
stress  reaction.  Free  amino  acids  in  the  livers  of  rats  given  HgClg  for  10 
days  were  determined  by  Thoelen  and  Pletscher  (1953),  and  it  was  shown 
that  although  serine,  leucine,  and  phenylalanine  do  not  change  significantly, 
cystine  rises  to  3  times  the  control  value.  This  was  interpreted  as  a  detoxi- 
fication response.  One  must  question  the  ability  of  animals  to  increase  the 
synthesis  of  specific  amino  acids  for  the  purpose  of  complexing  with  a  non- 


EFFECTS   OBSERVED   IN  THE   WHOLE   ANIMAL  955 

physiological  metal  ion.  Might  it  not  be  assumed,  with  as  little  evidence,  that 
reduction  of  the  free  cysteine-cystine  concentration  by  the  formation  of 
Hg++  complexes  would  stimulate  synthesis  of  these  amino  acids? 

Toxic   and    Lethal    Doses 

A  few  of  the  results  on  different  types  of  mercurial  are  summarized  in 
Table  7-22.  The  lethal  dose  will  depend  on  the  time  interval  chosen  for  de- 
termination of  the  mortality,  since  death  from  mercurial  poisoning  may 
occur  several  days  following  the  administration,  and  this  accounts  for  some 
of  the  variability  seen  in  the  table.  It  is  clear  that  the  organic  mercurials 
are  generally  less  toxic  than  HgCl,.  It  is  difficult  to  estimate  average  doses, 
but  roughly  the  LD50  for  HgClg  is  near  10  mg  Hg/kg  (0.05  millimole/kg) 
and  for  the  organic  mercurials  around  40  mg  Hg/kg  (0.2  millimole/kg); 
there  is  so  much  species  variation  and  differences  between  the  organic  mer- 
curials that  these  figures  are  to  be  taken  only  as  a  crude  basis  for  compari- 
son. Differences  between  routes  of  administration  are  not  as  marked  as  one 
might  expect.  One  of  the  most  important  factors  determining  the  toxicity  of 
mercurials  given  intravenously  is  the  state  of  dissociation  of  the  R — Hg — X 
bond,  where  X  represents  any  ion  or  thiol  either  introduced  with  the  mer- 
curial or  present  in  the  blood.  In  other  words,  the  concentration  of  the  free 
R — Hg+  ion,  which  is  able  to  react  with  the  SH  groups  of  the  tissue  cell 
membranes,  is  a  major  toxicity  determinant.  If  the  mercurial  is  already 
complexed  with  a  thiol,  as  in  mercaptomerin,  the  dissociation  of  the  Hg — S 
bond  will  be  slow  and  little  of  the  mercurial  will  be  bound  to  SH  groups, 
in  either  the  blood  or  the  tissues.  A  second  factor  of  undoubted  significance 
is  the  distribution  of  the  mercurials,  in  both  the  R — Hg — X  and  R — Hg+ 
forms.  An  R  group  or  a  slowly  dissociating  Hg — X  bond  will  favor  penetra- 
tion into  the  central  nervous  system  in  some  instances,  and  this  may  alter 
the  pattern  of  toxicity  from  a  rapid  cardiovascular  death  to  a  slowly  de- 
veloping degeneration  of  certain  nervous  pathways. 

Distribution,   Metabolism,   and    Excretion   of  Mercurials 

Certain  aspects  of  the  fate  of  the  mercurials  in  the  body  have  been  dis- 
cussed relative  to  the  diuretic  action,  and  these  will  be  briefly  summarized. 
(1)  All  mercurials  are  accumulated  in  the  kidney  and  reach  much  higher 
concentrations  in  this  tissue  than  in  others,  although  the  rate  and  degree 
of  accumulation  depend  on  the  structure  of  the  mercurial.  (2)  Mercurials 
are  to  a  great  extent  bound  to  the  plasma  proteins  and  erythrocytes  so 
that  only  a  small  fraction  is  free  to  enter  the  tissues  or  be  filtered  through 
the  glomeruli.  (3)  Some  of  the  mercurials  may  be  secreted  by  the  renal 
tubular  cells.  (4)  Mercurials  are  excreted  in  the  urine  mainly  complexed 
with  thiols  such  as  cysteine.  (5)  Some  organic  mercurials  are  split  to  form 
Hg++  in  the  body,  but  there  is  no  agreement  as  to  the  degree  to  which  this 


956 


7.  MERCURIALS 


S     M 

^    W 

CO 

Q  Tl 

IJI 

"     M 

< 

a 

s 

^—^ 

t> 

u 

p:} 

H 

Li 

P5 


p-( 


05  —   ^ 


r\  &H 


>  rt  PH  ^ 

Qi  rt  -H  F^ 


;C  ffl    5  =^  -Q  .':3  r,    ^      r.     r_     r^_ 


=«  ii; 


II 


3 


S,    i^    -G 


.2   d   3 


"^  ""  52  t; 

t^       -^       ,LJ       r^ 


^  J  ^  O  K  O  H 


^  ^  5  -  :S 


2     ^ 


iC'^t^ioiotMCOfr^co'^^'-H^co^ 


q"  p"  q"  q"  Q  H  C  CT  Q"  £^  £^  Q  Q  Q  Q 


M  02  K   02   H 


HH  O    I— I 


^         ^  >  >  > 

O         I— I  I— I  I— I  HH 


-t^ 

ClO 

ee 

o 

« 

Q 

-^  TO 


o 


o 

be 


w    w 


ft 

6 


EFFECTS   OBSERVED  IN  THE   WHOLE   ANIMAL 


957 


i-H 

JS 

rts 

T 

e 

a 

p-J 

■ 

eS 

o 

<u 

c 

C 

cS 

s 

ft 

^ 

rrt 

OJ 

^ 

rt 

hJ 

o 

PC 

o  o 

C5     05 


ui 


P5  § 


1U 

«i 

C 

C 

cS 

crt 

c« 

1—t 

C 

i= 

1> 

-a 

-C 

'(J 

o 

0) 

<D 

h-1 

hJ 

la^ 

j2  ft 

^  13  o 

^    ^  ^ 

*J     S  i3 

C  V 
O 


(M  M  -H 


h3  § 


P  Q  « 
hJ  J  h^ 


P  P 

P    hJ 


H  H  P 

<i1  -<  kJ 


>     >>     >>>     >> 


S  P 


42 

■^    -°  -u 

eS     c6  gs 

W  «  O 


5   1? 


Oh    P    O 


.2  .2      -S 


S      «      P5  05 


,J2 

be 

w 
6 


p  p 
h-3  J 


§   >   >       (In        >   > 


P5    Pi 


§     ^ 


n  to 

05     C5 


P  P 


05     Oi 

to  to 

05     05 


^^         ^     ^ 


P     P 
P     h^ 


bn 

M 

W 

W 

to 
O 

^ 

^ 

>> 

>> 

<o 

s 

c 

H 

a> 

<o 

^ 

^ 

-c 

Ah 

fM 

H 

w 


c 
o 

a 

3 

» 

u 

0) 

a> 

!-i 

ft 

SI 

•^ 

0) 

1=^    13 

I— I      2 


S    :S 


c    .2 


"ci    -« 


.5     ts 


a;      5 


958  7.   MERCURIALS 

occurs  or  whether  it  is  important  for  the  actions  of  the  mercurials.  (6)  Hg+ 
is  quite  rapidly  oxidized  to  Hg++  in  the  body  and  probably  acts  on  the  tis- 
sues in  the  oxidized  form. 

The  results  of  distribution  studies  are  shown  in  Table  7-23.  Further  data 
on  the  early  distribution  of  several  mercurials  may  be  found  in  Table  7-19; 
rough  values  for  tissues  levels  of  Hg++  in  fatal  human  poisonings  were  pre- 
sented in  Table  1-8-1.  Although  quantitative  comparisons  are  difficult  due 
to  the  widely  different  doses  and  the  various  routes  of  administration,  the 
general  picture  is  clear.  The  concentrations  in  most  tissues  are  not  mark- 
edly different  from  those  in  blood,  but  there  is  slight  accumulation  in  the 
spleen,  moderate  accumulation  in  the  liver,  and  striking  accumulation  in 
the  kidney.  The  central  nervous  system  levels  are  generally  low,  as  expected, 
and  this  must  be  due  mainly  to  the  small  unbound  fraction  in  the  blood, 
.since  even  the  more  lipid-soluble  mercurials  (e.g.  MM)  do  not  readily  pene- 
trate into  the  brain.  Berlin  and  Ullberg  (1963)  gave  single  doses  of  Hg^^^Clg 
intravenously  to  mice  and  determined  the  changing  tissue  levels  over  16 
days.  The  greatest  amount  in  the  central  nervous  system  occurs  in  the  brain 
stem  in  the  area  postrema,  in  the  hypothalamus,  and  in  sites  adjacent  to  the 
lateral  ventricles,  and  the  retention  in  these  regions  is  greater  than  in  other 
tissues.  Essentially  no  Hg++  appears  in  the  fetus  so  that  the  placenta  pre- 
sents a  barrier  to  penetration,  most  of  the  Hg++  being  bound  to  proteins 
and  the  cellular  elements  of  the  blood.  With  the  exception  of  the  kidney 
and  brain  there  appears  to  be  no  obvious  correlation  between  tissue  levels 
and  pharmacological  or  toxic  actions,  and  it  is  possible  that  the  acute  ef- 
-fects,  as  on  the  heart,  may  be  due  to  the  initial  binding  to  the  cell  mem- 
branes rather  than  the  result  of  intracellular  uptake.  Most  of  the  results  in 
the  tables  were  obtained  with  subtoxic  doses,  with  the  exception  of  those 
of  Galoyan  and  Turpaev  (1958)  and  the  cases  of  human  poisoning,  so  it  is 
not  possible  to  obtain  a  complete  picture  of  the  tissue  levels  during  periods 
of  toxic  reactions,  but  it  is  evident  that  rather  low  over-all  concentrations 
occur  in  most  tissues.  When  it  is  considered  that  probably  a  major  fraction 
of  the  tissue  mercurial  is  boiind  to  metabolically  or  functionally  inert  com- 
ponents, it  appears  that  very  little  mercurial  is  required  to  alter  tissue 
activity. 

Loss  of  mercurials  from  the  body  by  urinary  excretion  is  usually  slow. 
Rothstein  and  Hayes  (1960)  determined  the  total  body  content  of  Hg^"* 
in  rats  given  small  doses  of  HgCL  over  a  period  of  100  days,  and  found 
three  distinct  phases:  40%  is  lost  in  5-10  days,  45%  more  during  the  next 
40-50  days,  and  not  over  5%  more  in  the  next  50  days,  so  that  at  100  days 
there  is  still  some  10-15%  of  the  administered  mercury  in  the  body.  When 
jjg203Qj^  is  infused  intravenously  for  periods  up  to  4  hr  in  rabbits,  it  is  found 
that  the  renal  excretion  does  not  exceed  10%  of  the  total  amount  of  Hg^°^ 
passing  through  the  kidneys.  About  50%  of  the  total  dose  is  taken  up  in 


EFFECTS   OBSERVED   IN  THE   WHOLE   ANIMAL 


959 


Ph    s 


pq 


W 


to    Hh 


(M    O    f-^    O    CO 


CO 


rt-#OOOOOOOOOOOQ0 


-J    ^    o 

o  O'  CO 


-hOOOOOOOOO 


o 

o 

o 

o 

C<1 

lO 

o 

o 

o 

o 

lO      'M      -rJH      TjH 

O    O    05    fO 


O     C5 
O    iM 


O    ■<*    I>    o 

O    O    O    rt 

d  d  d  d 


-^   lO   »o   t^  cc   ■*   t^ 


C<l    >0    (M  O    O 

(M  -rj  O         i>  cq 


>  9.> 
1-^  M  I— I 


-tj  +2  -i^ 


p^  P5  p^         P5  05 


o 


O   O   O   i-H    o    o   o 


,i2 

.^ 

>o 

'Cr 

c3 

a1 

)n 

^ 

^ 

■ ' 

■ — ' 

^_^ 

.^-^ 

^^ 

Oi 

e 

LO 

ii-i 

rr^ 

H.^ 

^H 

1 

' 

^ — ' 

r^ 

M 

S 

ft 

eS 

I 

^ 

_ii! 

1 

S-i 

'ii 

<1 

ki 

S^ 

^^ 

fl"! 

00 

,_ 

-Q 

~ 

^H 

"2 

o 

^^ 

(i^ 

f^ 

O 

' ^ 

fO 

O 

Oi 

fM 

Oi 

"^ ' 

o 

:3^ 

''-' 

o 

CO 

S" 

'q 

^ 

c3 

lu 

o 

w 

k4 

(11 

o 

-a 

^ 

c 

Si 

n 

^ 

CO      CC      r-H      O      O 


o  r^  rt  CO  o  <M 

OOOCOiO-^CMO 


cS 

cS 

o3             03 

P5 

P^ 

P5            P5 

W 

-o 

o 

.    S    . 

M 

M    eS     M 

o" 

ffi 

W     "T^    W 

(D 

>> 

>.     o3      >, 

C. 

^ 

J3      >>     C 

+^ 

bC 

o 

«^    ~      ^ 

w 

S 

§  ^  s 

CC 

o 

>■ 

g 

o 

>■ 

i/j 

(—1 

CO 

.S     cj 

-u 

-u 

-2    c 

-O 

j3 

3     3 

"§ 

a 

X.  ^ 

P5      p:; 


►^     ■-     C5 


00  T3  - 

S^  M  O 

f— I  ^  CO 

H  -H  — 


t^     ,^       S 


r-H       -?'  ^       ^^ 


_  t-l 

^7     u      0      S 

^H         (1-1        f^        U 


S    c    c    ?« 


P5 


<u    a) 


cS     cS    .g 


960  7.  MERCURIALS 

the  kidneys  (Berlin  and  Gibson,  1963).  Less  than  1%  of  the  plasma  Hg^"^ 
is  filtered,  due  to  both  protein  binding  and  the  fact  that  nearly  50%  of  the 
Hg203  is  in  the  erythrocytes  and  is  slowly  exchangeable  with  the  plasma. 
Thus  much  of  the  Hg^''^  found  in  the  kidney  tissue  must  come  directly  from 
the  blood  since  the  glomerular  filtration  cannot  account  for  it.  In  the  case 
of  the  mercurial  diuretics,  around  20-40%  is  excreted  during  the  first  day 
(Borghgraef  and  Pitts,  1956;  Calesnick  et  at.,  1960),  but  progressively  smaller 
amounts  are  lost  each  day.  The  kidney  retains  appreciable  mercurial  for 
many  days;  following  injection  of  only  0.2  mg  Hg/kg  of  HgClg  into  rats,  the 
renal  level  is  around  12.4  //g/g  at  52  days  (Rothstein  and  Hayes,  1960),  this 
constituting  90%  of  the  body  mercury.  MM  is  surprisingly  well  retained  in 
the  kidney,  at  32  days  the  level  being  only  30%  reduced  from  that  after 
1  day  (Swensson  et  al.,  1959).  It  is  obvious  that  cumulation  invariably  oc- 
curs when  a  mercurial  is  administered  daily.  Indeed,  HgClg  given  intrave- 
nously every  21  days  leads  to  a  marked  cumulation,  the  total  body  mercury 
after  5  doses  being  twice  that  from  a  single  dose  (Rothstein  and  Hayes, 
1960).  Rats  given  HgClg  subcutaneously  daily  cumulate  mercury  in  several 
tissues  and  require  2  weeks  for  the  rates  of  intake  and  excretion  to  be  equal 
(Friberg,  1956). 

Normal  human  tissues  contain  mercury  because  there  is  a  daily  intake 
in  the  food.  Bread,  flour,  milk,  pork,  and  beef  contain  2-4  //g%  mercury, 
and  certain  vegetables  a  good  deal  more,  depending  on  soil  conditions  and 
sprays  used  (Szep,  1940).  Forney  and  Harger  (1949)  reported  a  wide  va- 
riation in  kidney  mercury  levels  in  normal  human  subjects  (from  0  to  12.7 
mg%  in  92  autopsies)  with  two  thirds  having  concentrations  greater  than 
0.1  mg%.  Liver  levels  were  less  (from  0  to  1.72  mg%)  with  values  greater 
than  0.1  mg%  in  one  third.  Those  having  received  mercurial  medication 
ranged  from  0.94  to  27.5  mg%  mercury  in  the  kidney.  Similar  results  were 
obtained  by  Griffith  et  al.  (1954),  the  mean  values  in  nonmercurialized  cases 
being  0.45  mg%  in  kidney,  0.10  mg%  in  liver,  and  0.026  mg%  in  spleen 
(these  values  are  in  terms  of  wet  weight  to  compare  with  the  results  of 
others,  and  were  calculated  from  the  dry  weight  figures  with  the  data  in 
Table  1-8-3).  Patients  receiving  large  amounts  of  mercurials  for  some  time 
prior  to  death  (mean  of  4.7  g  total)  had  much  higher  concentrations  in  the 
kidney  (3.4  mg%)  and  liver  (0.36  mg%).  It  is  very  interesting  that  normal 
human  liver  contains  around  1  //g/g  wet  weight  of  mercury.  From  the  values 
for  liver  in  Table  7-23  it  is  seen  that  in  many  cases  the  levels  are  lower  even 
though  the  animals  had  been  given  mercurials;  since  control  concentrations 
have  seldom  been  obtained,  it  is  questionable  how  much  of  the  mercury  in 
most  of  the  tissues  is  due  to  the  administered  mercurial  and  how  much  to 
other  sources.  Of  course,  this  does  not  apply  to  studies  with  Hg^*'^.  It  would 
appear  that  human  liver  normally  contains  more  mercury  than  the  rodent 
liver,  but  whether  this  is  of  dietary  origin  or  a  species  difference  is  not  known. 


EFFECTS   OBSERVED   IN  THE   WHOLE   ANIMAL  961 

The  metabolism  of  the  alkyl  mercurials  and  PM  in  the  body  is  of  some 
importance  in  understanding  the  pattern  of  their  effects.  PM  is  one  of  the 
least  stable  mercurials  in  the  body  and  much  of  it  is  split  to  inorganic 
mercury,  little  being  retained  in  any  tissue  but  the  kidneys  (Miller  et  al., 
1960;  Gage  and  Swan,  1961).  Very  little  MM,  on  the  other  hand,  is  split  to 
inorganic  mercury;  it  is  slowly  excreted  and  cumulates  in  certain  tissues, 
such  as  the  brain.  Ethyl-Hg  is  also  retained  well  by  the  tissues,  but  is  ap- 
parently split  to  inorganic  mercury  at  a  moderate  rate  (i.e.,  faster  than  MM 
and  slower  than  PM),  so  that  by  the  seventh  day  only  21%  of  the  total 
mercury  in  the  kidney  is  ethyl-Hg  (Miller  et  al.,  1961).  It  is  odd  that  at  the 
seventh  day  all  the  mercury  in  the  liver  is  ethyl-Hg,  so  that  one  concludes 
that  splitting  does  not  occur  in  the  liver.  However,  only  70%  of  the  blood 
mercury  is  ethyl-Hg,  so  it  seems  that  the  liver  takes  up  some  inorganic 
mercury.  It  is  possible,  of  course,  that  all  the  nonethyl-Hg  mercury  is  not 
inorganic  mercury. 

Toxicity  to  Aquatic  Organisms 

A  discussion  of  the  effects  of  mercurials  on  animals  would  not  be  complete 
without  mentioning  briefly  some  of  the  interesting  work  done  with  marine 
invertebrates,  mainly  in  connection  with  antifouling  programs,  and  with 
fish.  One  would  expect  sea  water  not  to  be  a  favorable  medium  for  the  ac- 
tion of  Hg++  because  of  the  high  concentrations  of  complexing  anions  and 
the  elevated  pH.  The  importance  of  the  medium  is  apparent  in  the  study 
of  the  amphipod  crustacean  Marinogammarus  marinus  by  Hunter  (1949). 
The  minimal  toxic  concentration  of  Hg++  in  sea  water  is  0.074  mM,  while 
in  distilled  water  it  is  only  0.0093  vaM.  Other  factors,  such  as  altered  trans- 
port activity,  may  contribute  to  the  increased  susceptibility.  Hg++  at  0.18 
milf  does  not  depress  the  respiration  of  this  organism,  indicating  that  the 
toxic  effect  is  not  to  be  attributed  to  a  general  metabolic  inhibition. 

Marine  invertebrates  often  show  marked  changes  in  susceptibility  to  Hg++ 
during  development.  This  is  well  illustrated  by  the  results  obtained  on  the 
barnacle  Balanus  balanoides,  the  sensitivity  to  Hg++  reaching  a  minimum 
during  the  free-swimming  cyprid  stage  (see  accompanying  tabulation)  (Pye- 


(Hg++)  for  50%  lethality 
(mM) 


Nauplii 

Stage  III  0.00033 

Stage  IV  0.00085 

Stage  V  0.0011 

Stage  VI  0.0011 

Cyprids  0.011 

Barnacles  0.0026 


962 


7.  MERCURIALS 


finch  and  Mott,  1948).  Although  the  cyprids  are  not  killed  so  readily  by 
Hg+"'",  their  settlement  on  the  substratum  is  reduced  appreciably  by  0.000037 
mM  and  completely  by  0.00018  mM.  The  lack  of  an  open  gut  in  the  cyprids 
is  suggested  as  a  possible  explanation  for  the  reduced  sensitivity.  Support 
for  the  law  that  nothing  is  simple  or  predictable  in  the  response  of  organisms 
to  metal  ions,  although  no  further  support  is  needed,  is  the  fact  that  the 
cyprids  are  less  sensitive  to  Hg++  in  50%  diluted  sea  water  relative  to  normal 
sea  water,  in  contrast  to  the  amphipod  discussed  above.  Another  interesting 
and  complex  phenomenon  was  discovered  by  Barnes  and  Stanbury  (1948) 
in  studying  the  effects  of  Hg++  and  Cu++  on  the  harpacticid  copepod  Nitocra 
spinipes.  Hg++  is  over  1000  times  more  toxic  than  Cu++,  but  when  these 
metal  ions  are  present  together  at  certain  concentrations  the  lethal  effect 
is  greater  than  would  be  expected  on  the  basis  of  their  actions  alone  (Ta- 
ble 7-24).  An  isobologram  for  50%  lethality  provides  a  curve  characteristic 

Table  7-24 
Lethal  Effects  of  Mercury,  Copper,  and  Their  Combinations  on  Nitocra  spinipes'^ 


Animals 

killed  in  24  hr 

(%) 

Hg++ 

(mM) 

No  Cu++ 

Cu++ 

Cu++ 

Cu++ 

Cu++ 

0.0041  mM 

0.041  mM 

0.41  mM 

4.1  mM 

0 

0 

1.3 

11.3 

21.2 

42.5 

0.00026 

0 

9.1 

11.9 

— 

— 

0.00056 

1.4 

14.5 

20.0 

78 

— 

0.0011 

10.0 

12.7 

45.6 

82 

— 

0.0015 

16.7 

50.0 

93.7 

98 

— 

0.0022 

60 

61.8 

100 

100 

— 

0.0026 

72 

76.4 

100 

100 

— 

0.0056 

78 

87.3 

100 

100 

— 

0.011 

84 

100 

100 

100 

— 

0.016 

100 

100 

100 

100 

— 

"  From  Barnes  and  Stanbury  (1948.) 


of  very  definite  synergism  (see  Fig.  1-10-8),  i.e.,  the  curve  is  extremely  con- 
cave upward.  It  was  postulated  that  lowered  vitality  due  to  one  metal  ion 
may  not  allow  the  animal  to  deal  effectively  with  the  other  metal  ion.  For 
example,  Hg++  might  impair  the  excretory  system  so  that  Cu++  would  be 
retained,  since  it  is  known  that  certain  crustaceans  and  mollusks  excrete 
Cu++.  The  synergistic  effects  indicate  that  Hg++  and  Cu++  may  act  by  dif- 


EFFECTS    ON    MITOSIS,   GROWTH,   DIFFERENTIATION  963 

ferent  mechanisms,  and  possibly  simultaneous  attacks  on  different  metabolic 
or  functional  systems  would  be  particularly  toxic. 

Jones  (1946)  studied  the  effects  of  several  metabolic  inhibitors  on  the 
fresh-water  stickleback  Gasterosteus  aculeatus.  When  Hg++  is  present  at  0.02 
mikf,  there  is  a  temporary  stimulation  of  respiration  (+  20-30%)  at  10-20 
min,  followed  by  a  depression  that  reaches  50%  at  55  min,  the  fish  surviv- 
ing for  110  min.  During  the  phase  of  respiratory  increase,  there  is  accelerat- 
ed motility,  a  greater  opercular  activity,  and  a  faster  heart  rate;  it  is  quite 
possible  that  the  rise  in  respiration  is  associated  with  the  greater  functional 
activity.  If  the  Hg++  is  removed  after  the  respiration  has  been  depressed 
50%,  recovery  is  slow  and  erratic,  and  the  respiration  never  recovers  its 
normal  level,  although  after  1  day  the  fish  appear  normal.  The  mechanism 
by  which  the  fish  are  killed  is  unknown  but  possibly  it  is  asphyxial. 


EFFECTS  ON    MITOSIS,  GROWTH,  AND   DIFFERENTIATION 

If  SH  groups  are  particularly  important  in  cell  cleavage,  as  many  have 
believed,  the  mercurials  should  be  effective  growth  inhibitors  and  perhaps 
useful  agents  to  determine  if  these  SH  groups  are  enzymic  or  involved  in 
cytoplasmic  structure.  The  rather  potent  inhibition  of  the  proliferation  of 
many  microorganisms  by  mercurials  has  been  known  for  almost  100  years 
and  will  be  discussed  in  the  following  section,  while  here  we  shall  attempt 
to  analyze  the  mechanisms  by  which  mitosis  of  plant  and  animal  cells  is 
disturbed  by  the  mercurials. 

Eggs  and   Embryos 

Mercurials  at  concentrations  in  the  range  0.001-0.01  mM  usually  inter- 
fere with  cleavage,  even  in  sea  water  and  despite  the  fact  that  much  of  the 
mercurial  in  most  work  is  removed  from  the  medium  because  of  binding. 
Thus  Mathews  (1904)  showed  that  the  formation  of  embryos  from  Fundulus 
heteroditus  eggs  is  50%  blocked  by  0.0048  mM  Hg++,  90%  blocked  by 
0.0095  mM,  and  completely  blocked  by  0.014  raM.  The  effects  of  mercurials 
on  eggs  and  embryos  at  different  stages  of  development  may  be  quite  com- 
plex. Hg++  may  be  parthenogenetic  in  that  it  induces  membrane  elevation 
in  Arbacia  eggs  and  initiates  a  form  of  cleavage  at  0.01-0.1  mM  (Heilbrunn, 
1925).  The  membranes  begin  to  rise  3-5  min  after  addition  of  Hg++  and 
after  12  min  the  cells  may  become  constricted  unequally  or  cleave,  but  the 
relation  of  this  to  normal  division  is  not  clear.  Hoadley  (1930)  studied  these 
effects  more  closely  and  observed  that  0.025  mM  Hg++  (a  concentration 
several  times  that  suppressing  cleavage  completely)  caused,  after  mem- 
brane elevation,  a  clumping  of  the  cortical  pigment  to  one  side  of  the  egg, 
followed  by  an  unequal  constriction  which  pinches  off  a  small  fragment 


964  7.  MERCURIALS 

containing  all  the  pigment,  this  fragment  later  cytolyzing.  Hoadley  claimed 
that  no  true  cleavage  occurs  and  called  the  Hg++-induced  behavior  pseu- 
docleavage.  Both  Heilbrunn  and  Hoadley  felt  that  the  Hg++  acts  primarily 
on  the  cortical  region,  and  Hoadley,  in  addition,  thought  that  the  Hg++ 
may  react  with  the  pigment  itself. 

Kriszat  and  Runnstrom  (1952)  reported  a  strange  phenomenon  occurring 
in  Arhacia  eggs  treated  with  0.028  mill  p-MB.  This  concentration  of  mer- 
curial rapidly  inactivates  the  spermatozoa,  but  some  fertilization  can  occur 
before  this  is  complete.  Fertilization  causes  a  strong  contraction  of  the  cor- 
tical layer,  squeezing  out  the  cytoplasm  into  a  number  of  pigment-free 
lobes,  the  cortex,  containing  all  the  pigment,  shrinking  to  a  small  folded 
sac.  It  was  postulated  that  p-MB  blocks  rather  specifically  those  processes 
reversing  the  surface  contraction  occurring  during  normal  fertilization,  it 
is  quite  possible  that  these  effects  are  exerted  directly  on  the  SH  groups 
of  the  protein  components  of  the  cortex  (or  plasma  membrane,  since  it  is 
difficult  to  differentiate  them),  rather  than  on  enzymes.  The  concentration 
of  mercurial  is  very  critical.  Hg++  at  0.02  n\M  acting  for  20  min  on  Arhacia 
eggs  prevents  development  beyond  the  early  blastula  stage,  but  acting  for 
6  min  has  no  effect  on  motility  or  larvae;  however,  0.025  vaM  acting  for 
3  min  reduces  cleavage  and  interferes  with  development  (Hoadley,  1930). 
If  the  mercurial  is  added  some  time  after  insemination,  the  effects  are 
modified.  Thus  0.05  vaM  p-MB  30  min  after  fertilization  scarcely  interferes 
with  cleavage,  the  delay  in  onset  being  only  1-2  min  and  95%  of  the  eggs 
dividing  (Zimmerman  et  al.,  1957).  Mersalyl  is  less  inhibitory  and  at  2  mM 
the  eggs  cleave  normally,  although  there  is  a  10-15  min  delay;  only  40% 
develop  to  the  blastula  stage.  PM,  on  the  other  hand,  is  very  potent,  indi- 
cating possible  permeability  factors  (Macfarlane  and  Nadeau,  1948).  De- 
velopment of  Tripneustes  esculentus  (sea  urchin)  embryos,  exposed  at  the 
2-4  cell  stage  for  1  hr  to  0.001  mill  PM,  is  inhibited  and  only  2%  reach  a 
motile  blastula  stage.  Many  of  the  embryos  are  abnormal  and  partial  cytol- 
ysis  occurs.  Even  0.00038  m3I  PM  slows  yolk  absorption  although  cleavage 
is  not  affected.  Echinus  miliaris  larvae  exposed  to  0.0005  mill  Hg++  meta- 
morphose, but  there  is  dedifferentiation  of  the  tissues  so  that  the  young 
echinus  is  often  abnormal,  perhaps  possessing  rudimentary  tube-feet  or 
spines  (Huxley,  1928).  Gastrulation  is  a  process  generally  sensitive  to  toxic 
substances  and  this  is  true  for  the  mercurials.  For  example,  frog  dorsal  lip 
explants  are  depressed  rapidly  by  0.1  milf  p-MB  so  that  little  further  de- 
velopment takes  place  (Ornstein  and  Gregg,  1952;  Gregg  and  Ornstein, 
1953).  The  mercurial  seems  to  prevent  certain  movements  and  spreading 
associated  with  gastrulation,  e.g.,  the  stretching  of  the  mesoderm  within 
the  endoderm  and  the  ectodermal  flow  over  the  endoderm.  Not  much  has 
been  done  on  later  embryonic  development,  but  p-MB  injected  into  newborn 
mice  brings  about  varying  degrees  of  neuroblastic  necrosis,  chiefly  in  the 


EFFECTS    ON    MITOSIS,   GROWTH,   DIFFERENTIATION  965 

outer  cortical  zones,  this  to  some  extent  simulating  radiation  injury  (Hicks, 
1953). 

There  is  no  evidence  that  mercurials  disturb  development  by  generally 
depressing  metabolism.  Ornstein  and  Gregg  (1952)  observed  no  effect  of 
p-MB  on  dorsal  lip  explant  respiration  at  a  concentration  blocking  differen- 
tiation, and  Brock  et  al.  (1939)  found  that  to  inhibit  sea  urchin  egg  respira- 
tion requires  20  times  the  Hg++  concentration  necessary  for  cleavage  block. 
The  latter  concluded  that  this  points  to  a  nuclear  effect  as  the  basis  for  the 
inhibition  of  division,  but  this  does  not  follow.  However,  the  difference  be- 
tween SH  reagents  is  well  shown  here,  in  that  arsenite  depresses  respiration 
more  readily  than  division.  Haas  (1941)  also  inclined  to  a  primary  nuclear 
effect,  since  Hg++  produces  demonstrable  damage  to  the  nucleus  at  0.00074 
vaM,  while  0.037  mM  is  required  for  cytoplasmic  damage  in  Anodonta 
(fresh-water  clam)  eggs.  Without  denying  that  Hg++  can  damage  the  nu- 
cleus, one  must  be  pessimistic  as  to  the  reliability  of  determining  the  site 
of  action  of  an  inhibitor  by  visual  inspection;  e.g.,  the  action  could  have  been 
on  the  plasma  membrane  and  be  microscopically  undetectable,  or  a  good 
deal  of  disturbance  in  the  cytoplasm  might  have  been  caused  without  being 
immediately  evident.  Landau  et  al.  (1954)  stated  that  mersalyl  is  an  effec- 
tive ATPase  inhibitor  and  hence  was  tried  on  the  cleavage  of  Arbacia  and 
Chaetopterus  eggs.  Fertilized  eggs  placed  in  2  mM  mersalyl  complete  the 
first  3-4  cleavages,  but  the  furrowing  strength  is  reduced,  as  measured  by 
the  pressure  increases  required  to  prevent  furrowing,  so  an  inhibition  of  the 
gelation  of  the  cortex  in  the  equatorial  region  by  mersalyl  was  postulated, 
this  presumably  being  mediated  through  an  interference  with  ATP  utiliza- 
tion. Since  the  ATPases  from  different  sources  vary  a  good  deal  in  sensi- 
tivity to  mercurials,  one  does  not  know  what  inhibition  to  expect  in  these 
eggs,  and  the  mercurial  concentration  is  so  extremely  high  that  many  me- 
tabolic and  functional  processes  must  be  affected.  Heilbrun  and  Wilson 
(1955)  explained  the  block  of  cleavage  by  mercurials  as  an  inhibition  of  the 
proteolytic  enzyme  system  involved  in  gelation,  without  obvious  evidence. 
The  direct  effects  of  Hg++  on  fibrous  proteins  extracted  by  sea  urchin  eggs 
by  Sakai  (1962)  are  meaningless  because  a  concentration  of  10  mM  was 
used. 

Plants 

Many  organic  mercurials  are  applied  to  seeds,  bulbs,  or  plants  as  fungi- 
cides, but  occasionally  the  plant  tissues  may  be  damaged  and  growth  de- 
pressed. The  persistence  of  the  mercurial  in  the  plant  is  often  remarkable. 
For  example,  carnation  seeds  treated  with  radioactive  PM  at  a  concentra- 
tion causing  growth  abnormalities  in  the  seedlings  produces  plants  which 
at  8-9  weeks  contain  the  mercurial  in  the  cotyledon  leaves,  the  hypocotyl, 
and  the  root  adjacent  to  the  hypocotyl  (Robson  and  Fenn,  1961).  A  very 
interesting  effect,  and  one  illustrating  that  the  mechanisms  by  which  mer- 


966  7.  MERCURIALS 

curials  act  may  often  be  unexpected,  is  the  zinc-deficiency  disease  of  coffee 
trees  in  Kenya  due  to  spraying  with  mercurial  fungicides  (Bock  et  al.,  1958). 
Not  only  do  the  plants  exhibit  typical  signs  of  zinc  deficiency  —  chlorosis, 
abnormal  growth  of  shoots  and  leaves,  and  short  internodes  —  but  the  zinc 
content  is  reduced  to  25%  of  normal.  In  the  promotion  program  for  mer- 
bromin,  Macht  (1931  a)  purported  to  show  that  organic  mercurials  are  less 
toxic  than  Hg++  to  plants,  but  his  data  are  equivocal,  since  Hg++  is  toxic 
around  0.1  mM,  while  merbromin  inhibits  growth  slightly  at  0.0013  mM, 
50%  at  0.043  mM,  and  81%  at  1.29  mM.  The  growth  of  Avena  coleoptiles 
and  of  pea  stems  is  readily  inhibited  by  the  mercurials,  PM  being  much 
more  potent  than  p-MB  (see  accompanying  tabulation)  (Thimann  and  Bon- 


Concentration   for 

Tissue 

Mercurial 

50%  inhibition 
(mM) 

Avena  coleoptile 

p-MB 

0.035 

PM 

0.007 

Pea  stems 

p-MB 

0.4 

PM 

0.02 

ner,  1949).  It  is  likely  that  the  carboxylate  group  prevents  the  p-MB  from 
penetrating  as  well  as  PM.  These  results  will  suffice  to  demonstrate  growth 
inhibition  by  the  mercurials,  and  we  shall  turn  to  what  little  evidence  is 
available  for  the  mechanisms  involved. 

Onion  roots  exposed  to  0.0075  mM  PM  develop  terminal  swellings  and 
growth  is  immediately  stopped;  however,  after  a  day  new  growth  starts 
distal  to  the  swelling  (Macfarlane  and  Nadeau,  1948).  Doubling  the  con- 
centration leads  to  100%  mortality  of  the  roots.  Hg++  even  at  0.05  mM 
does  not  cause  terminal  swelling,  inhibits  growth  only  15%,  and  does  not 
kill  any  of  the  roots.  Macfarlane  (1951)  pointed  out  that  PM  acts  on  onion 
roots  cytologically  like  colchicine,  in  that  spindles  are  abnormal,  chromo- 
some movement  is  impeded,  and  polyploidy  results  in  the  zone  of  cell  en- 
largement proximal  to  the  meristem.  In  addition,  there  is  chromosome 
stickiness,  fragmentation,  and  aggregation.  Although  mitotic  and  chromo- 
somal disturbances  certainly  occur,  there  may  be  some  question  as  to  the 
validity  of  terming  these  effects  mutagenic  or  radiomimetic  (Macfarlane, 
1953).  Meyer  (1948)  also  observed  such  changes  in  the  root  tips  of  Crepis 
capillaris  exposed  to  0.013  mM  Hg++,  the  sister  telophase  nuclei  often  being 
connected  by  chromatin  bridges,  with  some  breakage  and  recombination, 
leading  to  80%  diploid  metaphases  and  1.5%  tetraploid  metaphases.  The 
formation  of  cell  wall  material  in  the  microspores  of  excised  Lilium  henryi 
anthers  is  reversibly  blocked  by  0.01  mM  p-MB  and  the  progress  of  meiosis 


EFFECTS    ON    MITOSIS,   GROWTH,   DIFFERENTIATION 


967 


is  slightly  retarded  (Pereira  and  Linskens,  1963).  Similar  chromosomal 
changes  induced  by  mercurials  have  not  been  reported,  as  far  as  I  know, 
for  animal  cells.  The  growth  stimulation  by  auxin  applied  to  Avena  coleop- 
tiles  is  inhibited  by  p-MB  at  0.3  raM  (Cleland  and  Bonner,  1956),  but  the 
effects  on  auxin  transport  in  sunflower  stem  section  are  complex  in  that 
0.01  mM  p-MB  accelerates  transport  125%,  0.1  mM  depresses  it  25%,  and 
1  mM  blocks  it  completely  (Niedergang-Kamien  and  Leopold,  1957).  It  is 
not  known  if  interference  with  auxin  transport  or  action  is  involved  in 
growth  inhibition. 

The  structure-action  relationships  of  mercurials  acting  on  the  sporelings 
of  the  marine  red  alga  Flumaria  elegans,  reported  by  Boney  et  al.  (1959), 
were  believed  to  demonstrate  the  importance  of  lipophilicity  and  penetra- 
tion (Table  7-25).  The  alkyl  mercurials  are  often  200-300  times  more  toxic 
than  HgClg,  the  branched  chain  compounds  being  less  toxic  than  the  straight 

Table  7-25 

Lethal  Concentrations  for  Plumaria  Sporelings  Exposed 
TO  Mercurials  for  18  Hr." 


Mercurial 


Concentration  for 

50%  lethality 

(mM) 

Potency 
relative  to  HgCl. 

0.0115 

1 

0.000344 

33 

0.000176 

65 

0.000097 

119 

0.000046 

250 

0.000043 

268 

0.000041 

280 

0.000099 

116 

0.000060 

192 

0.000173 

67 

0.000260 

45 

HgCl, 

Hgl, 

Methyl-HgCl 

Ethyl-HgCI 

n-Propyl-HgCl 

n-Butyl-HgCl 

«-Amyl-HgCl 

Isopropyl-HgCl 

Isoaniyl-HgCl 

Phenyl-HgCl 

Phenyl-Hgl 


«  From  Boney  et  al.  (1959). 


chain.  Some  correlation  between  potency  and  the  distribution  ratios  be- 
tween ether  and  water,  and  between  methyloleate  and  water,  was  claimed, 
but  discrepancies  exist.  Similar  relationships  have  been  reported  for  certain 
marine  crustaceans  (e.g.,  Artemia  salina),  but  in  others  (e.g.,  Acartia  clausi) 
there  is  little  difference  in  toxicity  between  the  mercurials.  Inasmuch  as 


968  7.  MERCURIALS 

there  has  been  no  work  on  the  relative  inherent  or  direct  toxicities  to  cel- 
lular processes,  metabolic  or  functional,  or  adequate  comparison  of  their 
abilities  to  react  with  relevant  SH  groups,  it  is  impossible  to  be  certain 
that  the  differences  are  due  solely  to  variation  in  the  penetration.  Indeed, 
one  is  not  sure  that  the  mercurials  all  act  by  the  same  mechanism.  One 
factor  which  is  often  ignored  is  the  role  of  the  size  of  the  side  chain  in 
membrane  processes,  assuming  that  they  all  react  with  identical  SH  groups 
in  the  membrane.  Nevertheless,  such  quantitative  studies  are  valuable  in 
establishing  a  necessary  basis  for  understanding  the  mechanisms  by  which 
the  mercurials  act;  further  work  will  undoubtedly  allow  these  results  to 
be  interpreted  more  readily. 

Mammalian   Cells  and  Tissue  Cultures 

We  must  note  that  growth,  in  common  with  many  other  processes,  may 
be  accelerated  by  low  concentrations  of  the  mercurials,  as  noted  by  Hira- 
shima  (1934)  in  chick  fibroblast  cultures,  and  by  others.  This  has  also  been 
reported  for  plant  tissues,  p-MB  at  0.001-0.005  mM  stimulating  the  growth 
oi  Arena  coleoptiles  some  20-25%  (Thimann  and  Bonner,  1949).  It  is  inter- 
esting to  recall  that  Mallus  (1931)  gave  HgClg  at  0.25-0.3  mg/kg  to  atrophic 
but  otherwise  healthy  children  and  observed  certain  changes  —  increased 
chest  measurements,  rise  in  erythrocyte  count  and  hemoglobin,  and  eleva- 
tion of  urinary  nitrogen  —  indicating  a  stimulation  of  growth  and  metabo- 
lism. The  mechanisms  by  which  mercurials  can  stimulate  growth  are  un- 
known, but  it  is  possibly  not  a  specific  action  since  many  types  of  cells  ex- 
hibit an  increased  proliferative  activity  when  disturbed  slightly  by  irritant 
substances  (our  terminology  in  this  field  is  admittedly  inadequate). 

Inhibition  of  growth  is  invariable  when  the  mercurial  concentration  is 
increased  beyond  a  certain  level,  which  is  frequently  quite  low.  Many  of 
the  experiments  with  tissue  cultures  have  been  done  of  necessity  in  complex 
media  (e.g.,  embryo  extract)  which  must  bind  a  large  fraction  of  the  mer- 
curial, so  the  true  inhibitory  potency  must  be  much  greater  than  is  indi- 
cated by  the  concentrations  used.  Fibroblastic  and  leucocytic  migration  is 
depressed  50%  by  Hg++  at  concentrations  near  0.08  mM,  and  growth  is 
somewhat  modified  even  at  0.0037  mM  (Meier,  1933).  Chick  embryo  heart 
cultures  fail  to  grow  in  0.08  mM  Hg++  (Salle  and  Lazarus,  1936)  and  here 
the  organic  mercurials  are  less  toxic  (Salle,  1943).  Pulsations  of  the  cardiac 
cells  are  stopped  by  Hg++  before  growth  is  affected,  but  the  organic  mer- 
curials (thimerosal  and  nitromersol)  stop  growth  and  cause  cytolysis  with- 
out previously  interfering  with  the  contractile  activity.  The  concentration 
of  Hg++  for  50%  inhibition  of  Eagle's  KB  strain  of  human  carcinoma  cells 
is  0.037  mM  (Smith  et  al.,  1959),  and  75%  reduction  in  the  mitoses  of  mouse 
ear  epidermis  requires  0.01  mM  p-MB  (Gelfant,  1960).  We  may  thus  con- 
clude that  mercurials  at  concentrations  around  0.005-0.05  mM  seriously 


EFFECTS    ON    MITOSIS,   GROWTH,    DIFFERENTIATION  969 

restrict  growth  of  mammalian  cells  in  culture,  higher  concentrations  usually 
killing  the  cells  directly'.  If  the  mechanism  of  this  growth  inhibition  is  to  be 
elucidated  and  correlated  with  metabolic  alterations,  it  will  be  necessary 
to  work  within  this  range  if  the  conclusions  are  to  be  valid. 

Some  General  Aspects  of  the  Effects  of  Mercurials  on  Mitosis  and  Growth 

A  few  general  comments  on  the  localization  of  the  sites  of  growth  inhi- 
bition were  made  on  page  1-531  and  reference  to  these  will  make  it  evident 
that  at  the  present  time  we  have  little  hope  of  explaining  the  actions  of  the 
mercurials.  Essentially  nothing  is  known  of  the  possible  effects  of  mer- 
curials on  protein,  nucleic  acid,  or  coenzyme  synthesis,  or  whether  the 
demonstrated  inhibitions  of  active  transport  are  in  any  way  related  to  the 
growth  depression.  We  have  seen  that  respiration  is  not  significantly  de- 
pressed during  growth  inhibition  in  the  few  instances  in  which  it  has  been 
determined;  however,  in  view  of  the  rather  potent  inhibition  of  the  cycle, 
it  would  be  worthwhile  to  pursue  this  question  further.  Mercurials  are  not 
efficient  uncoupling  agents  and  could  scarcely  act  in  this  way.  The  signifi- 
cance as  a  possible  metabolic  mechanism  of  mercurial  action  of  the  obser- 
vation by  Hirashima  (1935),  that  glucose  reduces  the  toxicity  of  Hg++  for 
fibroblast  cultures,  cannot  be  evaluated. 

A  direct  action  on  the  sol-gel  transformations  and  protoplasmic  move- 
ments during  furrowing,  spindle  formation,  and  cleavage  has  been  postulat- 
ed and  discussed  briefly  in  previous  sections.  Mazia  (1959)  believes  that  the 
mitotic  apparatus  may  be  an  S — S  bonded  structure  because  of  the  ability 
of  agents  splitting  S — S  bonds  to  dissolve  the  structure,  and  states  that 
p-MB  and  mersalyl  bring  about  the  dissolution  of  the  freshly  isolated  spindle. 
The  question  is  whether  such  an  action  can  be  exerted  at  the  concentrations 
occurring  within  cells  during  mitotic  inhibition.  We  have  also  mentioned 
that  several  workers  favor  a  nuclear  site  for  the  mercurials  but  that  the 
evidence  is  insufficient,  as  is  that  of  Meyer  (1960),  who  showed  that  the 
conidia  of  Fusarium  decemcellnlare  incubated  with  0.037  mM  Hg++  accu- 
mulate mercury  in  some  form  either  on  or  within  the  nucleus,  since  there 
is  no  necessary  correlation  between  relative  intracellular  concentrations  and 
the  site  of  action.  It  is  worth  noting  that  p-MB  interferes  with  the  synthesis 
of  RNA  from  nucleosides,  as  determined  by  the  uptake  of  cytidine  into  the 
nuclear  RNA  of  HeLa  cells,  and  an  inhibition  of  the  RNA  polymerase  was 
suggested  (Srinivasan  ct  al.,  1964).  HeLa  cells  are  blocked  in  metaphase  by 
0.02  mM  7)-MB.  In  connection  with  the  selective  accumulation  of  mercurials, 
it  is  worth  noting  that  merbromin  appears  to  be  localized  in  tumor  tissues 
of  both  mouse  and  man,  as  indicated  by  fluorescence  several  days  after 
initiation  of  intravenous  or  oral  administration  (Katsuya  et  al.,  1963).  Al- 
though kidney  exhibits  the  highest  concentration  of  mercurial  initially,  it 
and  other  normal  tissues  lose  the  merbromin  much  faster  than  tumors.  It  is 


970  7.  MERCURIALS 

somewhat  surprising  that  tumors  could  contain  a  component  holding  mer- 
curials more  tightly  than  in  normal  tissues,  or  at  least  more  of  a  component 
binding  the  mercurial  strongly,  but  this  would  support  ideas  which  have 
been  advanced  relative  to  the  abnormal  state  of  SH  group-containing  sub- 
stances in  tumor  cells. 


EFFECTS  ON  THE  GROWTH   OF  MICROORGANISMS 

Mercurials  have  been  used  for  years  to  control  the  growth  of  many  types 
of  microorganism,  invertebrate,  and  plant,  and  have  been  applied  commer- 
cially as  fruit  sprays,  paint  preservatives,  mothproofers,  grain  insecticides, 
anthelmintics,  as  well  as  antiseptics  and  disinfectants,  and  in  antifouling, 
crab  grass  control,  bacterial  plant  diseases,  and  nematode  control.  One  of 
the  most  important  commercial  uses  at  the  present  is  as  fungicides  in  the 
treatment  of  seeds,  fruits,  and  plants,  and  for  the  most  part  certain  organic 
mercurials  have  been  developed  for  this  purpose.  The  clinical  use  of  mer- 
curials as  antiseptics,  first  popularized  by  Koch  (1881),  has  dechned  some- 
what due  to  the  discovery  of  generally  more  selective  antibacterial  agents, 
but  most  of  the  experimental  work  has  been  done  with  this  group  of  aro- 
matic mercurials.  Inasmuch  as  the  relation  between  these  actions  and  met- 


NO, 


\^      /^  Hg       NO3 

H3C 

Phenylmercuric  nitrate  Nitromersol 

(Merphenyl  nitrate,  Merphene)  (Metaphen) 


HgOH 


Merbromin  Thimerosal 

(Mercurochrome)  (Merthiolate) 

abolic  interference  is  vague,  only  a  cursory  treatment  of  the  mercurials 
as  inhibitors  of  microorganism  growth  will  be  given. 


EFFECTS    ON    THE    GROWTH    OF    MICROORGANISMS  971 

Bacteria,   Fungi,  and   Yeast 

Koch  originally  claimed  that  Hg++  possesses  the  ability  to  kill  various 
bacteria  and  their  spores,  but  it  was  soon  shown  by  Geppert  (1889)  that 
the  proliferative  activities  of  the  treated  bacteria  can  be  restored  by  remov- 
ing the  Hg++  with  sulfides,  and  that  the  action  of  the  Hg++  is  primarily 
bacteriostatic,  a  concept  confirmed  many  times  and  extended  to  other  or- 
ganisms. Dilution  or  washing  is  not  sufficient  to  extract  the  Hg++  bound 
to  the  cells  (Chick,  1908),  but  sulfides  and  various  thiols  (Fildes,  1940)  can 
readily  reverse  the  bacteriostatic  action.  With  high  enough  concentrations 
and  prolonged  exposure,  of  course,  bacteria  may  be  killed,  especially  at 
elevated  temperatures.  Yeast  cells  incubated  with  0.93  mM  Hg++  are  killed 
progressively  over  an  hour  as  determined  by  staining  with  methylene  blue 
or  Congo  red  (Rahn  and  Barnes,  1933).  One  of  the  first  effects  noted  at 
minimal  concentrations  of  the  mercurials  is  a  prolongation  of  the  lag  phase 
of  proliferation  (Cook  and  Steel,  1959).  Increasing  the  concentration  of  mer- 
curial progressively  delays  and  slows  the  growth,  and  eventually  stops  it, 
at  which  point  the  cells  can  remain  viable  often  for  quite  extended  periods 
of  time,  as  exhibited  by  the  renewal  of  proliferation  when  the  mercurial  is 
removed  with  a  thiol. 

The  sensitivities  of  various  microorganisms  to  the  mercurials  are  shown 
in  Table  7-26,  in  which  only  a  few  of  the  reported  results  have  been  pre- 
sented because  the  purpose  is  mainly  to  illustrate  that  proliferation  is 
usually  depressed  at  fairly  low  concentrations,  and  that  the  data  often  vary 
due  to  the  conditions  of  the  testing.  One  of  the  most  important  factors  is 
the  medium  used,  inasmuch  as  most  growth  media  contain  many  substances 
capable  of  complexing  with  the  mercurials  and  reducing  their  effective 
concentration.  Claus  (1956)  demonstrated  that  the  minimal  inhibitory  con- 
centration of  Hg++  varies  over  a  10-fold  range  depending  on  whether  pep- 
tone media  or  simple  nitrogen  sources  are  supplied,  and  Cook  and  Steel 
(1959)  also  presented  evidence  that  the  usual  culture  media  exert  a  pro- 
tective action.  Another  important  factor  is  the  temperature,  increase  in 
temperature  markedly  enhancing  the  bacteriostatic  effect  (Cianci,  1940; 
Cook  and  Steel,  1959),  the  Q^q  being  3-4  (Chick,  1908).  There  is  unquestion- 
ably species  variation  in  susceptibility  but  no  correlation  with  bacterial 
metabolic  or  growth  habits  has  been  made.  It  should  be  noted  that  very 
low  concentrations  of  the  mercurials  occasionally  stimulate  growth,  this 
being  perhaps  more  true  for  fungi  than  bacteria.  Robertson  (1943)  noted 
that  a  mercurial  used  for  preserving  leather  actually  accelerated  the  growth 
of  certain  fungi  within  a  particular  concentration  range,  and  Converse  and 
Besemer  (1959)  reported  that  p-MB  at  0.00028  mM  stimulates  the  growth 
of  Coccidioides  immitis,  although  spherulation  is  definitely  inhibited.  This 
phenomenon  would  probably  be  seen  more  commonly  if  low  concentrations 
of  the  mercurials  were  more  frequently  tested,  and  clinically  and  commer- 


972 


7.  MEECURIALS 


P5 


1^ 


>1 

Tl 

4> 

■t^ 

r, 

3 

W 

o 

o 

IC 

iCi 

lO 

?D 

^-^ 

^— s 

CO 

-* 

■* 

Tt< 

^■^ 

^-v 

Oi 

i> 

t- 

C5 

05 

Oi 

C5 

t- 

o 

t-H 

(M 

-* 

F-H 

f— ( 

f-H 

l-H 

-* 

fO 

■ 

05 

>-*^ 

05 

^-^ 

^^ 

■ 

^-^ 

05 

a> 

^^ 

-2 

:3^ 

■* 

C5 

to 

Tt< 

Cl. 

^ ^ 

-2 

0) 

C5 

^-^ 

^-^ 

liO 

Oi 

>o 

o 

-— ^ 

"C 

G 
H 

a 

a 

eS 

H 

C 

c 

^-^ 

lO 

■* 

a; 

c 

-* 

Oi 

Ci 

^ 

•* 

TtH 

OJ 

O 

^ 

(O 

05 

TfH 

3 

o 

■^ 

■ — ' 

o 

Si 

Tt< 

3 

C<5 

o 

3 

Oi 

CO 

a 

Oi 

' — 

'o 

' — ' 

O 

05 

05 

o 
O 

Eh 

o 

2 

4^ 

C 

'O 

c 
ft 

O 
O 

+2 

-2 

75 

5 

CO 

CO 

C 

a 

.2 
■-3 

> 

o3 

00 
o 

C5 

C 

CO 

C 
4) 

T3 
C 

CO 

C 

c6 

8 

"3 

C 

13 
> 
o 

o 
o 

> 

O 

g 

s 

2 

0) 
4) 

> 

u 
0) 

> 

4) 

S 

o 

r2 

hlH 

o 

^ 

m 

O 

^ 

1-:! 

O 

O 

f-l 

H 

§    <5 

m 

O 

OJ 

m 

02 

CQ 

H 

02 

OOSOOO'^OOOOO^O-^OO^^—HOiOtO 
OOeMOOOOOOOOO-HOOOOO-navJ 

-hOo— <ooooooooooooooodd 


<^mmmmmmmmmmxnmifimmvi^j^^mm 


P5  +    +         + 

S     be    M  g     M 
4  K    W    Ph    K 


CO  +      +  +      +      +      + 

5^+    +    ^+    +    +    + 


s 

CO 

s 

"S 

3' 

;«^ 

00 

5 

i< 

00 
•<s» 

.'°° 

"sS 

I1 

.O 

^ 

s 

CO 

S 

s 

.2 

=0 

■« 

•^ 

'C 

5^ 

s^ 

■^   ^ 

t) 
e 

o 

o 

SU       CO 

c« 

oq 

O 

O 

q    Et] 

?  s  £  s 


s    g    e    e    e    e 

»    2  ~  ~  ~  ss 
g    ^    (i)    ^    ^ 


"S  "o    <i>  ~:  -S  r^  -S 
e    *~  -=0    «    e    «    B 

1^    Q^    ft,    (^2    G5    CQ    CQ 


^ 


EFFECTS  ON  THE  GROWTH  OF  MICROORGANISMS 


973 


t^   r^   t^    ir^ 


_i^  c5      cd      c^      cd    Ci 

o  .t^  .ti  .-ti  .-S  S 

-H  ^^     M     M     M    M   ^^ 


c3      cS      c3      c5      ^    ^- 


o3    Tj< 


3;  :5h    c    c    c    c 


(^      c^      c3      c^      c3 


2  -^  >^  >^  >H  >H 


O     -H 


cS      c6     o3     c3 


^   -C   T3 
»     c3     S 


^    _S    .5      C      C 


c3     cS    ^    SS 


ffO 

■* 

CO 

10 

01 

(M 

CO 

C5 

0 

0 

0 

CO   0 

0 

f^i 

H      0 

0 

0 

0 

rt   0 

0 

0 

000000000 


00202:000020203 


ffi 


CQ 


O^  Gi  O^ 
iO  »0  L^ 

Ci  Ci  Oi 


4) 

« 

0) 

■    ' 

■    ' 

a 

s 

s 

^ 

0) 

0) 

05 

P5 

T3 
e8 

c 

s 

c3 

0 

0 

tj 

-0 

C5 

C5 

e3 

ca 

g 

ki 

1 

i 

cS 

cS 

-a 

CM  <M  ^  (M  (N  M  5^1 

CD  CO  C^  O  O  CO  CO 
Oi  O  CO  CI  C5  05  C5 


"         '        CS      O  CJ  O  O  05 

I  S  tf  S  2  ^ 

c3    'C  TJ  "C  "C  ~ 

03    c  c  c  c  e 

^  ^  ci  ci  ,«^ 


«  -^   -^   -r  -r 


(B    <D    aj     <u 

M    W)    br    M 
'C   "C   "O   TS 


OOOc3c3c3c30 

OOOHHPLiPhH 


cfi      e3      Hi      cS 

Pli   Ph   P^   Ph 


00 

O    <M  I> 

o  o  00        o 

O    O    <M    O    O 

00000 


05     00     »C     05     00     05     00 

00   -^   00   00  ■*   00   ■* 
O'    O    -H    o    o   o    o 


000000 
V 


000000 


O2O2O20Ca2a2O202a2CZ202 


pq 


S    5 


<^  ^ 


o    o 


m 


CQ 


02   e 


S  to 


O 


73 

-rs 

_!.; 

JS 

■t-i 

ii 

is 

^ 

H 

0 

d 

be 

3 

,__, 

C 

c 

e3 

^ 

A 

■4J 

0 

-0 

S 
« 

0 
to 

3 
03 

0 

^ 

4) 

•w 

+i 

X 

3 

H 

0 

2    c    o 


974  7.  MERCURIALS 

cially  it  might  be  an  important  aspect  of  their  use.  Despite  the  many  com- 
parisons of  the  relative  activities  of  different  mercurials,  very  few  interesting 
correlations  between  structure  and  effectiveness  have  emerged.  Krahe  (1924) 
found  that  the  bacteriostatic  activity  of  Hg++  is  reduced  by  increasing  the 
concentration  of  NaCl  and  postulated  this  to  be  due  to  the  formation  of 
HgClg""  and  HgCl4=  complexes,  these  being  less  lipid-soluble,  and  showed 
that  the  distribution  coefficient  between  ether  and  water  is  reduced  parallel 
with  the  antibacterial  potency.  Coleman  et  al.  (1937)  observed  in  the  ali- 
phatic mercurials  that  the  antibacterial  activity  increases  with  the  length 
of  the  side  chain.  Such  relationships  have  been  found  for  many  actions  of 
the  mercurials  and  are  probably  based  on  differences  in  penetration  into 
the  cells  rather  than  to  fundamental  differences  in  action  on  the  susceptible 
cellular  mechanisms. 

The  uptake  and  distribution  of  mercurials  have  been  well  studied  and 
several  facts  relevant  to  the  mechanism  of  their  action  have  emerged.  The 
amount  taken  up  in  any  case  will  depend  on  the  relative  quantities  of  cells 
and  mercurial  present.  Herzog  and  Betzel  (1911)  incubated  10  g  of  pressed 
yeast  (2.6  g  dry  weight)  with  various  concentrations  of  Hg++  and  found 
the  cellular  Hg++  concentration  to  increase  with  the  total  amount  of  H.g++ 
present,  but  the  percentage  taken  up  falls  (see  accompanying  tabulation). 


Total  Hg++  present 
(g/100  ml) 

Hg++ 

taken   up    by   yeast   cells 

Total 

(g) 

Concentration 
(g/g  dry  weight) 

0/ 

/o 

0.092 

0.077 

0.030 

84 

0.460 

0.168 

0.065 

37 

0.921 

0.219 

0.084 

24 

1.341 

0.304 

0.117 

16 

3.683 

0.449 

0.173 

12 

Since  there  are  roughly  10^''  cells  in  1  g  of  dry  yeast,  these  uptakes  would 
correspond  to  between  2  x  10*  and  10®  Hg++  ions/cell;  they  would  also 
correspond  to  0.065-0.38  g  Hg++/g  yeast  protein  of  molecular  weight  100,000. 
The  amount  of  Hg++  accumulated  by  yeast  is  thus  very  considerable  and 
only  a  fraction  is  likely  to  be  bound  to  SH  groups.  E.  coli  binds  even  more 
Hg++,  since  Hahn  and  Remy  (1922)  found  an  uptake  of  around  0.5  g  Hg++/g 
dry  weight  (assuming  around  75%  water  content)  from  a  3.7  mM  solution. 
McCalla  and  Foltz  (1941)  calculated  that  E.  coli  possesses  around  10*  bind- 
ing sites/cell,  which  is  close  to  the  figure  to  be  estimated  from  the  uptake 
found  by  Hahn  and  Remy.  Steel  (1960)  claimed  that  a  cell  of  E.  coli  con- 
tains about  10*  SH  groups,  but  that  Hg++  does  not  react  with  all  of  them, 


EFFECTS    ON    THE    GROWTH    OF    MICROORGANISMS  975 

p-MB  giving  a  much  higher  accumulation.  It  is  more  difficult  to  determine 
the  distribution  of  Hg++  within  the  cells.  Siipfle  (1923)  treated  anthrax 
bacilli  with  Hg++  and  then  with  HgS,  and  showed  black  granules  within  the 
cells.  However,  the  Hg++  enters  much  more  slowly  than  the  antibacterial 
action  develops,  which  might  be  used  as  evidence  for  a  primary  membrane 
site.  Ruska  (1947)  examined  Hg++-treated  streptococci  and  E.  coli  with 
the  electron  microscope  and  observed  mercury  in  the  membrane  and  dif- 
fusely distributed  in  the  cytoplasm,  but  most  in  small  globular  masses  be- 
tween the  membrane  and  the  cytoplasm.  By  a  similar  technique,  Harris 
et  al.  (1954)  found  no  mercury  in  the  membrane  or  cell  wall  of  E.  coli  — 
confirming  the  absence  of  electrophoretic  change  in  cells  treated  with  Hg++ 
—  and  most  deposited  as  granules  within  the  cytoplasm.  Troger  (1959) 
localized  mercury  by  the  diphenylcarbazone  method  and  found  accumula- 
tion in  certain  areas.  It  should  be  noted  that  visualization  either  electron 
microscopically  or  histochemically  is  difficult  in  bacteria  and,  furthermore, 
that  generally  quite  high  concentrations  have  been  used  so  that  the  pattern 
of  distribution  cannot  apply  directly  to  the  bacteriostatic  situation. 

The  possible  mechanisms  for  the  bacteriostatic  action  of  the  mercurials 
have  been  debated  for  years  and  many  theories  have  been  proposed  without 
benefit  of  experimental  evidence.  The  amount  of  valuable  work  on  the  mer- 
curials from  the  standpoint  of  basic  actions  is  almost  negligible,  due  prob- 
ably to  the  fact  that  when  bacterial  metabolism  and  proliferation  began 
to  be  investigated  seriously,  attention  was  turned  to  the  sulfonamides  and 
antibiotics.  From  Tables  7-13  and  7-17  one  might  conclude  that  metabolism 
must  certainly  be  depressed  in  some  manner  during  the  action  of  the  mer- 
curials on  bacteria,  and  this  may  well  be  in  many  cases,  but  Yamada  and 
Yanagita  (1957)  showed  quite  conclusively  that  the  growth  of  staphylococci 
is  140  times  and  57  times  more  sensitive  than  respiration  to  thimerosal  and 
Hg++,  respectively.  Indeed,  it  is  possible  to  stop  growth  essentially  com- 
pletely without  affecting  respiration  significantly.  Despite  the  lack  of  critical 
experiments  in  other  organisms,  it  is  safe  to  say  that  the  mercurials  do  not 
inhibit  the  growth  of  microorganisms  by  simply  suppressing  oxidative  proc- 
esses and  the  supply  of  energy  for  growth  and  division.  Fildes  (1940)  on 
the  basis  of  irrelevant  evidence  concluded  that  the  antibacterial  mechanism 
is  based  on  reaction  with  SH  groups,  and  the  impression  is  gained  that  he 
was  thinking  of  the  smaller  thiols  rather  than  proteins  and  enzymes.  Lou- 
reiro  and  Lito  (1946)  put  this  theory  on  a  better  basis  by  demonstrating 
some  correlation  between  the  fraction  of  bacterial  SH  groups  reacted  and 
the  bactericidal  activity.  However,  even  here  all  one  can  do  is  to  increase 
the  mercurial  concentration  so  that  more  and  more  SH  groups  are  reacted, 
and  it  is  not  surprising  that  more  and  more  cells  are  inhibited  or  killed; 
one  cannot  say  what  fraction  of  SH  groups  should  be  reacted  before  an 
effect  on  the  bacteria  is  observed,  and  indeed  it  is  very  unlikely  that  it 


976  7,  MERCURIALS 

requires  a  1  :  1  ratio  of  mercurial  to  SH  groups,  as  assumed  by  these  work- 
ers. Nevertheless,  in  the  face  of  no  negative  evidence,  it  is  felt  that,  the 
mercurials  do  inhibit  growth  by  reacting  with  some  SH  groups  —  the  prob- 
lem is  with  what  SH  groups,  since  there  are  many  different  SH-containing 
substances  in  the  cell.  Do  the  mercurials  inactivate  some  SH  enzyme,  or 
enzymes,  involved  in  an  important  metabolic  pathway,  or  react  with  SH 
groups  in  the  membrane  to  block  active  transport  of  necessary  substances 
into  the  cell,  or  alter  permeability  so  that  intracellular  components  are  lost, 
or  directly  stabilize  the  membrane  to  prevent  division,  or  interfere  with  the 
utilization  of  ATP,  or  disturb  metabolism  by  reacting  with  some  thiol  co- 
enzyme ?  Experiments  showing  that  certain  substances  protect  against  mer- 
curials are  not  easy  to  interpret.  Thus  Pershin  and  Shcherbakova  (1958) 
found  that  histidine,  glutamate,  methionine,  and  particularly  thiamine  pro- 
tect E.  coli  against  Hg++,  and  interpreted  this  as  indicating  that  the  metab- 
olism of  these  compounds  is  interfered  with  by  Hg++,  but  it  is  also  possible 
that  the  protection  is  simply  due  to  complexes  formed  with  the  Hg++. 
Theories  involving  various  physicochemical  properties  of  Hg++  and  other 
heavy  metal  ions  —  such  as  solution  pressure,  solubility  products,  electro- 
negativity, and  ionization  potential  (e.g.,  Shaw,  1954;  Somers,  1959)  — 
do  not  warrant  serious  consideration  since  they  simplify  the  biological 
system  beyond  recognition  and,  even  if  true,  would  not  help  us  appreciably 
to  understand  how  the  mercurials  act. 

Viruses 

Most  viruses  and  phages  can  be  inactivated  by  the  mercurials  but  it 
requires  fairly  high  concentrations  relative  to  those  inhibiting  bacterial 
growth  (Table  7-27).  In  most  work  a  virus  suspension  is  incubated  with  the 
mercurial  for  a  certain  period  and  the  infectivity  is  then  tested.  As  with 
the  effects  on  microorganisms  in  general,  the  degree  of  inactivation  by  the 
mercurials  depends  on  many  factors,  particularly  the  medium  in  which  the 
virus  is  suspended,  the  temperature,  and  the  exposure  time.  The  rates  of 
inactivation  are  quite  different  for  various  viruses:  ECHO  7  virus  is  50% 
inactivated  by  0.05  roM  p-MB  in  1  min  and  99%  inactivated  in  6  min 
(Choppin  and  Philipson,  1961),  whereas  tobacco  mosaic  virus  is  not  com- 
pletely inactivated  after  24  hr  exposure  to  18.5  mM  Hg++  (Kassanis  and 
Kleczkowski,  1944).  This  would  be  expected  since  the  virus  SH  groups 
must  vary  widely  in  reactivity  as  do  the  SH  groups  of  proteins  in  general. 
The  inactivation  is  first  order  with  respect  to  virus.  Staphylococcus  phage 
infectivity  declines  exponentially  when  exposed  to  Hg++  according  to  the 
equation: 

dPIdt  -  kiRgCU)  (Po  -  Pi) 

where  Pq  is  the  phage  initially  present  and  P,  the  inactivated  phage  (Krueger 
and  Baldwin,  1933).  This  equation  holds  fairly  well  over  most  of  the  range. 


EFFECTS    ON    THE    GROWTH    OF    MICROORGANISMS 


977 


o     2     2 

m     —     — 

-(3         ~         ~ 

See 


X.       C-- 


W      ^  ^ 


=  C5      C5 


1^ 


> 


02 


D5 


& 

>.. 

^ 

o- 

bti 

hf 

o 

o 

D 

3 

:::) 

§ 

OJ 

V3 

02    D5 


5    ;=      :o 


-H         O  C         -H         C 


Z  V 


m    m    pq 


'> 

o 

c 
o 

^ 

o 

c 
o 

«3 
!»1 

o 

O 

< 

M 

O 

O 

H 


«) 


^     ^5     S     S     ^  .2     Mp^ffl     ►^ 


P>q 


W 


w 


j2 

C 

C 

O 

c 

o 

00 

b     <  <  ^ 


t^     rt  ^  rt 

M        O     O     O 


+   m 


978 


7.  MERCURIALS 


^ 


o    .2 


S   P^ 


5    >-- 


'-^     ^      C3       >0 


CO       00       -5**       C5 


fP        Ml         03       Ci       -H 


e    ^     cs      ^ 


■5      •?      <4i 


O    cc      M    tsi    PM    O    O 


3     ^ 


e 

HH 

e 

.^^ 

-c 

■*^ 

ii) 

s 

■4) 

f^ 

C 

o 

o 

t« 

^ 

m 

o 

<: 

H 

< 

c 
o 
o 


d    o 


t-H  O    — I  o    ^ 


-H  -H  O       -^ 


S    o 


S     M    52 


CO 

O 

00 

>o 

LO 

Tt( 

c^ 

II 

M 

11 

II 

-' 

(M 

II 

P5 

H 

M 

ll 

H 

H 

H 

D5 
H 

P5 
H 

P5  p? 


HHh5      h^hShS^h5     -2     ,2^     -2^     S,    '-'     2.      i-H-2,2 


t  s 


4      W    Ph    ffi 


fa, 


P3 


P5 


PP 
a. 


S      h^ 


a 

o 

\< 


EFFECTS    ON     THE    GROWTH    OF    MICROORGANISMS 


979 


^^  GO       00 


'^i    .- 


■^       rC    x; 


W 


^    Ci    -- 
^    S     >. 


TS    e 


c   55   2    ^ 


PC    pq 


d    =2  ^ 


c    c 


W      '^  ^  < 


.  >     o     o     o 
^      a:      K      oQ 

O    *<    "^    "^ 


rt    o  d       —I 


00    cc    CO       >n) 

O'    o    o       00 


o   —I        — 


O 

o 

O 

g 

O^ 

o^ 

^o 

o 

o 

o 

o 

05 

f-H 

o  ■ 

o 

a 

'2 

C5 

■* 

O 

c 

(M 

^ 

d 

II 

c 

.2 

11 

c 

^ 

II 

c 

S 

c 
o 

II 

c 
o 

II 

c 
.2 

V 

II 

2 

to 

GO 

so 

to 

O 

II 

II 

11 

'-i3 
c3 

o 

'-2 

II 

'-3 

'-13 

'-H 

03 

'-D 

c3 

_2 

'-3 

'-+3 
c3 

II 

II 

II 

II 

II 

K 

P2 

P5 

> 

c3 

_> 

P5 

_> 

_> 

_> 

> 

o 

a 

_> 

05 

K 

05 

P5 

P5 

H 

H 

H 

o 

'43 

o 

H 

-<;> 

o 

'•*3 

-^ 

'■^ 

o 

H 

H 

H 

H 

H 

M 

M 

M 

3 

O 

c3 

M 

C 

a 

cS 

S 

<v 

o 

ei 

M 

M 

M 

M 

ti> 

_o 

_o 

_o 

1^ 

^3 

^ 

c 

C 

»5 

^ 

Z 

C 

_2 

^ 

^ 

^ 

_o 

m 

m 

+ 

S 

S 

+ 

C5, 

4 

K 

CQ 


pa 


pq 


biD 


PQ 


O 
'>, 
ft 


CO       ^j 


p-i    fii 


PM 


980  7.  MEKCURIALS 

but  there  appears  to  be  a  small  resistant  fraction  which  remains  infectious 
over  several  days  (Krueger  and  Baldwin,  1934).  Similar  kinetics  of  inactiva- 
tion  have  been  reported  by  Moriyama  and  Ohashi  (1941)  for  E.  coli  phage, 
and  by  Allison  (1962)  for  fowl  plague  and  vaccinia  viruses. 

The  mercurials  usually  do  not  destroy  the  viruses  or  produce  irreversible 
structural  changes  in  them,  since  reactivation  with  thiols  has  been  observed 
with  staphylococcus  phage  (Wahl,  1939),  influenza  virus  (Klein  et  at.,  1948; 
Perez  et  al.,  1949),  psittacosis  virus  (Burney  and  Golub,  1948),  vaccinia 
virus  (Kaplan,  1959;  Allison,  1962),  streptococcal  phage  (Kessler  and 
Krause,  1963),  and  various  enteroviruses  (Choppin  and  Philipson,  1961).  The 
results  of  Krueger  and  Baldwin  (1933,  1934)  are  particularly  impressive; 
reactivation  with  sulfide  occurred  even  after  exposure  of  phage  to  around 
100  mM  Hg++  for  9  days  at  22o.  It  is  remarkable  that  viruses,  like  certain 
enzymes,  can  be  reactivated  readily  with  thiols  (especially  dimercaprol) 
although  no  reactivation  occurs  by  washing  or  dilution.  Dimercaprol  is 
able  to  reactivate  influenza  virus  in  vivo  when  injected  after  the  treated 
virus  in  animals  (Klein  et  al.,  1948)  or  chick  embryos  (Perez  et  al.,  1949). 

The  reactivation  with  thiols  does  not  prove  that  the  mercurials  react 
with  virus  SH  groups,  as  has  been  concluded,  but  there  is  evidence  for  the 
importance  of  SH  groups.  The  relative  resistance  of  tobacco  mosaic  virus 
to  the  mercurials  is  probably  due  to  the  unavailability  of  the  SH  groups, 
since  Anson  and  Stanley  (1941)  showed  that  p-MB  reacts  with  all  the  SH 
groups  of  denatured  virus  but  does  not  inactivate  native  virus.  Some  steric 
factor  preventing  reaction  with  p-MB  was  postulated  by  Fraenkel-Conrat 
(1959),  since  MM  reacts  stoichiometrically  in  a  1:1  ratio  with  the  SH 
groups.  The  restriction  may  be  imposed  by  hydrogen  bonding  to  adjacent 
groups.  Some  plant  viruses  are  structurally  altered  by  mercurials.  Solutions 
of  potato  virus  X  lose  their  flow  birefringence  when  treated  with  p-MB,  and 
sedimentation  studies  indicate  disintegration  into  subunits  (Reichmann  and 
Hatt,  1961).  It  was  concluded  that  the  SH  groups  occur  near  the  linkage 
sites  holding  the  units  together,  rather  than  participating  in  the  linkage, 
and  that  the  bulky  p-MB  molecule  splits  the  links.  Turnip  yellow  mosaic 
virus  is  also  split  into  subunits  by  p-MB,  and  RNA  is  liberated  simultane- 
ously (Kaper  and  Houwing,  1962  a).  The  artificial  top  component  (empty 
virus  protein  shells)  binds  645-660  molecules  of  mercurial  per  particle.  As 
structural  changes  occur,  new  SH  groups  are  unmasked  and  react  with 
p-MB  (Kaper  and  Houwing,  1962  b).  Finally,  one  must  consider  the  reac- 
tion of  mercurials  with  the  nucleic  acid  components  of  the  viruses,  since 
such  complexes  have  been  established  (Katz,  1962).  Tobacco  mosaic  virus 
RNA  complexes  with  Hg++  (Katz  and  Santilli,  1962  b)  but  much  of  the 
infectivity  remains  in  this  case,  although  retention  of  specific  infectivity 
was  not  demonstrated  (Katz  and  Santilli,  1962  a). 

We  shall  now  inquire  into  the  particular  phases  of  virus  multiplication 


EFFECTS    ON    THE    GROWTH    OF    MICROORGANISMS  981 

inhibited  by  the  mercurials.  It  is  clear  that  there  is  little  or  no  selective  ac- 
tion on  viruses  growing  in  vivo,  and  the  mercurials  have  not  been  found  to 
be  effective  virucidal  or  virustatic  agents.  Thus,  although  p-MB  depresses 
psittacosis  virus  formation  in  chick  embryo  cultures,  it  also  inhibits  tissue 
growth,  and  it  is  quite  possible  that  the  effect  on  the  virus  is  secondary  to 
that  on  the  host  cells  (Burney  and  Golub,  1948).  Corneal  infections  with 
herpes  virus  are  not  benefited  by  application  of  p-MB,  although  the  virus 
is  readily  inactivated  in  vitro  (Sery  and  Furgiuele,  1961),  and  mercurials 
are  not  effective  in  preventing  or  treating  plant  virus  infections.  Kaplan 
(1959)  concluded  that  mercurials  react  with  SH  groups  on  the  surface  of  the 
virus  and  thus  prevent  attachment  to  the  host  cell.  Certainly  the  hemagglu- 
tinating  activity  and  the  adsorption  onto  erythrocytes  are  depressed  along 
with  the  infectivity  (Choppin  and  Philipson,  1961).  Adsorption  of  entero- 
viruses onto  renal  cells  may  also  be  reduced,  but  p-MB  does  not  prevent 
adsorption  of  influenza  virus  onto  the  chorioallantoic  membrane.  Further- 
more, p-MB  does  not  prevent  infection  of  E.  coli  by  phage  but  inhibits  the 
proliferation.  Allison  (1962)  holds  that  the  mercurials  do  not  affect  the 
primary  attachment  of  the  virus  to  the  host  cell,  but  may  prevent  the  un- 
coating  of  the  virus,  an  event  which  precedes  multiplication.  The  studies 
of  Shug  et  al.  (1960)  on  T2  phage  show  that  here  the  inhibition  by  p-MB  is 
exerted  early  in  the  development,  the  maximal  inhibition  being  10-20  min 
after  infection.  This  excludes  the  energy-yielding  host  metabolism  as  a  pri- 
mary site  of  action,  since  the  energy  requirements  are  greater  after  25  min 
when  the  phage  is  being  synthesized  and  assembled.  They  conclude  that 
the  site  of  attack  is  a  protein  concerned  with  the  initial  phases  of  replica- 
tion and  possibly  involved  in  the  assembly  of  the  components  into  an  in- 
tact phage.  It  would  not  be  surprising  if  the  site  of  mercurial  action,  or 
the  phase  disturbed,  is  different  for  the  various  phages  and  viruses. 

Protozoa 

Ciliates  are  immobilized  and  killed  by  the  mercurials  but  the  results  re- 
ported are  quantitatively  discrepant.  Nuhaus  (1910)  found  that  0.15  milf 
Hg++  paralyzes  paramecia  in  50  min  and  kills  them  in  70  min,  but  Wood- 
ruff and  Bunzel  (1910)  stated  that  0.175  mM  Hg++  stops  all  motion  within 
2  sec.  Even  greater  sensitivity  was  reported  by  Gause  (1933),  as  indicated 
in  the  accompanying  tabulation.  It  should  be  noted  that  actually  death 


Hg++  (milf)  Duration  of  life  (sec) 


0.01 

1080 

0.015 

360 

0.02 

116 

0.03 

30 

982  7.  MERCURIALS 

was  not  the  criterion,  but  cessation  of  movement,  since  no  attempt  to  reac- 
tivate was  made.  When  the  log  of  survival  time  was  plotted  against  log 
(Hg++),  two  linear  segments  were  obtained,  which  led  Gause  to  conclude 
that  two  different  processes  are  responsible  for  the  paralysis.  The  treat- 
ment is  based  on  an  equation  of  the  type  1-12-89  and  plotting  as  in  Fig.  I- 
12-38.  A  break  in  the  curve  would  suggest  a  change  of  the  exponent  n  and, 
since  it  is  possible  for  one  process  or  response  to  exhibit  different  values  of 
n,  it  is  not  necessary  to  conclude  that  Hg++  kills  by  two  processes.  High 
sensitivity  of  paramecia  to  Hg++  was  also  noted  by  Calcutt  (1950),  who 
found  0.001  raM  to  paralyze  within  6-14  min  depending  on  illumination. 
Paralysis  of  Colpidium  colpoda  occurs  in  3  min  after  exposure  to  0.00087 
mM  PM  (Walker,  1928).  Some  of  the  variations  in  sensitivity  are  due  to 
the  strains  used,  and  probably  some  to  the  media  in  which  the  ciliates  were 
suspended.  It  is  probably  safe  to  say  that  ciliary  movement  is  very  sensitive 
to  mercurials  and  is  stopped  within  a  few  minutes  by  concentrations  in  the 
range  0.001-0.01  mM.  Reversal  of  ciliary  beat  by  Hg++  is  not  observed 
(Oliphant,  1942).  The  classic  death  rate  curves  for  Colpidium  exposed  to 
0.2  mM  Hg++  and  their  relation  to  population  variation  were  discussed 
previously  (page  1-593  and  Fig.  1-12-36). 

The  effects  of  the  mercurials  on  ameboid  movement  are  interesting  be- 
cause of  the  possible  bearing  on  muscle  contractility.  Reznikoff  (1926)  used 
rather  high  concentrations  of  Hg++  and  consequently  found  only  a  pinch- 
ing off  of  the  region  into  which  the  Hg++  was  injected  (0.62  mM),  or  a 
break  in  the  membrane  (up  to  10  mM),  or  an  immediate  gelation  or  coagu- 
lation of  the  protoplasm  (up  to  200  mM).  Kappner  (1961)  used  mersalyl 
since  this  mercurial  has  been  a  favorite  with  myologists,  and  the  changes 
he  observed  in  amebas  with  increasing  concentration  are  worth  describing 
briefly.  With  0.001  mM  mersalyl  there  are  no  immediate  changes  but  after 
several  days  some  damage  is  evident.  At  0.01  mM  there  is  restriction  of 
normal  pseudopodial  response  and  fewer  pseudopods  are  formed,  while 
clumping  of  the  cytoplasmic  crystals  occurs.  At  0.1  mM  the  pseudopods 
withdraw,  the  ceUs  soon  form  numerous  small  pseudopods  at  the  end  of 
which  appear  tiny  spheres,  the  cortex  appears  to  be  thicker,  and  eventually 
the  cells  round  up.  At  1  mM  the  response  is  not  so  specific,  the  surface 
bubbles,  the  cells  round  up,  and  soon  the  membrane  dissolves.  Higher  con- 
centrations produce  vacuolization  and  cytolysis.  Abe  (1963)  reported  a  sim- 
ilar study  but  with  p-MB  to  which  amebas  seem  to  be  more  sensitive  than 
to  mersalyl,  since  0.1  mM  causes  cytolysis  within  8  min.  Further  investiga- 
tion of  this  interesting  problem  is  warranted  and  a  closer  analysis  of  the 
effects  of  low  concentrations  on  the  sol-gel  transformation  might  provide 
useful  information  on  the  role  of  SH  groups  in  protoplasmic  movement. 


DEVELOPMENT    OF    KESISTANCE    TO    MERCURIALS 


983 


DEVELOPMENT  OF  RESISTANCE  TO   MERCURIALS 

Most  types  of  microorganism  appear  to  be  able  to  adapt  to  the  presence 
of  mercurials,  but  usually  not  as  readily  or  to  such  a  degree  as  to  arseni- 
cals,  sulfonamides,  or  antibiotics.  Some  resistance  factors  are  given  in  the 
accompanying  tabulation,  but  it  is  likely  that  greater  tolerance  could  have 


Organism 


Mercurial 


Resistance 
factor 


Reference 


Staphylococcus  aureus 

Hg++ 

>50 

Benigno   and   Santi    (1946) 

1.8 

Klimek  et  al.  (1948) 

Escherichia  coli 

PM 

2.6 

Akiba  and  Ishii  (1952) 

Salmonella  pullorum 

Hg++ 

12 

Severens  and  Tanner  (1945) 

Salmonella  typhosa 

Hg++ 

6 

Severens  and  Tanner  (1945) 

Penicillium  notatum 

Hg++ 

2.5 

Partridge  and  Rich   (1962) 

Sclerotinia  fructicola 

Hg++ 

2.5 

Partridge  and  Rich   (1962) 

Yeast 

Hg++ 

>10 

Imshenetsky  and  Perova  (1957) 

Candida,  utilis 

Hg++ 

7 

Gerardin  and  Kayser  (1959) 

Treponema  pallidum 

Hg++ 

75 

Nogiichi  and  Akatsu   (1917) 

been  developed  in  some  instances  if  training  had  been  prolonged.  There  are 
also  naturally  occurring  resistant  strains  and  species.  An  interesting  exam- 
ple is  the  relative  tolerance  of  Penicillium  roqueforti  to  PM.  and  this  has 
bearing  on  the  preservation  of  groundwood  pulp  (Russell,  1955).  Most  fungi 
fail  to  grow  in  0.006-0.030  mM  PM,  but  this  species  grows  well  in  a  con- 
centration of  0.06  mM  and  furthermore  accumulates  sufficient  mercurial 
to  allow  the  less  resistant  organisms  to  grow.  The  number  of  serial  cultures 
in  increasing  mercurial  concentrations  required  to  produce  tolerance  varies 
with  the  organism:  It  was  20  transfers  for  the  fungi  in  the  above  table 
(Partridge  and  Rich,  1962),  70-100  transfers  for  the  species  of  Salmonella 
(Severens  and  Tanner,  1945),  and  up  to  500  transfers  for  yeast  (Imshenetsky 
and  Perova,  1957).  Occasionally  no  transfers  are  required  and  the  organism 
begins  to  grow  normally  after  a  prolonged  lag  period,  as  is  the  case  with 
Aspergillus  glaucus  where  hyphal  inoculations  fail  to  grow  for  periods  of 
up  to  14  days  in  0.033  mM  Hg++,  and  then  proliferate  without  loss  of  vigor 
(Briault,  1956).  This  indicates  that  resistance  can  develop  in  nonprolif crat- 
ing organisms. 

Inasmuch  as  we  do  not  understand  how  the  mercurials  depress  growth, 
it  is  clear  that  we  cannot  immediately  postulate  logical  mechanisms  for  the 
developed  resistance.  However,  some  interesting  observations  may  contrib- 
ute to  the  elucidation  of  the  mechanisms  of  inhibition.  The  resistance  is 
apparently  not  due  to  a  reduction  of  permeability  to  the  mercurials,  as 


984 


7.  MERCURIALS 


occurs  with  the  arsenicals,  since  Benigno  and  Santi  (1946)  found  that 
staphylococci  tolerant  to  Hg++  grow  when  they  have  taken  up  much  more 
Hg++  than  is  required  to  prevent  growth  of  the  normal  strain.  The  situation 
with  respect  to  the  thiol  content  of  the  resistant  organisms  is  confused, 
since  Akiba  and  Ishii  (1952)  showed  that  E.  coli  tolerant  to  PM  have  less 
SH  groups  than  normally,  and  Gerardin  and  Kayser  (1959)  found  that  tol- 
erant Candida  utilis  contained  6  times  more  SH  groups  than  the  normal 
strain.  Zambonelli  (1958  a,b)  has  obtained  evidence  that  adapted  yeast 
produces  more  HgS  and  believed  that  this  inactivates  much  of  the  Hg++. 
Normal  yeast  produces  HgS  only  from  sulfite,  whereas  adapted  strains  pro- 
duce it  from  sulfate  and  hyposulfite  in  addition,  although  not  from  cysteine 
or  glutathione.  If  the  resistant  strains  are  grown  with  only  cysteine  or  glu- 
tathione as  the  source  of  sulfur,  Hg++  readily  inhibits  their  growth.  If 
metabolic  changes  occur  during  adaptation,  they  are  not  marked.  Thus 
resistant  Candida  respires  normally  (Gerardin  and  Kayser,  1959)  and  re- 
sistant yeast  ferments  glucose  at  the  normal  rate,  although  the  respiration 
may  be  slightly  higher  (Imshenetsky  and  Perova,  1957).  Claus  (1956)  has 
shown  that  the  respiration  of  Aerobacter  aerogenes  is  initially  depressed  by 
Hg++  but  recovers  after  several  hours  and  reaches  normal  levels  (Fig.  7-46). 


Fig.  7-46.  Effects  of  Hg++  on  the  respiration  of  Aerobacter 
aerogenes.  The  O2  uptake  is  given  as  mm^  Oj/ml/SO  min.  A: 
Initial  exposure  to  Hg++;  B:  inoculation  of  organisms  from 
3  in  A  and  re-exposure.  C,  control;  1,  0.0011  mM; 
2,  0.0022  mM;  3,  0.0044  mM;  4,  0.0088  mM;  5,  0.0178 
mM;      6,  0.037  mM.   (From  Claus,  1956.) 


Adaptation  to  mercurials  is  apparently  specific  in  most  cases,  since  Se- 
verens  and  Tanner  (1945)  found  that  Salmonella  sp.  tolerant  to  Hg++  are 
not  tolerant  to  Cu++,  and  vice  versa,  while  Launoy  and  Levaditi  (1913) 
showed  that  spirochetes  tolerant  to  antisyphilitic  mercurials  are  not 
tolerant  to  arsenicals.  However,  Blumenthal  and  Pan  (1963)  noted  that 
penicillin-resistant  strains  of  staphylococci  are  more  apt  to  be  resistant  to 


DEVELOPMENT    OF   EESISTANCE    TO    MERCURIALS  985 

Hg++  than  the  normal  strains.  Resistance  to  the  mercurials  is  usually  not 
lost  during  subsequent  culturing  in  mercurial-free  media,  as  shown  for  fungi 
(Partridge  and  Rich,  1962),  yeast  (Zambonelli,  1958  a),  and  spirochetes 
(Launoy  and  Levaditi,  1913).  The  most  striking  case  of  the  retention  of 
resistance  is  that  of  Salmonella,  tolerance  being  unchanged  during  55  trans- 
fers over  a  period  of  18  months  (Severens  and  Tanner,  1945).  It  has  thus 
been  concluded  that  the  tolerance  is  inherited.  Morphological  changes  dur- 
ing adaptation  generally  do  not  occur,  but  yeast  cells  tolerant  to  Hg++ 
are  smaller,  have  lost  their  smooth  contours,  contain  more  vacuoles,  and  are 
deficient  in  lipid  (Imshenetsky  and  Perova,  1957).  Growth  of  tolerant  or- 
ganisms in  normal  media  is  usually  not  different  from  that  of  sensitive 
strains,  and  no  instance  of  dependence  on  mercurials  has  been  reported. 

Development  of  resistance  to  mercurials  is  not  confined  to  microorgan- 
isms but  is  observed  in  certain  mammalian  tissues.  Gil  y  Gil  (1924)  claimed 
to  have  produced  renal  tolerance  to  Hg++  in  rabbits,  but  his  work  was 
criticized  by  Hunter  (1929),  who  repeated  and  extended  this  study  using 
more  approijriate  dosages.  Rabbits  given  nephrotoxic  doses  of  mercurial 
exhibited  a  regeneration  of  the  renal  epithelium  around  the  fourth  day, 
the  new  cells  being  elongated,  flattened,  with  hyperchromatic  nuclei,  and 
somewhat  resistant  to  Hg++.  Tsurumaki  et  al.  (1928)  confirmed  that  the 
kidneys  of  rabbits  surviving  toxic  injections  of  HgCl,  are  not  damaged  by 
the  same  doses  if  given  10-14  days  after  the  disappearance  of  the  original 
nephritis.  During  the  period  of  adaptation,  even  though  repeated  subcutane- 
ous injections  are  given,  the  nephritis  disappears,  but  there  is  no  alteration 
in  the  excretion  of  Hg++  (Miura,  1934).  MacNider  (1941)  repeated  such  ex- 
periments in  dogs  and  claimed  that  the  newly  regenerated  renal  epithelial 
cells  are  different  both  morphologically  and  chemically  from  the  normal 
cells.  The  depressant  effects  of  meralluride  on  intestinal  transport  of  Na+, 
CI",  and  water  also  disappear  with  repeated  administration  of  the  drug 
(Blickenstaff,  1954),  as  does  the  suppression  of  conditioned  reflexes  in  rats 
given  HgCL  (Galoyan,  1957). 


REFERENCES 

Abadom,  P.N.  and  Scholefield,  P.G.  (1962).  Can.  J.  Biochem.  Physiol.  40,  1591. 

Abe,  S.  (1963).  Biol.  Bull.  124,  107. 

Abood,  L.G.  and  Romanchek,  L.  (1955).  Biochem.  J.  60,  233. 

Abraham,  E.P.  and  Newton,  G.G.F.  (1956).  Biochem.  J.  63,  628. 

Abraham,  S.,  Matthes,  K.J.,  and  Chaikoff,  I.L.  (1960).  Biochem.  Biophys.  Res.  Com- 

mun.  3,  646. 
Abraham,  S.,  Matthes,  K.J.  and  Chaikoff,  I.L.  (1961).  Biochim.  Biophys.  Acta  49,  268. 
Abul-Fadl,  M.A.M.  and  King,  E.J.  (1949).  Biochem.  J.  45,  51. 
Ackermann,  W.W.  (1951).  J.  Biol.  Chem.  189,  421. 

Ackermann,  W.W.  and  Potter,  V.R.  (1949).  Proc.  Soc.  Exptl.  Biol.  Med.  11,  1. 
Ackermann,  W.W.  and  Taylor,  A.  (1948).  Proc.  Soc.  Exptl.  Biol.  Med.  67,  449. 
Acland,  J.D.  (1962).  Biochem.  J.  84,  77p. 
Ada,  G.L.  (1963).  Biochim.  Biophys.  Acta  73,  276. 
Adams,  E.  (1954).  J.  Biol.  Chem.  209,  829. 
Adams,  E.  (1955).  J.  Biol.  Chem.  217,  325. 
Adams,  E.  (1959).  J.  Biol.  Chem.  234,  2073. 
Adams,  E.  and  Goldstone,  A.  (1960  a).  J.  Biol.  Chem.  235,  3499. 
Adams,  E.  and  Goldstone,  A.  (1960  b).  J.  Biol.  Chem.  235,  3504. 
Adell,  B.  (1940).  Z.  Physik.  Chem.  A 185,  161. 

Adkins,  H.,  Cramer,  H.I.,  and  Connor,  R.  (1931).  J.  Am.  Chem.  Soc.  53,  1402. 
Adler,  E.,  von  Euler,  H.  Giinther,  G.,  and  Plass,  M.  (1939).  Biochem.  J.  33,  1028. 
Agarwala,  S.C,  Krishna  Murti,  C.R.,  and  Shrivastava,  D.L.  (1954).  Enzymologia  16, 

322. 
Agosin,  M.  and  Aravena,  L.  (1959).  Biochim.  Biophys.  Acta  34,  90. 
Agosin,  M.  and  Aravena,  L.  (1960).  Enzymologia  22,  281. 
Agosin,  M.  and  von  Brand,  T.  (1953).  J.  Infect.  Diseases  93,  101. 
Agosin,  M.  and  von  Brand,  T.  (1955).  Exptl.  Parasitol.  4,  548. 
Agosin,  M.  and  Weinbach,  E.C.  (1956).  Biochim.  Biophys.  Acta  21,  117. 
Agosin,  M.,  von  Brand,  T.,  Rivera,  G.Fr,  and  McMahon,  P.  (1957).  Exptl.  Parasitol. 

6,  37. 
Agranoff,  B.W.,  Eggerer,  H.,  Henning,  U.,  and  Lynen,  F.  (1960).  J.  Biol.  Chem.  235,  326. 
Ahmed,  K.  and  Scholefield,  P.G.  (1962).  Can.  J.  Biochem.  Physiol.  40,  1101. 
Aikawa,  J.K.,  Blumberg,  A.J.,   and   Catterson,  D.A.  (1955).   Proc.  Soc.  Exp.  Biol. 

Med.  89,  204. 
Airan,  J.W.  and  Barnabas,  J.  (1953).  Naturwissenschaften  40,  510. 
Airan,  J.W.,  Joshi,  J.V.,  Barnabas,  J.,  and  Master,  R.W.P.  (1953).  Anal.  Chem.  25,  659. 
Aisenberg,  A.C.,  Reinafarje,  B.,  and  Potter,  V.R.  (1957).  J.  Biol.  Chem.  224,  1099. 
Ajl,  S.J.  and  Werkman,  C.H.  (1948).  Proc.  Natl.  Acad.  Sci.  U.S.  34,  491. 
Akamatsu,  S.,  Kiyomoto,  A.,  Harigaya,  S.,  and  Ohshima,  S.  (1961).  Nature  191,  1298. 
Akiba,  T.  and  Ishii,  K.  (1952).  Japan.  J.  Exptl.  Med.  22,  241. 
Alanis,  J.  (1948).  Arch.  Inst.  Cardiol.  Mex.  18,  410. 
Albaum,  H.G.  and  Eichel,  B.  (1943).  Am.  J.  Botany  30,  18. 
Albers,  R.W.  and  Koval,  G.J.  (1961).  Biochim.  Biophys.  Ada  52,  29. 

987 


988  REFERENCES 

Albert,  A.  (1960).  "Selective  Toxicity,"  2nd  ed.  Wiley,  New  York. 

Albert,  A.,  Falk,  J.E.,  and  Rubbo,  S.D.  (1944).  Nature  153,  712. 

Albrecht,  A.M.  (1957).  Federation  Proc.  16,  144. 

Alburn,  H.E.  and  Whitley,  R.W.  (1954).  Federation  Proc.  13,  330. 

Aleem,  M.I.H.  and  Lees,  H.  (1963).  Can.  J.  Biochetn.  Physiol.  41,  763. 

Alivisatos,  S.G.A.  and  Denstedt,  O.F.  (1952).  J.  Biol.  Chem.  199,  493. 

Alivisatos,  S.G.A. ,  Kashket,  S.,  and  Denstedt,  O.F.  (1956).  Can.  J.  Biochem.  Physiol. 

34,  46. 
Allison,  A.C.  (1962).  Virology  17,  176. 
Allison,  A.C.  and  Cecil,  R.  (1958).  Biochem.  J.  69,  27. 

Allison,  A.C,  Buckland,  F.E.,  and  Andrews,  C.H.  (1962).  Virology  17,  171. 
Almkvist,  J.  (1922).  Arch.  Dermatol.  Syphilol.  141,  342. 
Altekar,  W.W.  and  Rao,  M.R.R.  (1963).  J.  Bacteriol.  85,  604. 
Altszuler,  N.,  Dunn,  A.,  Steele,  R.,  Bishop,  J.S.,  and  De  Bodo,  R.C.  (1963).  Am.  J. 

Physiol.  204,  1008. 
Amaha,  M.  and  Nakahara,  T.  (1959).  Nature  184,  1255. 

Ames,  S.R.  and  Elvehjem,  C.A.  (1944  a).  Proc.  Soc.  Exptl.  Biol.  Med.  57,  108. 
Ames,  S.R.  and  Elvehjem,  C.A.  (1944  b).  Arch.  Biochem.  5,  191. 
Ames,  S.R.  and  Elvehjem,  C.A.  (1945).  J.  Biol.  Chem.  159,  549. 
Anagnostopoulos,  C.  (1953  a).  Bull.  Soc.  Chim.  Biol.  35,  575. 
Anagnostopoulos,  C.  (1953  b).  Bull.  Soc.  Chim.  Biol.  35,  595. 
Anderson,  B.M.  and  Kaplan,  N.O.  (1959).  J.  Biol.  Chem.  234,  1226. 
Anderson,  D.G.,  Rice,  M.S.,  and  Porter,  J.W.  (1960).  Biochem.  Biophys.  Res.  Com- 

mun.  3,  591. 
Anderson,  E.P.  and  Heppel,  L.A.  (1960).  Biochim.  Biophys.  Acta  43,  79. 
Andrews,  G.S.  and  Taylor,  K.B.  (1955).  Brit.  J.  Exptl.  Pathol.  36,  611. 
Angielski,  S.,  Rogulski,  J.,  and  Madonska,  L.  (1960a).  Acta  Biochim.  Polon.  7,  269. 
Angielski,  S.,  Rogulski,  J.,  Mikulski,  P.,  and  Popinigis,  J.  (1960b).  Acta  Biochim. 

Polon.  7,  285. 
Angielski,  S.,  Rogulski,  J.,  and  Basciak,  J.  (1960  c).  Acta  Biochim.  Polon.  7,  295. 
Anichkov,  S.V.  (1953).  DoM.  na  Mezhdunar.  Fiziol.  Kongr.,  Montreal  19,  7. 
Annau,  E.  (1935).  Z.  Physiol.  Chem.  236,  58. 
Anson,  M.L.  (1939  a).  J.  Gen.  Physiol.  23,  239. 
Anson,  M.L.  (1939  b).  J.  Gen.  Physiol.  23,  247. 
Anson,  M.L.  (1940).  J.  Gen.  Physiol.  23,  321. 
Anson,  M.L.  (1941).  J.  Gen.  Physiol.  24,  399. 
Anson,  M.L.  and  Mirsky,  A.E.  (1931).  J.  Gen.  Physiol.  14,  605. 
Anson,  M.L.  and  Stanley,  W.M.  (1941).  J.  Gen.  Physiol.  24,  679. 
Anthony,  D.D.,  Starr,  J.L.,  Kerr,  D.S.,  and  Goldthwait,  D.A.  (1963).  J.  Biol  Chem. 

238,  690. 
Aposhian,  H.V.  and  Lambooy,  J. P.  (1955).  J.  Am.  Chem.  Soc.  77,  6368. 
Aposhian,  H.V.  and  Pointer,  N.S.  (1958).  Federation  Proc.  17,  344. 
App,  A. A.  and  Meiss,  A.N.  (1958).  Arch.  Biochem.  Biophys.  77,  181. 
Applewhite,  T.H.,  Martin,  R.B.,  and  Niemann,  C.  (1958).  J.  Am..  Chem.  Soc.  80,  1457. 
Appleyard,  G.  and  Woods,  D.D.  (1956).  J.  Gen.  Microbiol.  14,  351. 
Araki,  K.  and  Myers,  D.K.  (1963).  Can.  J.  Biochem.  Physiol.  41,  2157. 
Arbuthnott,  J.P.  (1962).  Nature  196,  277. 
Archibald,  R.M.  (1944).  J.  Biol.  Chem.  154,  657. 


REFERENCES  989 

Arkin,  A.  (1911).  J.  Pharmacol.  Exptl.  Therap.  3,  145. 

Arkin,  A.  (1912).  J.  Infect.  Diseases  11,  427. 

Armstrong,  J.  McD.,  Coates,  J.H.,  and  Morton.  R.K.  (1960).  Xafure  186.  1033. 

Armstrong,  J.  McD.,  Coates,  J.H.,  and  Morton,  R.K.  (1963).  Biochem.  J.  88,  266. 

Arndt,  F.  (1955).  Org.  Syn.  20,  26. 

Amon,  D.I.,  Allen,  M.B.,  and  Whatley,  F.R.  (1956).  Biochim.  Biophys.  Acta  20,  449. 

Asahi,  T.,  Bandurski,  R.S.,  and  Wilson,  L.G.  (1961).  J.  Biol.  Chem.  236,  1830. 

Asano,  A.  and  Brodie,  A.F.  (1963).  Biochem.  Biophys.  Res.  Commun.  13,  423. 

Asnis,  R.E.,  Vely,  V.G.,  and  Glick,  M.C.  (1956).  J.  Bacteriol.  11,  314. 

Astrup,  T.  and  Thorsell,  W.  (1954).  Acta  Chem.  Scand.  8,  1859. 

Atkinson,  D.E.  (1956).  J.  Biol.  Chem.  218,  557. 

Atkinson,  M.R.,  Jackson,  J.F.,  and  Morton,  R.K.  (1961).  Nature  192,  946. 

Atkinson,  M.R.,  Morton,  R.K.,  and  Murray,  A.W.  (1963).  Biochem.  J.  89,  167. 

Aubel,  E.  and  Szulmajster,  J.  (1950).  Biochim.  Biophys.  Acta  5,  255. 

Aiibel,  E.,  Rosenberg.  A.J.,  and  Szulmajster,  J.  (1950).  Biochim.  Biophys.  Acta  5,  228. 

Auditore,  G.V.  and  Holland,  W.C.  (1956).  Am.  J.  Physiol.  187,  57. 

Augustinsson,  K.-B.  and  Heimbiirger,  G.  (1955).  Acta  Chem.  Scand.  9,  383. 

Austen,  K.F.  and  Brocklehurst,  W.E.  (1960).  Nature  186,  866. 

Avi-Dor,  Y.  and  Gonda,  O.  (1959).  Biochem.  J.  72,  8. 

Avi-Dor,  Y..  KuczjTiski,  M.,  Schatzberg,  G-,  and  Mager,  J.  (1956).  J.  Gen.  Microbiol. 

14,  76. 
Avi-Dor,  Y.,  Traub,  A.,  and  Mager,  J.  (1958).  Biochim.  Biophys.  Acta  30,  164. 
Avron,  M.  and  Biale.  J.B.  (1957).  Plant  Physiol.  32,  100. 
Avron,  M.  and  Jagendorf,  A.T.  (1956).  Arch.  Biochem.  Biophys.  65,  475. 
Axelrod,  A.E.,  Purvis,  S.E.,  and  Hofmann,  K.  (1948).  J.  Biol.  Chem.  176,  695. 
Axelrod,  B.  and  Jang,  R.  (1954).  J.  Biol.  Chem.  209,  847. 
Axelrod,  B.,  Saltman,  P.,  Bandurski,  R.S.,  and  Baker,  R.S.  (1952).  J.  Biol.  Chem. 

197,  89. 
Axelrod,  J.  and  Cochin,  J.  (1956).  Federation  Proc.  15,  395. 
Axelrod,  J.  and  Cochin,  J.  (1957).  J.  Pharmacol.  Exptl.  Therap.  121,  107. 
Axelrod,  J.  and  Laroche,  M.-J.  (1959).  Science  130,  800. 
Ayengar,  P.  and  Roberts,  E.  (1952).  Proc.  Soc.  Exptl.  Biol.  Med.  79,  476. 
Azoulay,  E.  and  Heydeman,  M.T.  (1963).  Biochim.  Biophys.  Acta  73,  1. 
Azzone,  G.F.  and  Carafoli,  E.  (1960).  Exptl.  Cell  Res.  21,  447. 
Azzone,  G.F.  and  Ernster,  L.  (1961).  J.  Biol.  Chem.  lib,  1518. 
Azzone,  G.F.,  Carafoli,  E.,  and  Muscatello,  U.  (1960).  Exptl.  Cell  Res.  21,  456. 

Bach,  M.K.  (1961).  Plant  Physiol.  36,  317. 

Bach,  S.J.  and  Williamson,  S.  (1942).  Nature  150,  575. 

Bachmann,  H.  (1938).  Arch.  Exptl.  Pathol.  Pharmakol.  190,  345. 

Bacq,  Z.M.  (1936).  Arch.  Intern.  Physiol.  42,  340. 

Bacq,  Z.M.  (1942).  Bull.  Acad.  Roy.  Med.  Belg.  7,  108. 

Bacq,  Z.M.,  Gosselin,  L.,  Dresse,  A.,  and  Renson,  J.  (1959).  Science  130,  453. 

Baer,  H.  (1948).  J.  Biol.  Chem.  173,  211. 

Baernstein,  H.D.  (1936).  J.  Biol.  Chem.  115,  25. 

Bahn,  R.C.  and  Longley,  J.B.  (1956).  J.  Pharmacol.  Exptl.  Therap.  118,  365. 

Bailey,  K.  and  Marsh,  B.B.  (1952).  Biochim.  Biophys.  Acta  9,  133. 

Bailey,  K.  and  Perry,  S.V.  (1947).  Biochim.  Biophys.  Acta  1,  506. 


990  REFERENCES 

Bain,  J.A.  and  Williams,  H.L.  (1960).  In  "Inhibition  in  the  Nervous  System  and  y- 

Aminobutyric  Acid"  (E.  Roberts,  ed.),  p.  275.  Macmillan  (Pergamon),  New   York 
Baker,  N.  and  Wilson,  L.  (1963).  Biochem.  Biophys.  Res.  Commun.  11,  60. 
Baker,  Z.,  Fazekas,  J.F.,  and  Himwich,  H.E.  (1938).  J.  Biol.  Chem.  125,  545. 
Baldwin,  E.  (1938).  Biochem.  J.  32,  1225. 
Balinsky,  D.  and  Davies,  D.D.  (1961a).  Biochem.  J.  80,  292. 
Balinsky,  D.  and  Davies,  D.D.  (1961b).  Biochem.  J.  80,  296. 
Balis,  M.E.  and  Dancis,  J.  (19.55).  Cancer  Res.  15,  603. 
Ball,  H.A.  and  Sanders,  C.R.  (1958).  Federation  Proc.  17,  427. 
Balls,  A.K.  and  Lineweaver,  H.  (1939  a).  Nature  144,  513. 
Balls,  A.K.  and  Lineweaver,  H.  (1939  b).  J.  Biol.  Chem.  130,  669. 
Baltscheffsky,  H.  (1960  a).  Biochim,.  Biophys.  Acta  40,  1. 
Baltscheffsky,  H.  (1960  b).  Acta  Chem.  Scand.  14,  264. 

Baltscheffsky,  H.  and  Baltscheffsky,  M.  (1958).  Acta  Chem.  Scand.  12,  1333. 
Bandelin,  F.J.  (1958).  J.  Am.  Pharm.  Assoc,  Sci.  Ed.  47,  691. 
Bandurski,  R.S.  (1955).  J.  Biol.  Chem.  217,  137. 
Banga,  I.,  Ochoa,  S.,  and  Peters,  R.A.  (1939).  Biochem.  J.  33,  1109. 
Baptist,  J.N.  and  Ve.stling,  C.S.  (1957).  Federation  Proc.  16,  150. 
Barany,  M.  (1959).  In  "Sulfur  in  Proteins"  (R.  Benesch  et  al.,  eds.),  p.  317.  Academic 

Press,  New  York. 
Barany,  M.  and  Barany,  K.  (1959  a).  Biochim.  Biophys.  Acta  35,  293. 
Barany,  M.  and  Barany,  K.  (1959  b).  Biochim.  Biophys.  Acta  35,  544. 
Barany,  M.,  Finkelman,  F.,  and  Therattil-Antony,  T.  (1962).  Arch.  Biochem.  Biophys. 

98,  28. 
Barban,  S.  (1961).  Biochim.  Biophys.  Acta  47,  604. 
Barban,  S.  (1962  a).  Biochim.  Biophys.  Acta  65,  376. 
Barban,  S.  (1962  b).  J.  Biol.  Chem.  237,  291. 
Barban,  S.  and  Schulze,  H.O.  (19.56).  J.  Biol.  Chem.  Ill,  665. 
Barban,  S.  and  Schulze,  H.O.  (1961).  J.  Biol.  Chem.  236,  1887. 
Barbate,  L.M.  and  Abood,  L.G.  (1963).  Biochim.  Biophys.  Acta  67,  531. 
Barber,  G.A.  (1959).  Federation  Proc.  18,  185. 

Barbour,  H.G.  and  Hjort,  A.M.  (1921).  J.  Pharmacol.  Exptl.  Therap.  18,  201. 
Bargoni,  N.  (1963).  Z.  Physiol.  Chem.  333,  242. 
Barker,  S.B.  (1953).  Federation  Proc.  12,  9. 

Barman,  T.E.,  Parke,  D.V.,  and  Williams,  R.T.  (1961).  Biochem.  Pharmacol.  8,  46. 
Barman,  T.E.,  Parke,  D.V.,  and  Williams,  R.T.  (1963).  Toxicol.  Appl.  Pharmacol.  5, 

545. 
Barnabas,  J.  (1955).  Anal.  Chem.  27,  443. 
Barnard,  E.A.  and  Ramel,  A.  (1962).  Biochem.  J.  84,  72p. 
Barnard,  E.A.  and  Stein,  W.D.  (1958).  Advan.  Enzymol.  20,  51. 
Barnes,  H.  and  Stanbury,  F.A.  (1948).  J.  Exptl.  Biol.  25,  270. 
Barnes,  J.M.,  Magee,  P.N.,  Boyland,  E.,  Haddow,  A.,  Passey,  R.D.,  Bullough,  W.S., 

Cruickshank,  C.N.D.,  Salaman,  M.H.,  and  Williams,  R.T.  (1957).  Nature  180,  62. 
Barnett,  R.C.  (1953).  Biol.  Bull.  104,  263. 
Barnett,  R.C.  and  Downey,  M.  (1955).  Federation  Proc.  14,  9. 
Barone,  J.A.,  Tieckelmann,  H.,  Guthrie,  R.,  and  Holland,  J.F.  (1960).  J.  Org.  Chem. 

25,  211. 
Barrett,  F.R.  (1957).  Med.  J.  Australia  p.  242. 


REFERENCES  991 

Barron,  E.S.G.  (1951).  Advan.  Enzymol.  11,  201. 

Barron,  E.S.G.  and  Dickman,  S.  (1949).  J.  Gen.  Phijsiol  32,  595. 

Barron,  E.S.G.  and  Ghiretti,  F.  (1953).  Biochim.  Biophys.  Acta  12,  239. 

Barron,  E.S.G.  and  Goldinger,  J.M.  (1941).  Proc.  Soc.  Exptl.  Biol.  Med.  48,  570. 

Barron,  E.S.G.  and  Kalnitsky,  G.  (1947).  Biochem.  J.  41,  346. 

Barron,  E.S.G.  and  Levine,  S.  (1952).  Arch.  Biochem.  Biophys.  41,  175. 

Barron,  E.S.G.  and  Singer,  T.P.  (1945).  J.  Biol.  Chem.  157,  221. 

Barron,  E.S.G.  and  Tahmisian,  T.N.  (1948).  J.  Cellular  Comp.  Physiol.  32,  57. 

Barron,  E.S.G.,  Dick,  G.F.,  and  Lyman,  CM.  (1941).  J.  Biol.  Chem.  137,  267. 

Barron,  E.S.G.,  Nelson,  L.,  and  Ardao,  M.I.  (1948).  J.  Gen.  Physiol.  32,  179. 

Barron,  E.S.G.,  Ardao,  M.I.,  and  Hearon,  M.  (1950).  Arch.  Biochem.  Biophys.  29,  130. 

Barron,  E.S.G.,  Sights,  W.P.,  and  Wilder,  V.  (1953).  Arch.  Exptl.  Pathol.  Pharmakol. 

219,  338. 
Barry,  R.J.C.,  Dikstein,  S.,  Matthews,  J.,  Smyth,  D.H.,  and  Wright,  E.M.  (1964). 

J.  Physiol.  [London)  171,  316. 
Bartlett,  G.R.  (1948).  J.  Am.  Chem.  Soc.  70,  1010. 
Bartley,  W.  and  Davies,  R.E.  (1954).  Biochem.  J.  57,  37. 

Bassham,  J.A.,  Benson,  A.A.,  and  Calvin,  M.   (1950).  J.  Biol.  Chem.  185,  781. 
Bastarrachea,  F.,  Anderson,  D.G.,  and  Goldman,  D.S.  (1961).  J.  Bacteriol.  82,  94. 
Battaglia,  F.C.  and  Randle,  P.J.  (1960).  Biochem.  J.  75,  408. 
Bauchop,  T.  and  Dawes,  E.A.  (1959).  Biochim.  Biophys.  Acta  36,  294. 
Bauer,  C.W.  and  La  Sala,  E.F.  (1956).  J.  Am.  Pharm.  Assoc,  Sci.  Ed.  45,  675. 
Baugh,  C.L.,  Claus,  G.W.,  and  Werkman,  C.H.  (1960).  Arch.  Biochem.  Biophys.  86, 

255. 
Baumann,  C.A.  and  Stare,  F.J.  (1940).  J.  Biol.  Chem.  133,  183. 
Baxter,  C.F.  and  Roberts,  E.  (1958).  J.  Biol.  Chem.  233,  1135. 
Baxter,  R.M.  and  Quastel,  J.H.  (1953).  J.  Biol.  Chem.  201,  751. 
Bayne,  S.  (1952).  Biochem.  J.  50,  xxvii. 
Bean,  R.C.  and  Hassid,  W.Z.  (1956).  Science  124,  171. 
Beaton,  G.H.  and  McHenry,  E.W.  (1953).  Brit.  J.  Nutr.  7,  357. 
Beaton,  J.R.  and  Goodwin,  M.E.  (1955).  J.  Biol.  Chem.  212,  195. 
Beaton,  J.R.,  Goodwin,  M.E.,  Ozawa,  G.,  and  McHenry,  E.W.  (1954).  Arch.  Biochem. 

Biophys.  51,  94. 
Beck,  G.E.  and  Bein,  H.J.  (1948).  Helv.  Physiol.  Pharmacol.  Acta  6,  398. 
Becker,  B.  (1961).  Am.  J.  Physiol.  200,  804. 
Becker,  B.  and  Friedenwald,  J.S.  (1949).  Arch.  Biochem.  22,  101. 
Becker,  C.E.  (1954).  Federation  Proc.  13,  180. 
Becker,  C.E.  and  Day,  H.G.  (1953).  J.  Biol.  Chem.  201,  795. 
Becker,  C.E.  and  May,  C.E.  (1949).  J.  Am.  Chem.  Soc.  71,  1491. 
Becker,  V.  and  Rauschke,  J.  (1951).  Z.  Ges.  Exptl.  Med.  117,  374. 
Becker,  V.  and  Rieken,  E.  (1954).  Arch.  Pathol.  Anat.  Physiol.  325,  1. 
Beckmann,  F.  (1934).  Arch.  Ges.  Physiol.  234,  469. 
Beechey,  R.B.  (1961).  Nature  192,  975. 
Beers,  R.F.,  Jr.  (1961).  J.  Biol.  Chem.  236,  2703. 
Beevers,  H.  (1952).  Plant  Physiol.  TJ,  725. 
Beevers,  H.  (1954).  Plant  Physiol.  29,  265. 
Beevers,  H.  (1956).  Plant  Physiol.  31,  440. 
Beevers,  H.  and  French,  R.C.  (1954).  Arch.  Biochem.  Biophys.  50,  427. 


992  REFERENCES 

Beevers,  H.,  Goldschmidt,  E.P.,  and  Koffler,  H.  (1952).  Arch.  Biochem..  Biophys.  39, 

236. 
Behal,  F.J.  (1959).  Federation  Proc.  18,  188. 

Behal,  F.J.  and  Hamilton,  R.D.  (1962).  Arch.  Biochem.  Biophys.  96,  530. 
Beher,  W.T.  and  Anthony,  W.L.  (1953).  J.  Biol.  Chem.  203,  895. 
Beher,  W.T.,  Holliday,  W.M.,  and  Gaebler,  O.H.  (1952).  J.  Biol.  Chem.  198,  573. 
Beher,  W.T.,  Baker,  G.D.,  and  Madoff,  M.  (1959).  J.  Biol.  Chem.  234,  2388. 
Beinert,  H.  (1960).  In  "The  Enzymes"  (P.D.  Boyer,  H.  Lardy,  and  K.  Myrback,  eds.), 

Vol.  2,  p.  339.  Academic  Press,  New  York. 
Beinert,  H.  and  Stansly,  P.G.  (19.53).  J.  Biol.  Chem.  204,  67. 
Beling,  C.G.  and  Diczfalusy,  E.  (1959).  Biochem.  J.  71,  229. 
Belkin,  M.  and  Hardy,  W.G.  (1961).  J.  Biophys.  Biochem.  Cytol.  9,  733. 
Bell,  F.E.  and  Mounter,  L.A.  (1958).  J.  Biol.  Chem.  233,  900. 
Bellin,  S.A.  and  Smeby,  R.R.  (1958).  Arch.  Biochem.  Biophys.  75,  1. 
Bello,  L.J.,  van  Bibber,  M.J.,  and  Bessman,  M.J.  (1961).  Biochim.  Biophys.  Acta  53, 

194. 
Benesch,  R.  and  Benesch,  R.E.  (1952).  Arch.  Biochem.  Biophys.  38,  425. 
Benesch,  R.  and  Benesch,  R.E.  (1962).  In  "Methods  of  Biochemical  Analysis"  (D. 

Glick  ed.),  Vol.  10,  p.  43.  Wiley  (Interscience),  New  York. 
Benesch,  R.,  Benesch,  R.E.,  and  Rogers,  W.I.  (1954).  In  "Glutathione"  (S.P.  Colo- 
wick,  ed.),  p.  31.  Academic  Press,  New  York. 
Benesch,  R.E.  and  Benesch,  R.  (1953).  J.  Am.  Chem.  8oc.  75,  4367. 
Benesch,  R.E.  and  Benesch,  R.  (1954).  Arch.  Biochem.  Biophys.  48,  38. 
Benesch,  R.E.  and  Benesch,  R.  (19.55).  J.  Am..  Chem..  Soc.  77,  5877. 
Benesch,  R.E.  and  Benesch,  R.  (1962).  Biochemistry  1,  735. 
Benigno,  P.  and  Santi,  R.  (1946).  Ric.  Sci.  16,  1779. 
Bennett,  D.R.,  Andersen,  K.S.,  Andersen,  M.V.,  Jr.,  Robertson,  D.N.,  and  Cheno- 

weth,  M.B.  (1958).  J.  Pharmacol.  Exptl.  Therap.  '[11,  489. 
Bennett,  E.D.  and  Bauerle,  R.H.  (1960).  Australian  J.  Biol.  Sci.  13,  142. 
Bennett,  H.S.  (1948  a).  Anat.  Record  100,  640. 
Bennett,  H.S.  (1948  b).  Anat.  Record  100,  732. 
Bennett,  H.S.  (1951).  Anat.  Record  110,  231. 
Bentley,  L.E.  (1952).  Nature  170,  847. 
Bentley,  R.  and  Keil,  J.G.  (1961).  Proc.  Chem.  Soc.  p.  111. 
Bentley,  R.  and  Thiessen,  C.P.  (1957).  J.  Biol.  Chem..  lib,  689. 
Beppu,  T.  (1961).  J.  Biochem.  [Tokyo)  SO,  386. 

Ber,  A.  and  Singer-Altbeker,  R.  (1961).  Biochem.  Pharmacol.  7,  169. 
Bergel,  F.  and  Todd,  A.R.  (1937).  J.  Chem.  Soc.  p.  1504. 
Berger,  J.  and  Avery,  G.S.,  Jr.  (1943  a).  Am.  J.  Botany  30,  290. 
Berger,  J.  and  Avery,  G.S.,  Jr.  (1943  b).  Am.  J.  Botany  30,  297. 
Berger,  M.  and  Harman,  J.W.  (1954).  Federation  Proc.  13,  183. 
Bergersen,  F.J.  (1960).  J.  Gen.  Microbiol.  11,  671. 
Bergersen,  F.J.  (1962).  J.  Gen.  Microbiol.  29,  113. 
Bergkvist,  R.  (1963).  Acta  Chem.  Scand.  17,  1541. 
Bergmann,  F.,  Segal,  R.,  and  Rimon,  S.  (1957).  Biochem.  J.  67,  481. 
Bergmann,  F.,  Levin,  G.,  and  Kwietny,  H.  (1959).  Arch.  Biochem.  Biophys.  80,  318. 
Bergmann,  F.,  Kwietny-Govrin,  H.,  Ungar-Waron,  H.,  Kalmus,  A.,  and  Tamari,  M. 

(1963  a).  Biochem.  J.  86,  567. 


REFERENCES  993 

Bergmann,  F.,  Ungar-Waron,  H.,  and  Kwietny-Govrin,  H.  (1963  b).  Biochem.  J.  86, 

292. 
Bergmeyer,  H.-U.,  Holz,  G.,  Kauder,  E.M.,  Mollering,  H.,  and  Wieland,  O.  (1961). 

Biochem.  Z.  333,  471. 
Bergquist,  P.L.  (1958).  Physiol.  Plantarum  11,  760. 
Berk,  S.,  Ebert,  H.,  and  Teitell,  L.  (1957).  Ind.  Eng.  Chem.  49,  1115. 
Berlin,  M.  and  Gibson,  S.  (1963).  Arch.  Environ.  Health  6,  617. 
Berlin,  M.  and  Ullberg,  S.  (1963).  Arch.  Environ.  Health  6,  589. 
Berman,  D.A.  (1951).  Ph.D.  Thesis,  University  of  Southern  CaHfornia. 
Berman,  D.A.  and  Saunders,  P.R.  (1955).  Circulation  Res.  3,  559. 
Bernfeld,  P.  (1963).  In  "Metabolic  Inhibitors"  (R.M.  Hochster  and  J.H.  Quastal,  eds.), 

Voh  2,  p.  437.  Academic  Press,  New  York. 
Bernfeld,  P.  and  Kelley,  T.F.  (1963).  J.  Biol.  Chem.  238,  1236. 
Bernfekl,  P.,  Tuttle,  L.P.,  and  Hubbard,  Pv.W.  (1961).  Arch.  Biochern.  Biophys.  92, 

232. 
Bernhard,  S.A.  and  Gutfreund,  H.  (1956).  Biochem.  J.  63,  61. 
Bernheim,  A. I.,  Hirschhorn,  L.,  and  Muhnos,  M.G.  (1932).  J.  Pharmacol.  Exptl.  Therap. 

44,  81. 
Bernheim,  F.  (1928).  Biochem.  J.  22,  1178. 
Bernheim,  F.  (1963).  Proc.  Soc.  Exptl.  Biol.  Med.  113,  411. 
Bernheim,  F.  and  Dixon,  M.  (1928).  Biochem.  J.  22,  113. 
Bernheim,  F.,  DeTurk,  W.E.,  and  Pope,  H.  (1953).  J.  Bacteriol.  65,  544. 
Bernstein,  D.E.  (1944).  Arch.  Biochem.  3,  445. 

Bernstein,  G.S.  and  Black,  R.E.  (1959).  Proc.  Soc.  Exptl.  Biol.  Med.  102,  531. 
Berry,  L.J.  (1955).  Ann.  N.  Y.  Acad.  Sci.  62,  327. 
Berry,  L.J.  and  Beuzeville,  C.  (1960).  J.  Infect.  Diseases  106,  183. 
Berry,  L.J.  and  Derbyshire,  J.E.  (1956).  Proc.  Soc.  Exptl.  Biol.  Med.  92,  315. 
Berry,  L.J.  and  Mitchell,  R.B.  (1953  a).  Science  118,  140. 
Berry,  L.J.  and  Mitchell,  R.B.  (1953  b).  J.  Infect.  Diseases  93,  75. 
Berry,  L.J.  and  Mitchell,  R.B.  (1953  c).  J.  Infect.  Diseases  93,  83. 
Berry,  L.J.  and  Mitchell,  R.B.  (1954).  Proc.  Soc.  Exptl.  Biol.  Med.  85,  209. 
Berry,  L.J.,  Merritt,  P.,  and  Mitchell,  R.B.  (1954  a).  J.  Infect.  Diseases  94,  144. 
Berry,  L.J.,  Ehlers,  K.H.,  and  Mitchell,  R.B.  (1954b).  J.  Infect.  Diseases  94,  152. 
Bersin,  T.  and  Logemann,  W.  (1933).  Z.  Physiol.  Chem.  220,  209. 
Bertino,  J.R.,  Gabrio,  B.W.,  and  Huennekens,  F.M.  (1960).  Biochem.  Biophys.  Res. 

Commun.  3,  461. 
Beyer,  K.H.  and  Baer,  J.E.  (1960).  Progr.  Drug  Res.  2,  9. 
Beyer,  K.H.,  Wright,  L.D.,  Skeggs,  H.R.,  Russo,  H.F.,  and  Shaner,  G.A.  (1947).  Am. 

J.  Physiol.  1 51 ,  202. 
Beyer,  R.E.  (1960).  Biochim.  Biophys.  Acta  41,  552. 
Beyer,  R.E.,  Lamberg,  S.L.,  and  Neyman,  M.A.  (1961).  Can.  J.  Biochem.  Physiol.  39, 

73. 
Bhuvaneswaran,  C,  Sreenivasan,  A.,  and  Rege,  D.V.  (1961).  Enzymologia  23,  185. 
Bhuyan,  B.K.  and  Simpson,  F.J.  (1962).  Can.  J.  Microbiol.  8,  737. 
Bickers,  J.N.,  Bresler,  E.H.,  and  Weinberger,  R.  (1960).  J.  Pharmacol.  Exptl.  Therap. 

128,  283. 
Bide,  R.W.  and  Collier,  H.B.  (1964).  Can.  J.  Biochem.  Physiol.  42,  669. 
Bieleski,  R.L.  (1960).  Australian  J.  Biol.  Sci.  13,  221. 


994  REFEKENCES 

Bilinski,  E.  and  Jonas,  R.E.E.  (1964).  Can.  J.  Biochem.  Physiol.  42,  345. 

Bilodeau,  F.  and  Elliott,  K.A.C.  (1963).  Can.  J.  Biochem.  Physiol.  41,  779. 

Birkenhager,  J.C.  (1959).  Biochim.  Biophys.  Acta  31,  595. 

Birkenhager,  J.C.  (1960).  Biochim.  Biophys.  Acta  40,  182. 

Bivings,  L.  and  Lewis,  G.  (1948).  J.  Pediat.  32,  63. 

Bjerrum,  J.   (1941).   "Metal  Ammine  Formation  in  Aqueous  Solution."  P.   Haase, 

Copenhagen. 
Bjerrum,  J.,  Schwarzenbach,  G.,  and  Sillen,  L.G.  (1957).  "Stability  Constants,"  Part 

1.  Chemical  Soc,  London. 
Black,  M.M.  and  Kleiner,  I.S.  (1947).  Cancer  Res.  7,  717. 
Black,  M.M.,  Kleiner,  I.S.,  and  Bolker,  H.  (1947).  Cancer  Res.  7,  818. 
Black,  S.  (1951).  Arch.  Biochem.  Biophys.  34,  86. 
Blair,  D.G.R.  and  Potter,  V.R.  (1961).  J.  Biol.  Chem.  236,  2503. 
Blakley,  E.R.  and  Boyer,  P.D.  (1955).  Biochim.  Biophys.  Acta  16,  576. 
Blakley,  E.R.  and  Ciferri,  0.  (1961).  Can.  J.  Microbiol.  7,  61. 
Blakley,  R.L.  (1952).  Biochem.  J.  52,  269. 
Blakley,  R.L.  (1954).  Biochem.  J.  58,  448. 
Blakley,  R.L.  (1957).  Biochem.  J.  65,  342. 

Blakley,  R.L.  and  McDougall,  B.M.  (1961).  J.  Biol.  Chem.  236,  1163. 
Blanchard,  M.,  Green,  D.E.,  Nocito,  V.,  and  Ratner,  S.  (1944).  J.  Biol.  Chem.  155, 

421. 
Blaschko,  H.  (1935).  Biochem.  J.  29,  2303. 
Blaschko,  H.  (1942).  J.  Physiol.  (London)  101,  337. 
Blaschko,  H.  and  Duthie,  R.  (1945).  Biochem.  J.  39,  347. 
Blaschko,  H.  and  Hinims,  J.M.  (1955).  Brit.  J.  Pharmacol.  10,  451. 
Blaschko,  H.,  Fastier,  F.N.,  and  Wajda,  I.  (1951).  Biochem.  J.  49,  250. 
Blaylock,  B.A.  and  Nason,  A.  (1963).  J.  Biol.  Chem.  238,  3453. 
Bhckenstaff,  D.D.  (19.54).  Am.  J.  Physiol.  178,  371. 
Blum,  J.J.  (1955).  Arch.  Biochem.  Biophys.  55,  486. 
Blum,  J.J.  (1960).  Arch.  Biochem.  Biophys.  87,  104. 
Blum,  J.J.  (1962  a).  Arch.  Biochem.  Biophys.  97,  309. 
Blum,  J.J.  (1962  b).  Arch.  Biochem.  Biophys.  97,  321. 

Blumenstein,  J.  and  Williams,  G.R.  (1963).  Can.  J.  Biochem.  Physiol.  41,  201. 
Blumenthal,  H.J.  and  Pan,  Y.-L.  (1963).  Proc.  Soc.  Exptl.  Biol.  Med.  113,  322. 
Blumson,  N.L.  (1957).  Biochem.  J.  65,  138. 

Bock,  K.R.,  Robinson,  J.B.D.,  and  Chamberlain,  G.T.  (1958).  Nature  182,  1607. 
Bodansky,  O.  (1961).  J.  Biol.  Chem.  236,  328. 
Boeri,  E.  and  Tosi,  L.  (1954).  Arch.  Biochem.  Biophys.  52,  83. 
Boeri,  E.,  Cutolo,  E.,  Luzzati,  M.,  and  Tosi,  L.  (1955).  Arch.  Biochem.  Biophys.  56, 

487. 
Boeri,  E.,  Cremona,  T.,  and  Singer,  T.P.  (1960).  Biochem.  Biophys.  Res.  Commun.  2, 

298. 
Bone,  D.H.  and  Hochster,  R.M.  (1960).  Can.  J.  Biochem.  Physiol.  38,  193. 
Boney,  A.D.,  Corner,  E.D.S.,  and  Sparrow,  B.W.P.  (1959).  Biochem.  Pharmacol.  2,  37. 
Bonner,  B.A.  and  Machlis,  L.  (1957).  Plant  Physiol.  32,  291. 
Bonner,  J.  (1948).  Arch.  Biochem.  17,  311. 
Bonner,  J.  (1949).  Am.  J.  Botany  36,  429. 
Bonner,  J.  and  Millerd,  A.  (1953).  Arch.  Biochem.  Biophys.  42,  135. 


REFEKENCES  995 

Bonner,  J.  and  Wildman,  S.G.  (1946).  Arch.  Biochem.  10,  497 

Bonner,  J.,  Bandurski,  R.S.,  and  Millerd,  A.  (1953).  Physiol.  Plantarum  6,  511. 

Bonner,  W.D.,  Jr.  (1954).  Biochem.  J.  56,  274. 

Bonsignore,  A.,  Pontremcli,  S.,  Grazi,  E.,  and  Horecker,  B.L.  (1960).  J.  Biol.  Chem. 

235,  1888. 
Booth,  A.N.,  Taylor,  J.,  Wilson,  R.H.,  and  DeEds,  F.  (1952).  J.  Biol.  Chem.  195,  697. 
Borghgraef,  R.R.M.  and  Pitts,  R.F.  (1956).  J.  Clin.  Invest.  35,  31. 
Borghgraef,  R.R.M.,  Kessler,  R.H.,  and  Pitts,  R.F.  (1956).  J.  Clin.  Invest.  35,  1055. 
Borsook,  H.,  Fischer,  E.H.,  and  Keighley,  G.  (1957).  J.  Biol.  Chem.  229,  1059. 
Borst,  P.  (1962).  Biochim.  Biophys.  Acta  57,  256. 
Bortz,  W.M.  and  Lynen,  F.  (1963).  Biochem.  Z.  337,  305. 
Boser,  H.  (1959).  Z.  Physiol.  Chem.  315,  163. 
Bosund,  I.  (1959).  Acta  Chem.  Scand.  13,  803. 
Bosund.  I.  (1960  a).  Acta  Chem.  Scand.  14,  1231. 
Bosund,  I.  (1960  b).  Acta  Chem.  Scand.  14,  1243. 
Boucek,  R.J.  and  Paff,  G.H.  (1961).  Am.  J.  Physiol.  201,  668. 
Bovarnick,  M.R.,  Lindsay,  A.,  and  Hellerman,  L.  (1946).  J.  Biol.  Chem.  163,  523. 
Bowman,  I.B.R.,  Grant,  P.T.,  Kermack,  W.O.,  and  Ogston,  D.  (1961).  Biochem.  J. 

78,  472. 
Bowman,  I.B.R.,  Tobie,  E.J.,  and  von  Brand,  T.  (1963).  Comp.  Biochem..  Physiol.  9,  105. 
Boxer,  G.E.,  Shonk,  C.E.,  Stoerk,  H.C.,  Bielinski,  T.C.,  and  Budzilovich,  T.V.  (1955). 

J.  Biol.  Chem.  217,  541. 
Boyd,  E.S.  and  Neuman,  W.F.  (1954).  Arch.  Biochem.  Biophys.  51,  475. 
Boyer,  P.D.  (1954).  J.  Am.  Chem.  Soc.  76,  4331. 
Boyer,  P.D.  (1959).  In  "The  Enzymes"  (P.D.  Boyer,  H.  Lardy,  and  K.  Myrback,  eds.). 

Vol.  1,  p.  511.  Academic  Press,  New  York. 
Boyer,  P.D.  and  Segal,  H.L.  (1954).  In  "The  Mechanism  of  Enzyme  Action"  (W.D. 

McElroy  and  H.B.  Glass,  eds.),  p.  520.  Johns  Hopkins  Press,  Baltimore,  Maryland. 
Boyland,  E.  (1940).  Biochem.  J.  34,  1196. 
Boyland,  E.  and  Boyland,  M.E.  (1936).  Biochem.  J.  30,  224. 
Boyland,  E.  and  Levi,  A.A.  (1936).  Biochem.  J.  30,  2007. 

Boyland,  E.,  Wallace,  D.M.,  and  Williams,  D.C.  (1957).  Brit.  J.  Cancer  11,  578. 
Boyland,  E.,  Davies,  J.H.,  and  Williams,  K.  (1959).  Biochem.  J.  72,  28p. 
Boylen,  J.B.  and  Quastel,  J.H.  (1962).  Nature  193,  376. 
Bradshaw,  W.H.  and  Barker,  H.A.  (1960).  J.  Biol.  Chem.  235,  3620. 
Brady,  C.J.  (1961).  Biochem.  J.  78,  631. 
Brady,  R.O.  (1963).  Biochim.  Biophys.  Acta  70,  467. 
Brady,  R.O.  and  Gurin,  S.  (1952).  J.  Biol.  Chem.  199,  421. 

Brady,  R.O.,  Spyropoulos,  C.S.,  and  Tasaki,  I.  (1958).  Am.  J.  Physiol.  194,  207. 
Brand,  E.  and  Kassell,  B.  (1940).  J.  Biol.  Chem.  133,  437. 
Branster,  M.V.  and  Morton,  R.K.  (1956).  Biochem.  J.  63,  640. 
Braun,  K.  (1949).  J.  Pharmacol.  Exptl.  Therap.  95,  58. 
Braune,  W.  (1963).  Arch.  Mikrobiol.  45,  252. 
Bramistein,  A.E.  (1960).  In  "The  Enzymes"  (P.D.  Boyer,  H.  Lardy,  and  K.  Myrback, 

eds.),  Vol.  2,  p.  113.  Academic  Press,  New  York. 
Braunstein,  A.E.,  Azarkh,  R.M.,  and  Tin-Sen',  S.  (1962).  Biochemistry  (USSR)  (Eng- 
lish Transl.)  26,  760. 
Bray,  R.E.  and  VanArsdel,  P.P.  (1961).  Proc.  Soc.  Exptl.  Biol.  Med.  106,  255. 


996  REFERENCES 

Bregoff,  H.M.  and  Kamen,  M.D.  (1952).  Arch.  Biochem.  Biophys.  36,  202. 

Bresnick,  E.  (1962).  Cancer  Res.  22,  1246. 

Bresnick,  E.  (1963).  Biochim.  Biophys.  Acta  67,  425. 

Bressler,  R.  and  Wakil,  S.J.  (1961).  J.  Biol.  Chem.  236,  1643. 

Bressler,  R.  and  Wakil,  S.J.  (1962).  J.  Biol.  Chem.  237,  1441. 

Breusch,  F.L.  (1942).  Enzymologia  10,  165. 

Breusch,  F.L.  and  Keskin,  H.  (1944).  Enzymologia  11,  243. 

Brewer,  C.R.  and  Werkman,  C.H.  (1939).  Enzymologia  6,  273. 

Briault,  P.G.  (1956).  Nature  178,  1223. 

Bridger,  W.A.  and  Cohen,  L.H.  (1963).  Biochim.  Biophys.  Acta  73,  514. 

Briggs,  F.N.,  Chernick,  S.,  and  Chaikoff,  I.L.  (1949).  J.  Biol.  Chem.  179,  103. 

Brink,  N.G.  (1953  a).  Acta  Chem.  Scand.  7,  1081. 

Brink,  N.G.  (1953  b).  Acta  Chem.  Scand.  7,  1090. 

Brock,  N.,  Druckrey,  H.,  and  Herken,  H.  (1939).  Arch.  Exptl.  Pathol.  Pharmakol. 

193,  679. 
Brockman,  R.W.  and  Anderson,  E.P.  (1963).  In  "Metabolic  Inhibitors"  (R.M.  Hoch- 

ster  and  J.H.  Quastel,  eds.),  Vol.  1,  p.  239.  Academic  Press,  New  York. 
Brodersen,.R.  and  Kjaer,  A.  (1946).  Acta  Pharmacol.  Toxicol.  2,  109. 
Brodie,  A.F.  (1952).  J.  Biol.  Chem.  199,  835. 
Brody,  T.M.  (1956).  J.  Pharmacol.  Exptl.  Therap.  117,  39. 
Brooks,  J.L.,  Christensen,  E.,  Stewart,  C.J.,  and  Wick,  A.N.  (1961).  Proc.  Soc.  Exptl. 

Biol.  Med.  107,  395. 
Brooks,  P.  and  Davidson,  N.  (1960).  J.  Am.  Chem.  Soc.  82,  2118. 
Brooks,  S.A.,  Lawrence,  J.C,  and  Ricketts,  C.R.  (1959).  Biochem.  J.  73,  566. 
Brooks,  S.A.,  Lawrence,  J.C,  and  Ricketts,  C.R.  (1960).  Nature  187,  1028. 
Broquist,  H.P.,  Kohler,  A.R.,  Hutchinson,  D.J.,  and  Burchenal,  J.H.  (1953).  J.  Biol. 

Chem.  202,  59. 
Brown,  D.D.,  Tomchick,  R.,  and  Axelrod,  J.  (1959).  J.  Biol.  Chem.  234,  2948. 
Brown,  E.G.  (1958).  Biochem.  J.  70,  313. 
Brown,  LA.  (1954).  A.M.A.  Arch.  Neurol.  Psychiat.  72,  674. 
Brown,  J.  and  Bachrach,  H.L.  (1959).  Proc.  Soc.  Exptl.  Biol.  Med.  100,  641. 
Brown,  W.D.  and  Tappel,  A.L.  (1959).  Arch.  Biocheyn.  Biophys.  85,  149. 
Bruchmann,  E.-E.  (1961a).  Naturivissenschaften  48,  79. 
Bruchmann,  E.-E.  (1961b).  Naturwissenschaften  48,  305. 
Bruchmann,  E.-E.  (1961c).  Biochem.  Z.  335,  199. 
Bruemmer,  J.H.,  Wilson,  P.W.,  Glenn,  J.L.,  and  Crane,  F.L.  (1957).  J.  Bacterial. 

73,  113. 
Brun,  C,  Hilden,  T.,  and  Raaschou,  F.  (1947).  Acta  Pharmacol.  Toxicol.  3,  1. 
Brunnemann,  A.,  Coper,  H.,  and  Herken,  H.  (1962).  Arch.  Exptl.  Pathol.  Pharmakol. 

223,  223. 
Brunngraber,  E.G.  (1958).  J.  Biol.  Chem.  233,  472. 

Bruns,  F.H.,  Noltman,  E.,  and  Vahlhaus,  E.  (1958).  Biochem.  Z.  330,  483. 
Brust,  M.  and  Barnett,  R.C.  (1952).  Federation  Proc.  11,  19. 
Bryce,  G.F.  and  Rabin,  B.R.  (1964).  Biochem.  J.  90,  513. 
Bublitz,  C.  and  York,  J.L.  (1961).  Biochim.  Biophys.  Acta  48,  56. 
Buch,  M.L.,  Montgomery,  R.,  and  Porter,  W.L.  (1952).  Anal.  Chem.  24,  489. 
Buchman,  E.R.,  Heegaard,  E.,  and  Bonner,  J.  (1940).  Proc.  Natl.  Acad.  Sci.  U.S. 

26,  561. 


REFERENCES  997 

Bueding,  E.  and  MacKinnon,  J.A.  (1955).  J.  Biol.  Chem.  215,  495. 

Bulen,  W.A.  (1956).  Arch.  Biochem.  Biophys.  62,  173. 

Bulen,  W.A.  (1961).  J.  Baderiol.  82,  130. 

Bullough,  W.S.  and  Johnson,  M.  (1951).  Proc.  Boy.  Soc.  B138,  562. 

BuUough,  W.S.  and  Laurence,  E.B.  (1957).  Brit.  J.  Exptl.  Pathol.  38,  278. 

Bu'Lock,  J.D.,  Smalley,  H.M.,  and  Smith,  G.N.  (1962).  J.  Biol.  Chem.  237,  1778. 

Burch,  G.,  Ray,  T.,  Threefoot,  S.,  Kelly,  F.J.,  and  Svedberg,  A.  (1950).  J.  Clin. 

Invest.  29,  1131. 
Burch,  H.B.,  Lowry,  O.H.,  Padilla,  A.M.,  and  Combs,  A.M.  (1956).  J.  Biol.  Chem. 

223,  29. 
Burdette,  W.J.  (1952).  Am.  Heart  J.  44,  823. 

Burgess,  E.A.,  Kerly,  M.,  and  Rahman,  M.A.  (1960).  Biochem.  J.  74,  4p. 
Burnett,  G.H.  and  Cohen,  P.P.  (1957).  J.  Biol.  Chem.  229,  337. 
Burney,  T.E.  and  Golub,  O.J.  (1948).  J.  Immunol.  60,  213. 
Burr,  M.J.  and  Finamore,  F.J.  (1963).  Biochim.  Biophys.  Acta  68,  608. 
Burris,  R.H.  (1956).  In  "Inorganic  Nitrogen  MetaboHsm"  (W.D.  McElroy  and  B. 

Glass,  eds.),  p.  316.  Johns  Hopkins  Press,  Baltimore,  Maryland. 
Burris,  R.H.  and  Wilson,  P.W.  (1939).  Cold  Spring  Harbor  Symp.  Quant.  Biol.  7,  349. 
Burton,  K.  (1951a).  Biochem.  J.  48,  458. 
Burton,  K.  (1951b).  Biochem.  J.  50,  258. 
Busch,  H.  and  Nair,  P.V.  (1957).  J.  Biol.  Chem.  229,  377. 
Busch,  H.  and  Potter,  V.R.  (1952  a).  J.  Biol.  Chem.  198,  71. 
Busch,  H.  and  Potter,  V.R.  (1952  b).  Cancer  Res.  12,  660. 
Busch,  H.  and  Potter,  V.R.  (1953).  Cancer  Res.  13,  168. 

Busch,  H.,  Hurlbert,  R.B.,  and  Potter,  V.R.  (1952).  J.  Biol.  Chem.  196,  717. 
Butt,  W.D.  and  Lees,  H.  (1960).  Biochem.  J.  76,  425. 
Buzard,  J.A.  and  Nytch,  P.D.  (1957).  J.  Biol.  Chem.  Til,  225. 
Byers,  S.O.  (1952).  J.  Am.  Pharm.  Assoc,  Sci.  Ed.  41,  611. 

Cady,  P.,  Abraham,  S.,  and  Chaikoff,  I.L.  (1963).  Bicchim.  Biophys.  Acta  70,  118. 
Caffrey,  R.W.,  Trembley,  R.,  Gabrio,  B.W.,  and  Huennekens,  F.M.  (1956).  J.  Biol. 

Chem.  112,  1. 
Cafruny,  E.J.  (1962).  Biochem:.  Phmmacol.  9,  15. 

Cafruny,  E.J.  and  Farah,  A.  (1956).  J.  Pharmacol.  Exptl.  Therap.  117,  101. 
Cafruny,  E.J.  and  Palmer,  J.F.  (1961).  J.  Pharmacol.  Exptl.  Therap.  131,  250. 
Cafruny,  E.J.,  di  Stefano,  H.S.,  and  Farah,  A.  (1955  a).  J.  Histochem.  Cytochem.  3,  354. 
Cafruny,  E.J.,  Farah,  A.,  and  di  Stefano,  H.S.  (1955b).  J.  Pharmacol.  Exptl.  Therap. 

115,  390. 
Cain,  R.B.  (1958).  J.  Gen.  Microbiol.  19,  1. 
Calcutt,  G.  (1950).  Nature  166,  443. 

Caldwell,  E.F.  and  McHenry,  E.W.  (1953).  Arch.  Biochem.  Biophys.  45,  466. 
Calesnick,  B.,  Wase,  A.,  and  Altarelli,  V.R.  (1960).  Proc.  Soc.  Exptl.  Biol.  Med.  103, 

22. 
Callaghan,  O.H.  and  Weber,  G.  (1959).  Biochem.  J.  73,  473. 
Caltrider,  P.G.  and  Gottlieb,  D.  (1963).  Phytopathology  53,  1021. 
Calvin,  M.  (1954).  In  "Glutathione"  (S.P.  Colowick  et  al.,  eds.),  p.  3.  Academic  Press, 

New  York. 
Calvin,  M.,  Magel,  T.T.,  and  Hurd,  CD.  (1941).  J.  Am.  Chem.  Soc.  63,  2174. 


998  REFERENCES 

Campbell,  D.E.S.  (1959).  Acta  Pharmacol.  Toxicol.  16,  151. 

Campbell,  P.N.,  Creasey,  N.H.,  and  Parr,  C.W.  (1952).  Biochem.  J.  52,  448. 

Canellakis,  Z.N.  and  Cohen,  P.P.  (1956  a).  J.  Biol.  Chem.  222,  53. 

Canellakis,  Z.N.  and  Cohen,  P.P.  (1956  b).  J.  Biol.  Chem.  222,  63. 

Cannata,  J.J.B.  and  Stoppani,  A.O.M.  (1963  a).  Nature  200,  573. 

Cannata,  J.J.B.  and  Stoppani,  A.O.M.  (1963  b).  J.  Biol.  Chem.  238,  1208. 

Cannata,  J.J.B.  and  Stoppani,  A.O.M.  (1963  c).  J.  Biol.  Chem.  238,  1919. 

Caraway,  W.T.  and  Hellerman,  L.  (1953).  J.  Am.  Chem.  Soc.  75,  5334. 

Carbon,  J.A.  (1962).  Biochem.  Biophys.  Res.  Commun.  7,  366. 

Carney,  S.A.,  Lawrence,  J.C,  and  Ricketts,  C.R.  (1962).  Biochem.  J.  83,  533. 

Carpenter,  A.T.  and  Scott,  J.J.  (1961).  Biochiyn.  Biophys.  Acta  52,  195. 

Carpenter,  T.M.  and  Benedict,  F.G.  (1909).  Am.  J.  Physiol.  24,  187. 

Carpenter,  W.D.  and  Beevers,  H.  (1959).  Plant  Physiol.  34,  403. 

Carr,  A.J.  (1963).  Nature  198,  1104. 

Cartwrigt,  N.J.  and  Cain,  R.B.  (1959).  Biochem.  J.  71,  248. 

Cascarano,  J.  and  Zweifach,  B.W.  (1962).  Federation  Proc.  21,  151. 

Castelfranco,  P.,  Stumpf,  P.K.,  and  Contopoulou,  C.R.  (1955).  J.  Biol.  Chem.  214,  567. 

Caughey,  W.S.,  Smiley,  J.D.,  and  Hellerman,  L.  (1957).  J.  Biol.  Chem..  224,  591. 

Cavallito,  C.J.  and  Haskell,  T.H.  (1945).  J.  Am.  Chem.  Soc.  67,  1991. 

Cecil,  R.  and  Snow,  N.S.  (1962).  Biochem.  J.  82,  247. 

Cedrangolo,  F.  (1941).  Atti  Accad.  Italia,  Rend.  Classe  Sci.  Fis.,  Mat.  Nat.  2,  78. 

Cedrangolo,  F.  and  Adler,  E.  (1939).  Enzymologia  7,  297. 

Cerecedo,  L.R.  and  Eich,  S.  (19.55).  J.  Biol.  Chem.  213,  893. 

Cerecedo,  L.R.,  Soodak,  M.,  and  Eusebi,  A.J.  (1951).  J.  Biol.  Chem.  189,  293. 

Cerecedo,  L.R.,  Eich,  S.,  and  Bresnick,  E.  (1954).  Biochim.  Biophys.  Acta  15,  144. 

Cerfontain,  H.  and  van  Aken,  G.M.F.A.  (1956).  J.  Am.  Chem.  Soc.  78,  2094. 

Cha,  C.-Y.  and  Scheraga,  H.A.  (1961a).  Biochem.  Biophys.  Res.  Commun.  5,  67. 

Cha,  C.-Y.  and  Scheraga,  H.A.  (1961b).  Biochem.  Biophys.  Res.  Commun.  6,  369. 

Chaberek,  S.  and  Martell,  A.E.  (1959).  "Organic  Sequestering  Agents."  Wiley,  New 

York. 
Chaffee,  R.R.  and  Bartlett,  W.L.  (1960).  Biochim.  Biophys.  Acta  39,  370. 
ChaikoflF,  I.L.  and  Brown,  G.W.,  Jr.  (1954).  In  "Chemical  Pathways  of  Metabolism" 

(D.M.  Greenberg,  ed.).  Vol.  1,  p.  277.  Academic  Press,  New  York. 
Challenger,  F.,  Subramaniam,  V.,  and  Walker,  T.K.  (1927).  J.  Chem.  Soc.  p.  200. 
Chambers,  R.,  Cameron,  G.,  and  Kopac,  M.J.  (1943).  Cancer  Res.  3,  293. 
Chance,  B.  (1954).  In  "The  Mechanism  of  Enzyme  Action"  (W.D.  McEkoy  and  B, 

Glass,  eds.),  p.  399.  Johns  Hopkins  Press,  Baltimore,  Maryland. 
Chance,  B.  and  Hollunger,  G.  (1961a).  J.  Biol.  Chem.  236,  1534. 
Chance,  B.  and  Hollunger,  G.  (1961b).  J.  Biol.  Chem.  236,  1555. 
Chance,  B.  and  Hollunger,  G.  (1961c).  J.  Biol.  Chem.  236,  1562. 
Chapman,  D.W.  and  Shaffer,  C.F.  (1947).  Arch.  Internal  Med.  79,  449. 
Chappell,  J.B.  (1964  a).  Biochem.  J.  90,  225. 
Chappell,  J.B.  (1964  b).  Biochem.  J.  90,  237. 
Chappell,  J.B.  and  Greville,  G.D.  (1958).  Nature  182,  813. 
Chappell,  J.B.  and  Perry,  S.V.  (1955).  Biochim.  Biophys.  Acta  16,  285. 
Chari-Bitron,  A.  (1961).  Biochem.  Pharmacol.  6,  169. 

Chari-Bitron,  A.  and  Avi-Dor,  Y.  (1959  a).  Biochim.  Biophys.  Acta  32,  276. 
Chari-Bitron,  A.  and  Avi-Dor,  Y.  (1959b).  Biochem.  J.  71,  572. 


REFERENCES  999 

Charmandarjan,  M.O.  and  Tjutjunnikowa,  A.W.  (1930).  Biochem.  Z.  Ill,  284. 
Chattaway,  F.W.,  Thompson,  C.C,  and  Barlow,  A.J.E.  (1956).  Biochem.  J.  63,  648. 
Chatterjee,  G.C.  and  Haider,  D.K.  (1960).  Naturwissenschaften  47,  86. 
Chaudhuri,  D.K.  (1950  a).  Ann.  Biochem.  Ezptl.  Med.  (Calcutta)  10,  65. 
Chaudhuri,  D.K.  (1950  b).  Ann.  Biochem.  Exptl.  Med.  (Calcutta)  10,  71. 
Cheesman,  D.F.,  Keeler,  M.M.,  and  Sten-Knudsen,  O.  (1959).  Biochem..  J.  72,  357. 
Cheftel,  R.I.,  Munier,  R.,  and  Macheboeuf,  M.   (1952).  Bull.  Soc.  Chim.  Biol.  34, 

380. 
Cheldelin,  V.H.  and  Beinert,  H.  (1952).  Biochim..  Biophys.  Acta  9,  661. 
Chen,  K.K.,  Rose,  C.L.,  and  Clowes,  G.H.A.  (1934).  Am.  J.  Med.  Sci.  188,  767. 
Chen,  R.F.  and  Plant,  G.W.E.  (1963).  Biochemistry  2,  1023. 
Chen,  R.F.,  Brown,  D.M.,  and  Plant,  G.W.E.  (1964).  Biochemistry  3,  552. 
Cheniae,  G.  and  Evans,  H.J.  (1959).  Biochim.  Biophys.  Acta  35,  140. 
Chiba,  H.,  Kawai,  F.,  and  Kondo,  K.  (1954  a).  Bull.  Res.  Inst.  Food  Sci.,  Kyoto  Univ. 

No.  13,  p.  12. 
Chiba,  H.,  Kawai,  F.,  and  Kondo,  K.  (1954  b).  Bull.  Res.  Inst.  Food  Sci.,  Kyoto  Univ. 

13,  23. 
Chiba,  H.,  Kawai,  F.,  and  Kondo,  K.  (1954  c).  Bull.  Res.  Inst.  Food  Sci.,  Kyoto  Univ. 

13,  53. 
Chiba,  Y.,  Inada,  A.,  and  Yoshihara,  I.  (1953).  Enzymologia  16,  148. 
Chick,  H.  (1908).  J.  Hyg.  8,  92. 

Chiga,  M.  and  Plant,  G.W.E.  (1959).  J.  Biol.  Chem.  234,  3059. 
Chin,  H.P.  (1963).  Ph.D.  thesis.  University  of  Southern  California. 
Chinard,  F.P.  (1942).  Proc.  Soc.  Exptl.  Biol.  Med.  51,  317. 
Chinard,  F.P.  and  Hellerman,  L.  (1954).  In  "Methods  of  Biochemical  Analysis"  (D. 

Glick,  ed.).  Vol.  1,  p.  1.  Wiley  (Interscience),  New  York. 
Chirigos,  M.A.,  Greengard,  P.,  and  Udenfriend,  S.  (1960).  J.  Biol.  Chem.  235,  2075. 
Choppin,  P.W.  and  Phihpson,  L.  (1961).  J.  Exptl.  Med.  113,  713. 
Christensen,  E.  and  Wick,  A.N.  (1963).  Proc.  Soc.  Exptl.  Biol.  Med.  112,  430. 
Christensen,  E.,  Brooke,  J.L.,  Stewart,  C.J.,  and  Wick,  A.N.  (1961).  Biochem.  Biophys. 

Res.  Commun.  5,  209. 
Christensen,  H.N.  and  Riggs,  T.R.  (1952).  J.  Biol.  Chem.  194,  57. 
Christensen,  H.N.  and  Ross,  W.F.  (1941).  J.  Biol.  Chem.  137,  101. 
Christensen,  H.N.  and  Streicher,  J.A.  (1949).  Arch.  Biochem.  23,  96. 
Christensen,  H.N.,  Riggs,  T.R.,  and  Coyne,  B.A.  (1954).  J.  Biol.  Chem.  209,  413. 
Christie,  G.S.,  Judah,  J.D.,  and  Rees,  K.R.  (1953).  Proc.  Roy.  Soc.  B141,  523. 
Christman,  A.A.  and  Lewis,  H.B.  (1921).  J.  Biol.  Chem.  47,  495. 
Chung,  A.E.  and  Langdon,  R.G.  (1963).  J.  Biol.  Chem.  238,  2317. 
Cianci,  V.  (1940).  Boll.  Soc.  Ital.  Biol.  Sper.  15,  288. 
Ciferri,  O.  (1962).  Enzymologia  24,  283. 

Ciferri,  O.  and  Blakley,  E.R.  (1959).  Can.  J.  Microbiol.  5,  547. 
Citri,  N.  and  Garber,  N.  (1960).  Biochim.  Biophys.  Acta  38,  50. 
Citri,  N.  and  Garber,  N.  (1961).  Biochem.  Biophys.  Res.  Commun.  4,  143. 
Citri,  N.  and  Garber,  N.  (1963).  Biochim.  Biophys.  Acta  67,  64. 
Clagett,  CO.,  Tolbert,  N.E.,  and  Burris,  R.H.  (1949).  J.  Biol.  Chem.  178,  977. 
Clark,  D.S.  and  Wallace,  R.H.  (1958).  Can.  J.  Microbiol.  4,  125. 
Clark,  W.M.  (1960).  "Oxidation-reduction  Potentials  of  Organic  Systems."  Williams 

&  Wilkins,  Baltimore,  Maryland. 


1000  REFERENCES 

Clarke,  D.D.,  Mycek,  M.J.,  Neidle,  A.,  and  Waelsch,  H.  (1959).  Arch.  Biochem.  Bio- 

phys.  79,  338. 
Clarkson,  T.W.  and  Crpss,  A.C.  (1961).  AEC  Research  and  Development  Report  (UR- 

588),  University  of  Rochester,  Rochester,  New  York. 
Clarkson,  T.W.  and  Kench,  J.E.  (1956).  Biochem.  J.  62,  361. 
Claus,  D.  (1956).  Arch.  Mikrohiol.  25,  10. 
Clayton,  R.A.  (1959).  Arch.  Biochem.  Biophys.  79,  HI. 
Clayton,  R.K.  (1957).  Arch.  Mikrohiol.  26,  29. 

Clayton,  R.K.,  Dettmer,  F.H.,  and  Robertson,  A.E.,  Jr.  (1957).  Arch.  Mikrohiol.  26,  20. 
Cleland,  K.W.  (1949).  Proc.  Linnean  Soc.  N.  S.  Wales  74,  296. 
Cleland,  K.W.  and  Rothschild,  Lord  (1952  a).  J.  Exptl.  Biol.  29,  285. 
Cleland,  K.W.  and  Rothschild,  Lord  (1952  b).  J.  Exptl.  Biol.  29,  416. 
Cleland,  R.  and  Bonner,  J.  (1956).  Plant  Physiol.  31,  350. 
Cleland,  W.W.  (1964).  Biochemistry  3,  480. 
Clementi,  A.  (1929).  Boll.  Soc.  Ital.  Biol.  Sper.  4,  503. 
Clements,  A.N.  (1959).  J.  Exptl.  Biol.  36,  665. 
ClifTe,  E.E.  and  Waley,  S.G.  (1961).  Biochem.  J.  79,  475. 
Cochrane,  V.W.  (1952).  J.  Bacteriol.  63,  459. 
Cochrane,  V.W.  and  Peck,  H.D.,  Jr.  (1953).  J.  Bacteriol.  65,  37. 
Coggeshall,  R.E.  and  MacLean,  P.D.  (1958).  Proc.  Soc.  Exptl.  Biol.  Med.  98,  6871 
Cohen,  E.M.  (1953  a).  Acta  Physiol.  Pharmacol.  Neerl.  3,  45. 
Cohen,  E.M.  (1953  b).  Acta  Physiol.  Pharmacol.  Neerl.  3,  59. 
Cohen,  L.H.  and  Bridger,  W.A.  (1964).  Can.  J.  Biochem.  Physiol.  42,  715. 
Cohen,  L.H.  and  Parks,  R.E.,  Jr.  (1963).  Can.  J.  Biochem.  Physiol.  41,  1495. 
Cohen,  P.P.  and  Hayano,  M.  (1946).  J.  Biol.  Chem.  166,  239. 
Cohen,  R.A.  and  Gerard,  R.W.  (1937).  J.  Cellular  Cornp.  Physiol.  10,  223. 
Cohen,  S.  (1963).  Exptl.  Cell  Res.  29,  207. 
Cohn,  M.  and  Horibata,  K.  (1959).  J.  Bacteriol.  78,  624. 

Coleman,  G.H.,  Weed,  L.A.,  and  Myers,  CD.  (1937).  J.  Am.  Chem.  Soc.  59,  2703. 
Coleman,  J.E.  and  Vallee,  B.L.  (1961).  J.  Biol.  Chem.  236,  2244. 
Coleman,  R.  and  Hiibscher,  G.  (1962).  Biochim.  Biophys.  Acta  56,  479. 
Collander,  R.  (1951).  Acta  Chem.  Scand.  5,  774. 
Collias,  E.G.,  McShan,  W.H.,  and  Lilly,  J.H.  (1952).  J.  Cellular  Comp.  Physiol.  40, 

507. 
Colowick,  S.P.,  Lazarow,  A.,  Packer,  E.,  Schwarz,  D.R.,  Stadtman,  E.R.,  and  Waelsch, 

H.   (eds.)  (1954).  "Glutathione,"  Academic  Press,  New  York. 
Colter,  J.S.,  Kuhn,  J.,  and  Ellem,  K.A.O.  (1961).  Cancer  Res.  21,  48. 
Commoner,  B.  and  Hollocher,  T.C.,  Jr.  (1960).  Proc.  Natl.  Acad.  Sci.  U.  S.  46,  405. 
Commoner,  B.  and  Thimann,  K.V.  (1941).  J.  Gen.  Physiol.  24,  279. 
Conchie,  J.  (1953).  Biochem.  J.  55,  xxi. 
Conchie,  J.  (1954).  Biochem.  J.  58,  552. 
Conchie,  J.  and  Hay,  A.J.  (1959).  Biochem.  J.  73,  327. 
Conchie,  J.  and  Levvy,  G.A.  (1957).  Biochem.  J.  65,  389. 
Conn,  E.E.  and  Young,  L.C.T.  (1957).  J.  Biol.  Chem.  226,  23. 

Conover,  T.E.,  Danielson,  L.,  and  Ernster,  L.  (1963a).  Biochim.  Biophys.  Acta  67,  254. 
Conover,  T.E.,  Prairie,  R.L.,  and  Packer,  E.  (1963  b).  J.  Biol.  Chem.  238,  2831. 
Conrad,  (1874).  Ber.  7,  688.  (Cited  by  WisUcenus,  J.) 
Converse,  J.L.  and  Besemer,  A.R.  (1959).  J.  Bacteriol.  78,  231. 


REFERENCES  1001 

Conway,  T.W.,  Lansford,  E.M.,  Jr.,  and  Shive,  W.  (1964).  Arch.  Biochem.  Biophys. 

107,  120. 
Cooil,  B.J.  (1952).  Plant  Physiol.  27,  49. 

Cook,  A.M.  and  Steel,  K.J.  (1959).  J.  Pharm.  Pharmacol.  11,  666. 
Cook,  E.S.  and  Kreke,  C.W.  (1943).  Proc.  Sac.  Exptl.  Biol.  Med.  53,  222. 
Cook,  E.S.  and  Perisutti,  G.  (1947).  J.  Biol.  Chem.  167,  827. 
Cook,  E.S.,  Kreke,  C.W.,  McDevitt,  M.,  and  Bartlett,  M.D.  (1946).  J.  Biol.  Chem. 

162,  43. 
Cook,  R.P.  (1930).  Biochem.  J.  lA,  1538. 
Cook,  S.F.  (1926).  J.  Gen.  Physiol.  9,  575. 

Coombs,  T.L.,  Omote,  Y.,  and  Vallee,  B.L.  (1964).  Biochemistry  3,  653. 
Cooper,  C.  (1958  a).  Biochim.  Biophys.  Acta  30,  484. 
Cooper,  C.  (1958  b).  Biochim.  Biophys.  Acta  30,  529. 
Cooper,  C.  and  Kulka,  R.G.  (1961).  J.  Biol.  Chem.  236,  2351. 
Cooper,  C.  and  Lehninger,  A.L.  (1957).  J.  Biol.  Chem.  224,  561. 
Cooper,  E.A.  and  Goddard,  A.E.  (1957).  J.  Appl.  Chem.  7,  613. 
Cooper,  E.A.  and  Mason,  J.  (1927).  J.  Hyg.  26,  118. 
Cooper,  J.A.D.,  Smith,  W.,  Bacila,  M.,  and  Medina,  H.  (1959).  J.  Biol.  Chem.  234, 

445. 
Cooperstein,  S.J.  (1963).  J.  Biol.  Chem.  238,  3606. 
Coore,  H.G.  and  Randle,  P.J.  (1964).  Biochem.  J.  91,  56. 
Copenhaver,  J.H.,  Jr.  and  Lardy,  H.A.  (1952).  J.  Biol.  Chem.  195,  225. 
Cori,  G.T.  and  Cori,  C.F.  (1940).  J.  Biol.  Chem.  135,  733. 
Cori,  G.T.,  Colowick,  S.P.,  and  Cori,  C.F.  (1939).  J.  Biol.  Chem.  127,  771. 
Corley,  R.C.  and  Rose,  W.C.  (1926).  J.  Pharmacol.  Exptl.  Therap.  27,  165. 
Cormier,  M.J.  (1960).  Biochim.  Biophys.  Acta  42,  333. 
Cormier,  M.J.,  Totter,  J.R.,  and  Rostorfer,  H.H.  (1956).  Arch.  Biochem.  Biophys.  63, 

414. 
Corwin,  L.M.  (1959).  J.  Biol.  Chem.  234,  1338. 

Corwin,  L.M.  and  Schwarz,  K.  (1963).  Arch.  Biochem.  Biophys.  100,  385. 
Cottrell,  T.L.  (1954).  "The  Strengths  of  Chemical  Bonds."  Academic  Press,  New  York. 
Cousins,  F.B.  (1956).  Biochem.  J.  64,  297. 

Covin,  J.M.  (1961).  Ph.D.  thesis,  University  of  Southern  California,  Los  Angeles. 
Covin,  J.M.  and  Berman,  D.A.  (1956).  J.  Pharmacol.  Exptl.  Therap.  117,  443. 
Cowgill,  R.W.  (1959).  J.  Biol.  Chem.  234,  3146. 
Cowgill,  R.W.  and  Pizer,  L.I.  (1956).  J.  Biol.  Chem.  Ill,  885. 
Coxon,  R.V.  and  Robinson,  R.J.  (1956).  Proc.  Roy.  Soc.  B145,  232. 
Cramer,  F.B.  and  Woodward,  G.E.  (1952).  J.  Franklin  Inst.  253,  354. 
Crandall,  D.I.  (1955).  J.  Biol.  Chem.  212,  565. 
Crandall,  D.I.  (1959).  Federation  Proc.  18,  208. 
Crane,  R.K.  (1960).  Biochim.  Biophys.  Acta  45,  477. 
Crane,  R.K.  and  Sols,  A.  (1953).  J.  Biol.  Chem.  203,  273. 
Crane,  R.K.  and  Sols,  A.  (1954).  J.  Biol.  Chem.  210,  597. 
Cravens,  W.W.  and  Snell,  E.E.  (1949).  Proc.  Soc.  Exptl.  Biol.  Med.  71,  73. 
Crawford,  M.A.  (1963).  Biochem.  J.  88,  115. 
Crawhall,  J.C.  and  Watts,  R.W.E.  (1962).  Biochem.  J.  85,  163. 
Creaser,  E.H.  and  Scholefield,  P.G.  (1960).  Cancer  Res.  20,  257. 
Cramer,  J.E.  (1962).  J.  Neurochem.  9,  289. 


1002  REFEKENCES 

Cremer,  J.E.  (1964).  J.  Neurochem.  11,  165. 

Cremona,  T.  (J 964).  J.  Biol.  Chem.  239,  1457. 

Creveling,  C.R.,  Daly,  J.W.,  Witkop,  B.,  and  Udenfriend,  S.  (1962).  Biochim.  Biophys. 

Acta  64,  125. 
Crewther,  W.G.  (1953).  Australian  J.  Biol.  Sci.  6,  205. 
Crewther,  W.G.  (1956).  Biochim.  Biophys.  Acta  21,  178. 
Cross,  R.J.  and  Taggart,  J.V.  (1950).  Am.  J.  Physiol.  161,  181. 
Crout,  J.R.  (1961).  Biochem.  Pharmacol.  6,  47. 

Cruickshank,  D.H.  and  Isherwood,  F.A.  (1958).  Biochem.  J.  69,  189. 
Cumings,  J.N.  (1959).  "Heavy  Metals  and  the  Brain."  Thomas,  Springfield,  Illinois. 
Cummins,  J.T.,  Cheldelin,  V.H.,  and  King,  T.E.  (1957).  J.  Biol.  Chem.  226,  301. 
Cunningham,  L.W.  and  Nuenke,  B.J.  (1959).  J.  Biol.  Chem.  234,  1447. 
Cunningham,  L.W.  and  Nuenke,  B.J.  (1960).  Biochim.  Biophys.  Acta  39,  565. 
Cunningham,  L.W.  and  Nuenke,  B.J.  (1961).  J.  Biol.  Chem.  236,  1716. 
Cunningham,  L.W.,  Nuenke,  B.J.,  and  Strayhorn,  W.D.  (1957).  J.  Biol.  Chem.  228, 

835. 
Czekalowski,  J.W.  (1952).  Brit.  J.  Exptl.  Physiol.  33,  57. 

Dahl,  J.L.,  Way,  J.L.,  and  Parks,  R.E.,  Jr.  (1959).  J.  Biol.  Cheyn.  234,  2998. 

Dahlqvist,  A.  (1959).  Acta  Chem.  Scand.  13,  2156. 

Dajani,  R.M.,  Danielski,  J.,  Gamble,  W.,  and  Orten,  J.M.  (1961).  Biochem.  J.  81, 

494. 
Dale,  R.A.  and  Sanderson,  P.H.  (1954).  Brit.  J.  Pharmacol.  9,  210. 
Dalgarno,  L.  and  Birt,  L.M.  (1962).  Biochem.  J.  83,  202. 
Dalziel,  K.  (1962).  Nature  195,  384. 

Damaschke,  K.  and  Liibke,  M.  (1958).  Z.  Naturforsch.  13b,  172. 
Danforth,  W.  (1953).  Arch.  Biochem..  Biophys.  46,  164. 
Daniel,  L.J.  and  Norris,  L.C.  (1949).  Proc.  Soc.  Exptl.  Biol.  Med.  72,  165. 
Dann,  O.T.  and  Carter,  C.E.  (1964).  Biochem.  Pharmacol.  13,  677. 
Darlington,  W.A.  and  Quastel,  J.H.  (1953).  Arch.  Biochem.  Biophys.  43,  194. 
Das,  H.K.  and  Roy,  S.C.  (1961).  Biochim.  Biophys.  Acta  51,  378. 
Das,  H.K.  and  Roy,  S.C.  (1962).  Arch.  Biochem.  Biophys.  96,  445. 
Das,  N.B.  (1937  a).  Biochem.  J.  31,  1116. 
Das,  N.B.  (1937  b).  Biochem.  J.  31,  1124. 

Das,  N.B.  and  Roy,  A.  (1943).  Trans.  Bose  Res.  Inst.  (Calcutta)  15,  91. 
Das,  S.K.  and  Chatterjee,  G.C.  (1962).  J.  Bacteriol.  83,  1251. 
Das,  S.N.  and  Ives,  D.J.G.  (1961).  Proc.  Chem.  Soc.  p.  373. 
Datta,  A.G.  and  Racker,  E.  (1961).  J.  Biol.  Chem.  236,  617. 
Daus,  L.,  Meinke,  M.,  and  Calvin,  M.  (1952).  J.  Biol.  Chem.  196,  77. 
DaVanzo,  J.P.,  Greig,  M.E.,  and  Cronin,  M.A.  (1961).  Am.  J.  Physiol.  201,  833. 
DaVanzo,  J.P.,  Matthews,  R.J.,  and  Stafford,  J.E.  (1964a).  Toxicol.  Appl.  Pharmacol. 

6,  388. 
DaVanzo,  J. P.,  Matthews,  R.J.,  Young,  G.A.,  and  Wingerson,  F.  (1964  b).  Toxicol. 

Appl.  Pharmacol.  6,  396. 
Davenport,  H.W.  (1954).  Federation  Proc.  13,  32. 
Davenport,  H.W.  and  Chavre,  V.J.  (1956).  Am.  J.  Physiol.  187,  227. 
Davenport,  H.W.,  Chavre,  V.J.,  and  Davenport,  V.D.  (1954).  Am.  J.  Physiol.  177,  418. 
Davenport,  H.W.,  Chavre,  V.J.,  and  Davenport,  V.D.  (1955).  Am.  J.  Physiol.  182,  221. 


REFERENCES  1003 

Davies,  G.E.,  Driver,  G.W.,  Hoggarth,  E.,  Martin,  A.R.,  Paige,  M.F.C.,  Rose,  F.L., 

and  Wilson,  B.R.  (1956).  Brit.  J.  Pharmacol.  11,  351. 
Davis,  F.F.  and  Allen,  F.W.  (1955).  Federation  Proc.  14,  200. 
Davison,  D.C.  (1949).  Proc.  Linnean  Soc.  N.  8.  Wales  74,  37. 
Dawes,  E.A.  and  Holms,  W.H.  (1958).  J.  Bacterial.  75,  390. 
Dawson,  R.M.C.  (1953).  Biochem.  J.  55,  507. 
Dawson,  R.M.C.  and  Thompson,  W.  (1964).  Biochem.  J.  91,  244. 
Day,  M.D.  and  Rand,  M.J.  (1963).  Brit.  J.  Pharmacol.  22,  72. 
de  Caro,  L.,  Rindi,  G.,  and  Grana,  E.  (19.54).  Experientia  10,  140. 
de  Caro,  L.,  Rindi,  G.,  Perri,  V.,  and  Ferrari,  G.  (1956).  Expeientia  12,  300. 
Decker,  R.H.,  Kang,  H.H.,  Leach,  F.R..  and  Henderson,  L.M.  (1961).  J.  Biol.  Chem. 

236,  3076. 
Decker,  R.H..  Brown,  R.R.,  and  Price,  J.M.  (1963).  J.  Biol.  Chem.  238,  1049. 
DeGraff,  A.C.  and  Lehman,  R.A.  (1942).  J.  Am.  Med.  Assoc.  119,  998. 
DeGroot,  C.A.,  Weber,  J.F.,  and  Cohen,  E.M.  (1955).  Arch.  Intern.  Pharmacodyn.  102, 

459. 
DeGroot,  N.  and  Cohen,  P.P.  (1962).  Biochim.  Biophtjs.  Acta  59,  588. 
de  Issaly,  I.S.  and  Stoppani,  A.O.M.  (1964).  Proc.  Soc.  Exptl.  Biol.  Med.  115,  528. 
Deitrich,  R.A.  and  Hellerman,  L.  (1963).  J.  Biol.  Chem.  238,  1683. 
Dejung,  K.  (1963).  Z.  Ges.  Exptl.  Med.  137,  196. 
Dellert,  E.E.  and  Stahmann,  M.A.  (1955).  Nature  176,  1028. 
Delnionte,  L.  and  Jukes,  T.H.  (1962).  Pharmacol.  Rev.  14,  91. 
.  Demis,  D.J.  and  Rothstein,  A.  (1955).  Am.  J.  Physiol.  180,  566. 
DeMoss,  J.A.  and  Swim,  H.E.  (1957).  J.  Bacterial.  74,  445. 
DeMoss,  R.D.  (1955).  In  "Methods  in  Enzymology"  (S.P.  Colowiek  and  N.O.  Kaplan, 

eds.).  Vol.  1,  p.  504.  Academic  Press,  New  York. 
Dempsey,  W'.B.  and  Snell,  E.E.  (1963).  Biochemistry  2,  1414. 
Dengler,  H.  and  Reichel,  G.  (1958).  Arch.  Exptl.  Pathol.  Pharmakol.  234,  275. 
Denison,  F.W^,  Jr.  and  Phares,  E.F.  (1952).  Anal.  Chem.  24,  1628. 
c^ennis,  D.  and  Kaplan,  N.O.  (1960).  J.  Biol.  Chem.  235,  810. 
De  Robertis,  E.  and  Peluifo,  C.A.  (1951).  Proc.  Soc.  Exptl.  Biol.  Med.  78,  584. 
Dervartanian,  D.V.  and  Veeger,  C.  (1962).  Biochem.  J.  84,  65p. 
Desnuelle,  P.  and  Antonin,  S.  (1946).  Helv.  Chim.  Acta  29,  1306. 
Desnuelle,  P.  and  Rovery,  M.  (1949).  Biochim.  Biophys.  Acta  3,  26. 
Desnuelle,  P.,  Antonin,  S.,  and  Casal.  A.  (1947).  Bull.  Soc.  Chim.  Biol.  29,  694. 
DeTurk,  W.E.  and  Bernheim,  F.  (1960).  Arch.  Biochem.  Biophys.  90,  218. 
Dewey,  D.L.,  Hoare,  D.S.,  and  Work,  E.  (1954).  Biochem.  J.  58,  523. 
Deykin,  D.  and  Goodman,  D.S.  (1962).  J.  Biol.  Chem.  237,  3649. 
Dhar,  S.C.  and  Bose,  S.M.  (1962).  Enzymologia  24,  155. 
Di  Carlo,  F.J.  and  Redfern,  S.  (1947).  Arch.  Biochem.  15,  343. 
Dickens,  F.  (1946  a).  Biochem.  J.  40,  145. 
Dickens,  F.  (1946  b).  Biochem.  J.  40,  171. 
Dickens,  F.  and  Glock,  G.E.  (1951).  Biochem.  J.  50,  81. 
Dickens,  F.  and  Williamson,  D.H.  (1956).  Biochem.  J.  64,  567. 
Dickerman,  H.W.,  San  Pietro,  A.,  and  Kaplan,  N.O.  (1962).  Biochim.  Biophys.  Acta 

62,  230. 
Dickman,  S.R.  (1958).  Science  127,  1392. 
Dickman,  S.R.  and  W^estcott,  W.L.  (1953).  Federation  Proc.  12,  197. 


1004  REFERENCES 

Diczfalusy,  E.,  Ferno,  O.,  Fex,  H.,  Hogberg,  B.,  Linderot,  T.,  and  Rosenberg,  T. 
(1953).  Acta  Chem.  Scand.  7,  913. 

Dieterle,  W.  and  Wenzel,  F.  (1944).  Biochem.  Z.  316,  357. 

Dietrich,  E.V.,  Fisher,  A.L.,  Long,  J.P.,  and  Featherstone,  R.M.  (1952).  Arch.  Bio- 
chem. Biophys.  41,  118. 

Dietrich,  L.S.  and  Borries,  E.  (1956).  Arch.  Biochem.  Biophys.  64,  512. 

Dietrich,  L.S.  and  Shapiro,  D.M.  (1953  a).  Proc.  Soc.  Exptl.  Biol.  Med.  84,  555. 

Dietrich,  L.S.  and  Shapiro,  D.M.  (1953  b).  J.  Biol.  Chem.  203,  89. 

Dietrich,  L.S.,  Monson,  W.J.,  Williams,  J.N.,  Jr.,  and  Elvehjem,  C.A.  (1952).  J.  Biol. 
Chem.  1 97,  37. 

Dietrich,  L.S.,  Friedland,  I.M.,  and  Kaplan,  L.A.  (1958).  J.  Biol.  Chem..  233,  964. 

Dikstein,  S.  (1959).  Biochim.  Biophys.  Acta  36,  397. 

Dilley,  R.A.  and  Vernon,  L.P.  (1964).  Biochemistry  3,  817. 

Dils,  R.  and  Popjak,  G.  (1962).  Biochem.  J.  83,  41. 

D'lorio,  A.  (1957).  Ca7i.  J.  Biochem.  Physiol.  35,  395. 

D'lorio,  A.  and  Mavrides,  C.  (1963).  Can.  J.  Biochem.  Phijsiol.  41.  1779. 

Dippy,  J.F.J.  (1939).  Chem.  Rev.  25,  151. 

Di  Sabato,  G.  and  Kaplan,  N.O.  (1963).  Biochemistry  2,  776. 

Dische,  Z.  and  Ash  well,  G.  (1955).  Biochim.  Biophys.  Acta  17,  56. 

Dittrich,  S.  (1963).  J.  Chromatog.  12,  47. 

Dixon,  G.H.,  Kornberg,  H.L.,  and  Lund,  P.  (1960).  Biochim.  Biophys.  Acta  41,  217. 

Dixon,  K.C.  and  Harrison,  K.  (1932).  Biochem.  J.  26,  1954. 

Dixon,  M.  (1925).  Biochem.  J.  19,  507. 

Dixon,  M.  and  Thurlow,  S.  (1924).  Biochem.  J.  18,  976. 

Doctor,  V.M.  (1958).  J.  Biol.  Chem.  232,  617. 

Doctor,  V.M.  (1959).  Blood  14,  1244. 

Dodgson,  K.S.  (1959).  Enzymologia  20,  301. 

Dodgson,  K.S.  and  Powell,  G.M.  (1959).  Biochem.  J.  73,  672. 

Dodgson,  K.S.,  Spencer,  B.,  and  Williams,  K.  (1955).  Biochem.  J.  61,  374. 

Doi,  R.H.  and  Halvorson,  H.  (1961).  J.  Bacteriol.  81,  51. 

Doisy,  R.J.,  Richert,  D.A.,  and  Westerfeld,  W.W.  (1955).  J.  Biol.  Chem.  217,  307. 

Dolin,  M.I.  (1957).  J.  Biol.  Chem.  225,  557. 

Dolin,  M.I.  (1959).  J.  Bacteriol.  77,  383. 

DoHn,  M.I.  and  Gunsalus,  I.C.  (1951).  J.  Bacteriol.  62,  199. 

Dolin,  M.I.  and  Wood,  N.P.  (1960).  J.  Biol.  Chem.  235,  1809. 

Dominguez,  A.M.  and  Shideman,  F.E.  (1953).  Federation  Proc.  12,  316. 

Dominguez,  A.M.  and  Shideman,  F.E.  (19.55).  Proc.  Soc.  Exptl.  Biol.  Med.  90,  329. 

Domonkos,  J.  and  Latzkovits,  L.  (1961).  Arch.  Biochem.  Biophys.  95,  144. 

Donovan,  J.W.  (1964).  Biochemistry  3,  67. 

Doran,  D.J.  (1957).  J.  Protozool.  4,  182. 

Dorfman,  A.,  Roseman,  S.,  Ludowieg,  J.,  Mayeda,  M.,  Moses,  F.E.,  and  Cifonelli, 
J.A.  (19.55).  J.  Biol.  Chem.  216,  549. 

Doudoroff,  M.,  Contopoulou,  C.R.,  and  Burns,  S.  (1958).  Proc.  Intern.  Symp.  Enzyme 
Chem.,  Tokyo  Kyoto  1957  p.  313.  Maruzen,  Tokyo. 

Dove,  W.F.  and  Yamane,  T.  (1960).  Biochem.  Biophys.  Res.  Commun.  3,  608. 

Dowdle,  E.B.,  Schachter,  D.,  and  Schenker,  H.  (1960).  Am.  J.  Physiol.  198,  609. 

Doyle,  M.L.,  Katzman,  P.A.,  and  Doisy,  E.A.  (1955).  J.  Biol.  Chem.  217,  921. 

Drabikowski,  W.  and  Gergely,  J.  (1963).  J.  Biol.  Chem.  238,  640. 


REFERENCES  1005 

Dreser,  H.  (1893).  Arch.  Exptl.  Pathol.  Pharmakol.  32,  456. 
Dreyfus,  P.M.  and  Moniz,  R.  (1962).  Biochim.  Biophys.  Acta  65,  181. 
Dube,  S.K.,  Rohult,  O.,  and  Pressman,  D.  (1963).  J.  Biol.  Chem.  238,  613. 
du  Buy,  H.G.  and  Hesselbach,  M.L.  (1956).  J.  Histochem.  Cytochem.  4,  363. 
Duerkson,  J.D.  and  Halvorson,  H.  (1958).  J.  Biol.  Chem.  233,  1113. 
Duggan,  D.E.  and  Titus,  E.G.  (1959).  J.  Biol.  Chem.  234,  2100. 
Duggan,  D.E.,  Weigert,  M.G.,  and  Titus,  E.G.  (1961).  Cancer  Res.  21,  1047. 
Dumont,  J.E.  (1961).  Biochim.  Biophys.  Acta  50,  506. 
Dumont,  J.E.  (1962).  Biochim.  Biophys.  Acta  56,  382. 
Duperon,  R.  (1956).  Rev.  Gen.  Botan.  63,  137. 
Durham,  N.N.  and  Hubbard,  J.S.  (1959).  Nature  184,  1398. 
Durham,  N.N.  and  Hubbard,  J.S.  (1960).  J.  Bacteriol.  80,  225. 
Dyer,  H.M.  (1938).  J.  Biol.  Chem..  124,  519. 

Dziirik.  R.  and  Krajci-Lazary,  B.  (1962).  Arch.  Intern.  Pharmacodyn.  135,  1. 
Dzurik.  R..  Brixova,  E.,  Krajci-Lazary,  B.,  Hostynova,  D.,  Kolesar,  P.,  and  Nieder- 
land,  T.R.  (1963).  Arch.  Intern.  Pharmacodyn.  141,  396. 

Eaton,  M.D.  (1952).  Arch.  Ges.  Virusforsch.  5,  53. 

Eaton,  M.D.  and  Scala,  A.R.  (1957).  Arch.  Ges.  Virusforsch.  7,  274. 

Eberson,  L.  and  Wadso,  I.  (1963).  Acta  Chem.  Scand.  17,  1552. 

Eberts,  F.S.,  Burris,  R.H.,  and  Riker,  A.J.  (1951).  Am.  J.  Botany  38,  618. 

Edelhoch,  H.,  Katchalski,  E.,  Maybury,  R.H.,  Hughes,  W.L.,  Jr.,  and  Edsall,  J.T. 

(1953).  J.  Am.  Chem.  Soc.  75,  5058. 
Edlbacher,  S.  and  Zeller,  A.  (1936).  Z.  Physiol.  Chem.  242,  253. 
Edlbacher,  S.,  Baur,  H.,  and  Becker.  M.  (1940).  Z.  Physiol.  Chem.  265,  61. 
Edman,  K.A.P.  (1958).  Acta  Physiol.  Scand.  43,  275. 
Edman,  K.A.P.  (1959).  Acta  Physiol.  Scarid.  46,  209. 
Edsall,  J.T.  and  Wyman,  J.  (1958).  "Biophysical  Chemistry,"  Vol.  1.  Academic  Press, 

New  York. 
Edsall,  J.T.,  Maybury,  R.H.,  Simpson,  R.B.,  and  Straessle,  R.  (1954).  J.  Am.  Chem. 

Soc.  76,  3131. 
Edson,  N.L.  (1936).  Biochem.  J.  30,  1855. 
Edson,  N.L.  and  Hunter,  G.J.E.  (1947).  Biochem.  J.  41,  139. 
Edson,  N.L.  and  Leloir,  L.F.  (1936).  Biochem.  J.  30,  2319. 
Edwards,  J.G.  (1942).  Am.  J.  Pathol.  18,  1011. 

Eeg-Larsen,  N.  and  Laland,  S.G.  (1954).  Acta  Physiol.  Scand.  30,  295. 
Egami,  F.  and  Yagi,  K.  (1956).  J.  Biochem.  (Tokyo)  43,  153. 

Egami,  F.,  Naoi,  M.,  Tada,  M.,  and  Yagi,  K.  (1956).  J.  Biochem.  (Tokyo)  43,  669. 
Eggerer,  H.  and  Remberger,  U.  (1963).  Biochem.  Z.  337,  202. 
Ehrmantraut,  H.  and  Rabinowitch,  E.  (1952).  Arch.  Biochem.  Biophys.  38,  67. 
Eich,  S.  and  Cerecedo,  L.R.  (19.54).  J.  Biol.  Chem.  207,  295. 
Eich,  S.  and  Cerecedo,  L.R.  (1955).  Arch.  Biochem.  Biophys.  57,  285. 
Eichel,  B.  and  Wainio,  W.W.  (1948).  J.  Biol.  Chem.  175,  155. 
Eichel,  H.J.  (1956  a).  J.  Biol.  Chem.  Ill,  137. 
Eichel,  H.J.  (1956b).  Biochim.  Biophys.  Acta  11,  571. 
Eichel,  H.J.  (19.59).  Biochem.  J.  71,  106. 

Eichel,  H.J.  and  Rem,  L.T.  (1959).  Biochim.  Biophys.  Acta  35,  571. 
Eichel,  H.J.  and  Rem,  L.T.  (1962).  J.  Biol.  Chem.  237,  940. 


1006  REFERENCES 

Eisenstadt,  J.M.,  Grossman,  L.,  and  Klein,  H.P.  (1959).  Biochim.  Biophys.  Acta  36, 

292. 
Eldjarn,  L.  and  Bremer,  J.  (1962).  Biochem..  J.  84,  286. 
Eldjarn,  L.  and  Pihl,  A.  (1957  a).  J.  Am.  Chem.  Soc.  79,  4589. 
Eldjarn,  L.  and  Pihl,  A.  (1957  b).  J.  Biol.  Chem.  225,  499. 
El  Ha  wary,  M.F.S.  (1955).  Biochem.  J.  61,  348. 
Elion,  G.B.,  Callahan,  S.,  Nathan,  H.,  Bieber,  S.,  Bundles,  R.W.,  and  Hitchings,  G.H. 

(1963).  Biochem.  Pharmacol.  12,  85. 
Elkins-Kaufman,  E.  and  Naurath,  H.  (1949).  J.  Biol.  Chem.  178,  645. 
Ellem,  K.A.O.  and  Colter,  J.S.  (1961).  J.  Cellular  Comp.  Physiol.  58,  267. 
Ellfolk,  N.  (1953).  Acta  Chem.  Scand.  7,  1155. 
Ellfolk,  N.  (1954).  Acta  Chem.  Scand.  8,  151. 

ElUott,  H.W.  and  Sutherland,  V.C.   (1952).  J.  Cellular  Comp.  Physiol.  40,  221. 
Elliott,  K.A.C.  and  Greig,  M.E.  (1937).  Biochem.  J.  31,  1021. 
Elliott,  W.H.  (1955).  In  "Methods  in  Enzymology"  (S.P.  Colowick  and  N.O.  Kaplan, 

eds.).  Vol.  2,  p.  337.  Academic  Press,  New  York. 
Ellis,  R.J.  and  Davies,  D.D.  (1961).  Biochem.  J.  78,  615. 
Ellis,  S.  and  Anderson,  H.L.  (1951a).  J.  Pharmacol.  Exptl.  Therap.  101,  10. 
Ellis,  S.  and  Anderson,  H.L.  (1951b).  J.  Pharmacol.  Exptl.  Therap.  103,  342. 
Ellman,  G.L.  (1959).  Arch.  Biochem.  Biophys.  82,  70. 
Elodi,  P.  (1960).  Biochim.  Biophys.  Acta  40,  272. 
El'tsina,  N.V.  and  Beresotskaya,  N.A.  (1962).  Biochemistry  {USSR)  (English  Transl.) 

27,  386. 
Eltz,  R.W.  and  Vandemark,  P.J.  (1960).  J.  Bacteriol.  79,  763. 
Ely,  J.O.,  TuU,  F.A.,  and  Hard,  J.A.  (1952).  J.  Franklin  Inst.  253,  361. 
Emerson,  G.A.  and  Southwick,  P.L.  (1945).  J.  Biol.  Chem.  160,  169. 
Emerson,  G.A.  and  Tishler,  M.  (1944).  Proc.  Soc.  Exptl.  Biol.  Med.  55,  184. 
Emerson,  G.A.,  Wurtz,  E.,  and  Johnson,  O.H.  (1945).  J.  Biol.  Chem.  160,  165. 
Emerson,  P.M.,  Wilkinson,  J.H.,  and  Withycombe,  W.A.  (1964).  Nature  202,  1337. 
Emery,  T.F.  (1963).  Biochemistry  2,  1041. 

Endahl,  B.R.  and  Kochakian,  CD.  (1961).  Biochim.  Biophys.  Acta  54,  15. 
Engelhardt,  W.  (1958).  Proc.  Intern.  Symp.  Enzyme  Chem.,  Tokyo  Kyoto  1957  p.  163. 

Maruzen,  Tokyo. 
Englard,  S.  and  Colowick,  S.P.  (1957).  J.  Biol.  Chem.  lib,  1047. 
Englard,  S.,  Sorof,  S.,  and  Singer,  T.P.  (1951).  J.  Biol.  Chem.  189,  217. 
Ennor,  A.H.  and  Rosenberg,  H.  (1954).  Biochem.  J.  57,  203. 
Eny,  D.M.  (1951).  Biochem.  J.  50,  559. 
Eppley,  R.W.  (1958).  J.  Gen.  Physiol.  41,  901. 
Eppley,  R.W.  (1960).  Plant  Physiol.  35,  637. 

Erdos,  E.G.,  Debay,  C.R.,  and  Westerman,  M.P.  (1960).  Biochem.  Pharmacol.  5,  173. 
Erlenmeyer,  H.,  Waldi,  D.,  and  Sorkin,  E.  (1948).  Helv.  Chim.  Acta  31,  32. 
Ernster,  L.,  Danielson,  L.,  and  Ljunggren,  M.  (1962).  Biochim.  Biophys.  Acta  58, 

171. 
Estler,  C.-J.,  Heim,  F.,  Moosburger,  A.,  and  Herzog,  U.  (1960).  Arch.  Exptl.  Pathol. 

Pharmakol.  238,  301. 
Eubank,  L.L.,  Huf,  E.G.,  Campbell,  A.D.,  and  Taylor,  B.B.  (1962).  J.  Cellular  Comp. 

Physiol.  59,  129. 
Eusebi,  A.J.  and  Cerecedo,  L.R.  (1949).  Science  110,  162. 


Biol 

.  Chem. 

130. 

13. 

102, 

125. 

REFERENCES  1007 

Eusebi,  A.J.  and  Cerecedo,  L.R.  (1950).  Federation  Proc.  9,  169. 

Evans,  E.A.,  Jr.,  Vennesland,  B.,  and  Slotin,  L.  (1943).  J.  Biol.  Chem.  147,  771. 

Evans,  J.I.  and  Monk,  C.B.  (1952).  Trans.  Faraday  Soc.  48,  934. 

Evert,  H.E.  (1952).  Arch.  Biochem.  Biophys.  41,  29. 

Eyring,  H.  (1932).  Phys.  Rev.  39,  746. 

Ezaki,  S.  (1940).  J.  Biochem.  (Tokyo)  32,  107. 

Fahrlander,  H.,  Favarger,  P.,  Nielsen,  H.,  and  Leuthardt,  F.  (1947).  Helv.  Physiol. 

Pharmacol.  Acta  5,  202. 
Fahrlander,  H.,  Favarger,  P.,  and  Leuthardt,  F.  (1948).  Helv.  Chim.  Acta  31,  942. 
Fain,  J.N.  (1964).  J.  Biol.  Chem.  239,  958. 
Fairclough,  R.A.  (1938).  J.  Chem.  Soc.  p.  1186. 

Falcone,  G.,  Salvatore,  G.,  and  CoveUi,  I.  (1959).  Biochim.  Biophys.  Acta  36,  390. 
Fanestil,  D.D.,  Hastings,  A.B.,  and  Mahowald,  T.A.  (1963).  J.  Biol.  Chem.  238,  836. 
Farah,  A.  (1952).  Arch.  Exptl.  Pathol.  Pharmakol.  215,  29. 
Farah,  A.  and  Kruse,  R.  (1960).  J.  Pharmacol.  Exptl.  Therap. 
Farah,  A.  and  Mook,  W.  (1951).  J.  Pharmacol.  Exptl.  Therap. 
Farah,  A.  and  Rennick,  B.R.  (1954).  Federation  Proc.  13,  353. 
Farah,  A.  and  Rennick,  B.R.  (1956).  J.  Pharmacol.  Exptl.  Therap.  117,  478. 
Farah,  A.,  Mook,  W.,  and  Johnson,  R.  (1951).  Proc.  Soc.  Exptl.  Biol.  Med.  76,  403. 
Farah,  A.E.  (1957).  In  "MetaboHc  Aspects  of  Transport  Across  Cell  Membranes" 

(Q.R.  Murphy,  ed.),  p.  257.  Univ.  of  Wisconsin  Press,  Madison,  Wisconsin. 
Farah,  A.E.  and  Miller,  T.B.  (1962).  Ann.  Rev.  Pharmacol.  2,  269. 
Farkas,  G.L.  and  Ledingham,  G.A.  (1959).  Can.  J.  Microbiol.  5,  141. 
Farkas,  G.L.,  Konrad,  £.,  and  Kiraly,  Z.  (1957  a).  Naturwissenschaften  44,  65. 
Farkas,  G.L.,  Konrad,  fi.,  and  Kiraly,  Z.  (1957  b).  Physiol.  Plantarum  10,  346. 
Fasella,  P.  and  Hammes,  G.G.  (1963).  Arch.  Biochem.  Biophys.  100,  295. 
Fasold,  H.,  Groschel-Stewart,  U.,  and  Turba,  F.  (1964).  Biochem.  Z.  339,  487. 
Fastier,  F.N.  (1962).  Pharmacol.  Rev.  14.  37. 

Favelukes,  G.  and  Stoppani,  A.O.M.  (1958).  Biochim.  Biophys.  Acta  28,  654. 
Fawaz,  E.N.  and  Fawaz,  G.  (1962).  Biochem.  J.  83,  438. 
Fawaz,  E.N.,  Manoukian,  E.,  and  Fawaz,  G.  (1963).  Biochem.  Z.  337,  195. 
Fawaz,  G.  and  Fawaz,  E.N.  (1951).  Proc.  Soc.  Exptl.  Biol.  Med.  77,  239. 
Fawaz,  G.  and  Fawaz,  E.N.  (1954).  Proc.  Soc.  Exptl.  Biol.  Med.  87,  30. 
Fawaz,  G.,  Tutunji,  B.,  and  Fawaz,  E.N.  (1958).  Proc.  Soc.  Exptl.  Biol.  Med.  97,  770. 
Feigelson,  P.  and  Davidson,  J.D.  (1956  a).  Cancer  Res.  16,  352. 
Feigelson,  P.  and  Davidson,  J.D.  (1956  b).  J.  Biol.  Chem.  223,  65. 
Feigelson,  P.,  Williams,  J.N.,  Jr.,  and  Elvehjem,  C.A.  (1951).  J.  Biol.  Chem.  189, 

361. 
Feigelson,  P.,  Davidson,  J.D..  and  Robins.  R.K.  (1957).  J.  Biol.  Chem.  116,  993. 
Feig],  F.  (1960).  "Spot  Tests  in  Organic  Analysis,"  6th  ed.  Elsevier,  Amsterdam. 
Feinberg,  R.H.  and  Greenberg,  D.M.  (1959).  J.  Biol.  Chem.  234,  2670. 
Feinstein,  R.N.,  Cotter,  G.J.,  and  Hampton,  M.M.  (1954).  Am.  J.  Physiol.  177,  156. 
Feist,  F.  (1890).  Ann.  257,  253. 

Feldberg,  W.  and  Kellaway,  C.H.  (1938).  Australian  J.  Exptl.  Biol.  Med.  Sci.  16,  249. 
Feldman,  W.  and  O'Kane,  D.J.  (1960).  J.  Bacteriol.  80,  218. 
Felenbok,  B.  and  Neufeld,  E.F.  (1962).  Arch.  Biochem.  Bioj)hys.  97,  116. 
Feller,  D.D.  and  Feist,  E.  (1957).  Federation  Proc.  16,  36. 


1008  REFERENCES 

Fellig,  J.  and  Wiley,  C.E.  (1959).  Arch.  Biochem..  Biophys.  85,  313. 

Fellig,  J.  and  Wiley,  C.E.  (1960).  Science  132,  1835. 

Fellman,  J.H.  (1956).  Pr^oc.  Soc.  Exptl.  Biol.  Med.  93,  413. 

Fellman,  J.H.  (1959).  Enzymologia  20,  366. 

Felts,  J.M.,  Masoro,  E.J.,  Panagos,  S.S.,  and  Rapport,  D.  (1956).  Federation  Proc. 

15,  61. 
Fernley,  H.N.  (1962).  Biochem.  J.  82,  500. 
Ferno,  0.,  Fex,  H.,  Hogberg,  B.,  Linderot,  T.,  and  Rosenberg,  T.  (1953).  Acta  Chem. 

Scand.  7,  921. 
Ferrari,  R.A.  and  Arnold,  A.  (1963  a).  Biochim.  Biophys.  Acta  77,  349. 
Ferrari,  R.A.  and  Arnold,  A.  (1963  b).  Biochim..  Biophys.  Acta  77,  357. 
Ferrari,  R.A.,  Mandelstam,  P.,  and  Crane,  R.K.  (1959).  Arch.  Biochem.  Biophys.  80, 

372. 
Ferreira,  R.,  Ben-Zvi,  E.,  Yamane,  T.  Vasilevskis,  J.,  and  Davidson,  N.  (1961).  In 

"Advances  in  the  Chemistry  of  the  Coordination  Compounds,"  6th  Intern.  Conf. 

on  Coordination  Chem.,  p.  457.  Macmillan,  New  York. 
Festenstein,  G.N.  (1958).  Biochem.  J.  70,  49. 
Festenstein,  G.N.  (1959).  Biochem.  J.  72,  75. 

Fewson,  C.A.  and  Nicholas,  D.J.D.  (1961).  Biochim.  Biophys.  Acta  49,  335. 
Fewster,  J.A.  (1957).  Biochem.  J.  66,  9p. 
Fildes,  P.  (1940).  Brit.  J.  Exptl.  Pathol.  21,  67. 

Findlay,  J.,  Levvy,  G.A.,  and  Marsh,  C.A.  (1958).  Biochem.  J.  69,  467. 
Finkle,  B.J.  and  Smith,  E.L.  (1958).  J.  Biol.  Chem.  230,  669. 
Fischer,  A.  and  Herrmann,  H.  (1937).  Enzymologia  3,  180. 
Fischer,  F.G.  and  Eysenbach,  H.  (1937).  Ann.  530,  99. 
.^Fishbein,  W.N.  and  Bessman,  S.P.  (1964).  J.  Biol.  Chem.  239,  357. 
'Fisher,  F.M.,  Jr.  (1964).  Biol.  Bull.  126,  220. 
Fisher,  H.F.,  Krasna,  A.I.,  and  Rittenberg,  D.  (1954).  J.  Biol.  Chem.  209,  569. 
Fishgold,  J.T.  (1957).  Federation  Proc.  16,  296. 
Fishman,  R.A.  (1964).  Am.  J.  Physiol.  206,  836. 
Fishman,  W.H.  and  Green,  S.  (1957).  J.  Biol.  Chem.  225,  435. 
Fitzgerald,  R.J.  and  Jordan,  H.V.  (1953).  Antihiot.  Chemotherapy  3,  231. 
Fiume,  L.  (1960).  Nature  187,  792. 

Flaks,  J.G.  and  Cohen,  S.S.  (1959).  J.  Biol.  Chem.  234,  2981. 
Flamm,  W.G.  and  Crandall,  D.I.  (1963).  J.  Biol.  Chem.  238,  389. 
Flatow,  L.  (1928).  Biochem.  Z.  194,  132. 

Flavin,  M.  and  Slaughter,  C.  (1960).  J.  Biol.  Chem.  235,  1103. 
Flavin,  M.,  Ortiz,  P. J.,  and  Ochoa,  S.  (1955).  Nature  176,  823. 
Flesch,  P.  and  Kun,  E.  (1950).  Proc.  Soc.  Exptl.  Biol.  Med.  74,  249. 
Fhckinger,  R.A.,  Jr.  (1949).  J.  Exptl.  Zool.  112,  465. 

Florkin,  M.  and  Duchateua-Bosson,  G.  (1939).  Bull.  Soc.  Chim.  Biol.  21,  1204. 
Fluharty,  A.L.  and  Ballou,  C.E.  (1959).  J.  Biol.  Chem.  234,  2517. 
Fluri,  R.  (1959).  Arch.  Mikrobiol.  33,  195. 
Fodor,  P.J.  (1952).  Arch.  Biochem.  Biophys.  35,  311. 
Folk,  J.E.  (1956).  J.  Am.  Chem.  Soc.  78,  3541. 
Folk,  J.E.  and  Gladner,  J.A.  (1958).  J.  Biol.  Chem.  231,  379. 
Fonnum,  F.,  Haavaldsen,  R.,  and  Tangen,  0.  (1964).  J.  Neurochem.  11,  109, 
Forney,  R.B.  and  Harger,  R.N.  (1949).  Federation  Proc.  8,  292. 


REFERENCES  1009 

Forsen,  S.  and  Nilsson,  M.  (1961).  Arkiv  Kenii  17,  523. 

Forssman,  S.  (1941).  Acta  Physiol.  Scand.  2,  Suppl.  5. 

Forssman,  S.  and  Lindsten,  T.  (1946).  Acta  Pharmacol.  Toxicol.  2,  329. 

Forster,  R.P.  and  Goldstein,  L.  (1961).  J.  Cellular  Camp.  Physiol.  58,  247. 

Forster,  R.P.  and  Taggart,  J.V.  (1950).  J.  Cellular  Comp.  Physiol.  36,  251. 

Forti,  G.  (1957).  Physiol.  Plantarum  10,  898. 

Foss,  O.  (1961).  In  "Organic  Sulfur  Compounds"  (N.  Kharasch,  ed.).  Vol.  1,  p.  83 

Macmillan  (Pergamon),  New  York. 
Foster,  R.J.  (1961).  J.  Biol.  Chem.  236,  2461. 

Foster,  R.J.  and  Niemann,  C.  (1955a).  J.  Am.  Chem.  Soc.  77,  3365. 
Foster,  R.J.  and  Niemann,  C.  (1955b).  J.  Am.  Chem..  Soc.  77,  3370. 
Foster,  R.J.,  Shine,  H.J.,  and  Niemann,  C.  (1955).  J.  Am.  Chem.  Soc.  77,  2378. 
Foulerton,  A.G.R.  (1921).  J.  Pathol.  Bacteriol.  24,  257. 
Foulkes,  F.C.  and  Paine,  CM.  (1961).  J.  Biol.  Chem.  236,  1019. 
Fountain,  J.R.,  Hutchinson,  D.J.,  Waring,  G.B.,  and  Burchenal,  J.H.  (1953).  Proc. 

Soc.  Exptl.  Biol.  Med.  83,  369. 
Fourneau,  E.  and  Melville,  K.I.  (1931).  J.  Pharmacol.  Exptl.  Therap.  41,  21. 
Fournier,  P.,  d'Auzac,  J.,  Pujarniwcle,  S.,  and  Cuong,  T.C.  (1961).  Compt.  Rend.  252, 

796. 
Fraenkel-Conrat,  H.  (1955).  J.  Biol.  Chem.  217,  373. 
Fraenkel-Conrat,  H.  (1959).  In  "Sulfur  in  Proteins"  (R.  Benesch  et  al.,  eds.),  p.  339. 

Thomas,  Springfield,  Illinois. 
Francis,  M.J.O.,  Hughes,  D.E.,  Romberg,  H.L.,  and  Phizackerley,  P.J.R.  (1963). 

Biochem.  J.  89,  430. 
Franke,  W.  (1944  a).  Z.  Physiol.  Chem.  280,  76. 
Franks,  W.  (1944  b).  Z.  Physiol.  Chem.  281,  162. 

Fraser,  D.M.  and  Kermack,  W.O.  (1957).  Brit.  J.  Pharmacol.  12,  16. 
Freedland,  R.A.,  Wadzinski,  I.M.,  and  Waisman,  H.A.  (1961).  Biochem.  Biophys.  Res. 

Commun,  6,  227. 
Freedlander,  B.L.,  French,  F.A.,  and  Furst,  A.  (1956).  Proc.  Soc.  Exptl.  Biol.  Med. 

92,  533. 
French,  T.V.,  Dawid,  I.B.,  Day,  R.A.,  and  Buchanan,  J.M.  (1963  a).  J.  Biol.  Chem. 

238,  2171. 
French,  T.C,  Dawid,  I.B.,  and  Buchanan,  J.M.  (1963  b).  J.  Biol.  Chem.  238,  2186. 
Freudenberg,  K.  and  Eyer,  H.  (1932).  Z.  Physiol.  Chem.  213,  226. 
Friberg,  L.  (1956).  Acta  Pharmacol.  Toxicol.  12,  411. 
Friberg,  L.  (1959).  A.M.A.  Arch.  hid.  Healt  20,  42. 

Friberg,  L.,  Odeblad,  E.,  and  Forssman,  S.  (1957).  A.M.A.  Arch.  Ind.  Health  16,  163. 
Fridhandler,  L.  (1959).  Federation  Proc.  18,  48. 

Fridhandler,  L.  and  Quastel,  J.H.  (1955).  Arch.  Biochem.  Biophys.  56,  424. 
Fridhandler,  L.  Hafez,  E.S.E.,  and  Pincus,  G.  (1957).  Exptl.  Cell  Res.  13,  132. 
Fridovich,  I.  and  Handler,  P.  (1954).  Federation  Proc.  13,  212. 
Fridovich,  I.  and  Handler,  P.  (1956).  J.  Biol.  Chem.  223,  321. 
Fridovich,  I.  and  Handler,  P.  (1957).  J.  Biol.  Chem.  228,  67. 
Fridovich,  I.  and  Handler,  P.  (1958).  J.  Biol.  Chem.  231,  899. 
Fried,  G.H.  and  Antopol,  W.  (1957).  J.  Cellular  Comp.  Physiol  49,  201. 
Friedemann,  T.E.  (1934).  Science  80,  34. 
Frieden,  C  (1959).  J.  Biol.  Chem.  234,  809. 


1010  REFERENCES 

Frieden,  C.  (1962).  Biochim.  Biophys.  Acta  59,  484. 

Frieden,  C.  (1963).  J.  Biol.  Chem.  238,  3286. 

Frieden,  E.  and  Maggiolo,  I.W.  (1957).  Biochim.  Biophys.  Acta  24,  42. 

Frieden,  E.  and  Naile,  B.  (1954).  Arch.  Biochem.  Biophys.  48,  448. 

Frieden,  E.,  Westmark,  G.W.,  and  Schor,  J.M.  (1961).  Arch.  Biochem.  Biophys.  92, 

176. 
Friedland,  I.M.,  Kaplan,  L.A.,  Dietrich,  L.S.,  and  Shapiro,  D.M.  (1958).  Federation 

Proc.  17,  224. 
Friedmann,  H.C.  and  Vennesland,  B.  (1958).  J.  Biol.  Chem.  233,  1398. 
Friedmann,  H.C.  and  Vennesland,  B.  (1960).  J.  Biol.  Chem.  235,  1526. 
Frimmer,  M.  (1960).  Biochem.  Z.  332,  522. 
Frimmer,  M.  (1961).  Arch.  Exptl.  Pathol.  Pharmakol.  lAl,  96. 
Frisell,  W.R.  and  Hellerman,  L.  (1957).  J.  Biol.  Chem.  225,  53. 
Frisell,  W.R.  and  Mackenzie,  C.G.  (1955).  J.  Biol.  Chem.  217,  275. 
Frisell,  W.R.,  Weinbach,  E.G.,  and  Hellerman,  L.  (1949).  Federation  Proc.  8,  200. 
Frisell,  W.R.,  Lowe,  H.J.,  and  Hellerman,  L.  (1956).  J.  Biol.  Chem.  Ill,  75. 
Friz,  C.T.,  Lazarow,  A.,  and  Cooperstein,  S.J.  (1960).  Biol.  Bull.  119,  161. 
Frohman,  C.E.  and  Day,  H.G.  (1949).  J.  Biol.  Chem.  180,  93. 
Frohman,  C.E.,  Orten,  J.M.,  and  Smith,  A.H.  (1951).  J.  Biol.  Chem.  193,  277. 
Fromageot,  C.  and  Grand,  R.  (1944).  Enzymologia  11,  235. 
Fromm,  H.J.  and  Nordlie,  R.C.  (1959).  Arch.  Biochem.  Biophys.  81,  363. 
Fromni,  H.J.  and  Zewe,  V.  (1962).  J.  Biol.  Chem.  237.  3027. 
Frommel,  E.,  Herschberg,  A.D.,  and  Piquet,  J.  (1944).  Helv.  Physiol.  Pharmacol.  Acta 

2,  169. 
Fruton,  J.S.  and  Mycek,  M.J.  (1956).  Arch.  Biochem.  Biophys.  65,  11. 
Fuentes,  V.  and  Rubino,  R.  (1923).  Anales  Fac.  Med.  Montevideo  8,  360. 
Fujie,  Y.  (1959).  Japan.  J.  Pharmacol.  9,  6. 

Fujimoto,  D.  and  Ishimoto,  M.  (1961).  J.  Biochem,.  (Tokyo)  50,  533. 
Fuld,  M.  and  Paul,  M.H.  (1952).  Proc.  Soc.  Exptl.  Biol.  Med.  79,  355. 
Fulton,  J.D.  and  Christophers,  S.R.  (1938).  Ann.  Trop.  Med.  Parasitol.  32,  77. 
Fulton,  J.D.  and  Spooner,  D.F.  (1956  a).  Exptl.  Parasitol.  5,  59. 
Fulton,  J.D.  and  Spooner,  D.F.  (1956  b).  Exptl.  Parasitol.  5,  154. 
Fulton,  J.D.  and  Spooner,  D.F.  (1959).  Exptl.  Parasitol.  8,  137. 
Fulton,  J.D.  and  Spooner,  D.F.  (1960).  Exptl.  Parasitol.  9,  293. 
Futterman,  S.  (1957).  J.  Biol.  Chem.  218,  1031. 
Futterman,  S.  and  Silverman,  M.  (1957).  J.  Biol.  Chem.  114,  31. 

Gaebler,  O.H.  and  Beher,  W.T.  (1951).  J.  Biol.  Chem.  188,  343. 

GafFney,  T.J.,  Rosenberg,  H.,  and  Ennor,  A.H.  (1964).  Biochem.  J.  90,  170. 

Gage,  J.C.  and  Swan,  A.A.B.  (1961).  Biochem..  Pharmacol.  8,  77. 

Gajdos,  A.  (1939).  Compt.  Rend.  Soc.  Biol.  130,  1566. 

Gale,  G.R.  (1961).  Arch.  Biochem.  Biophys.  94,  236. 

Galoyan,  A.A.  (1959).  Tiolovye  Soedin.  v  Med.,   Ukr.  Nauchn.-Issled.  Sanit.-Khim. 

Inst.,  Tr.  Nauchn.  Konf.,  Kiev,  1957,  p.  79. 
Galoyan,  S.A.  and  Turpaev,  T.M.  (1958).  Dokl.  Akad.  Nauk  Arm.  SSR  27,  59. 
Galston,  A.W.,  Siegel,  S.M.,  and  Sharpensteen,  H.H.  (1955).  Federation  Proc.  14,  217. 
Gamble,  J.L.,  Jr.  (1957).  J.  Biol.  Chem.  228,  955. 
Gamble,  J.L.,  Jr.  and  Najjar,  V.A.  (1955).  J.  Biol.  Chem.  217,  595. 


REFERENCES  1011 

Gamborg,  O.L.,  Wetter,  L.R.,  and  Neish,  A.C.  (1961).  Can.  J.  Biochem.  Physiol.  39, 

1113. 
Gammon,  G.D.,  Gumnit,  R.,  Kanirin,  R.P.,  and  Kamrin,  A. A.  (1960).  In  "Inhibition  in 

the  Nervous  System  and  y-Aminobutyric  Acid"  (E.  Roberts,  ed.),  p.  328.  Macmillan 

(Pergamon),  New  York. 
Gane,  R.  and  Ingold,  O.K.  (1931).  J.  Chem.  Soc.  p.  2153. 
Garber,  N.  and  Citri,  N.  (1962).  Biochim.  Biophys.  Acta  62,  385. 
Garcia-Hernandez,  M.  and  Kun,  E.  (1957).  Biochim.  Biophys.  Acta  24,  78. 
Gardier,  R.W.  and  Woodbury,  R.A.  (1955).  J.  Pharmacol.  Exptl.  Therap.  114,  453. 
Gardner,  E.A.  and  Farah,  A.  (1954).  J.  Pharmacol.  Exptl.  Therap.  111,  255. 
Garen,  A.  and  Levinthal,  C.  (1960).  Biochim.  Biophys.  Acta  38,  470. 
Garfinkel,  D.  (1957).  Arch.  Biochem.  Biophys.  71,  100. 
Gause,  G.F.  (1933).  Protoplasma  17,  548. 

Gayer,  J.  and  Partowi,  R.  (1962).  Z.  Ges.  Exptl.  Med.  135,  419. 
Gelfant,  S.  (1960).  Exptl.  Cell  Res.  19,  72. 
Gelles,  E.  (1956).  J.  Chem.  Soc.  p.  4736. 
Gellhorn,  A.  and  Jones,  L.O.  (1949).  Blood  4,  60. 

Gemmill,  C.L.  and  Bowman,  E.M.  (1950).  J.  Pharmacol.  Exptl.  Therap.  100,  244. 
Gemmill,  C.L.  and  Hellerman,  L.  (1937).  Am.  J.  Physiol.  120,  522. 
Geppert,  J.  (1889).  Berlin.  Klin.  Wochschr.  26,  789. 

Gerarde,  H.W.,  Jones,  M.,  and  Winnick,  T.  (1952).  J.  Biol.  Chem.  196,  51. 
Gerardin.  C.  and  Kayser,  F.  (1959).  Compt.  Rend.  Soc.  Biol.  152.  1780. 
Gergely,  J.  (1955).  In  "Methods  in  Enzymology"  (S.P.  Colowick  and  N.O.  Kaplan, 

eds.).  Vol.  1,  p.  602.  Academic  Press,  New  York. 
Gergely,  J.  and  Maruyama,  K.  (1960).  J.  Biol.  Chem.  235,  3174. 
Gergely,  J.,  Martonosi,  A.,  and  Gouvea,  M.A.  (1959).  In  "Sulfur  in  Proteins"  (R. 

Benesch  et  al.,  eds.),  p.  297.  Academic  Press,  New  York. 
Gerhardt,  P.,  Levine,  H.B.,  and  Wilson,  J.B.  (1950).  J.  Bacteriol.  60,  459. 
Gerhart,  J.C.  and  Pardee,  A.B.  (1961).  Federation  Proc.  20,  224. 
Gerhart,  J.C.  and  Pardee,  A.B.  (1962).  J.  Biol.  Chem.  237,  891. 
Gerhart,  J.C.  and  Pardee,  A.B.  (1964).  Federation  Proc.  23,  727. 
Gershanovich,  V.N.  (1963).  Biochemistry  (URRS)  {English  Transl.)  27,  868. 
Gershenfeld,  L.  and  Witlin,  B.  (1950).  Ann.  N.  Y.  Acad.  Sci.  53,  172. 
Gessler,  U.  and  Bass,  L.  (1960).  Z.  Ges.  Exptl.  Med.  133,  18. 
Gessler,  U.  and  Kuner,  E.  (1960).  Z.  Ges.  Exptl.  Med.  133,  394. 
Gest,  H.,  Judis,  J.,  and  Peck,  H.D.,  Jr.  (1956).  In  "Inorganic  Nitrogen  Metabolism" 

(W.D.  McElroy  and  H.B.  Glass,  eds.),  p.  298.  Johns  Hopkins  Press,  Baltimore, 

Maryland. 
Geuther  (1866).  Z.  Chem.  2,  8.  (Cited  in  Dyson,  G.M.  (1950).  "A  Manual  of  Organic 

Chemistry,"  Vol.  1.  Longmans,  Green,  New  York.) 
Geyer,  R.P.  and  Cunningham,  M.  (1950).  J.  Biol.  Chem.  184,  641. 
Geyer,  R.P.,  Matthews,  L.W.,  and  Stare,  F.J.  (1950a).  J.  Biol.  Chem.  182,  101. 
Geyer,  R.P.,  Cunningham,  M.,  and  Pendergast,  J.  (1950  b).  J.  Biol.  Cheyn.  188,  185. 
Ghiretti,  F.,  Ghiretti-Magaldi.  A.,  and  Tosi,  L.  (1959).  J.  Gen.  Physiol.  Al,  1185. 
Ghosh,  A.K.  (1958).  Proc.  Intern.  Symp.  Enzyme  Chem.,  Tokyo  Kyoto  1957  p.  532. 

Maruzen,  Tokyo. 
Ghosh,  S.,  Blumenthal,  H.J.,  Davidson,  E.,  and  Roseman,  S.  (1960).  J.  Biol.  Chem. 

235,  1265. 


1012  REFERENCES 

Gibbins,  L.N.  and  Simpson,  F.J.  (1963).  Can.  J.  Microbiol.  9,  769. 

Gibbs,  M.  and  Calo,  N.  (1959  a).  Federation  Proc.  18,  235. 

Gibbs,  M.  and  Calo,  N.  (1959  b).  Plant  Physiol.  34,  318. 

Gibbs,  M.  and  Calo,  N.  (1960  a).  Federation  Proc.  19,  329. 

Gibbs,  M.  and  Calo,  N.  (1960  b).  Biochim.  Biophys.  Acta  44,  341. 

Gibbs,  M.,  Earl,  J.M.,  and  Ritchie,  J.L.  (1955).  Plant  Physiol.  30,  463. 

Gibson,  F.  and  Yanofsky,  C.  (1960).  Biochim..  Biophys.  Acta  43,  489. 

Gibson,  K.D.,  Laver,  W.G.,  and  Neuberger,  A.  (1958).  Biochem.  J.  70,  71. 

Gibson,  Q.H.  and  Roughton,  F.J.W.  (1955).  Proc.  Roy.  Soc.  B143,  310. 

Giebel,  O.  and  Passow,  H.  (1960).  Arch.  Oes.  Physiol.  T7^,  378. 

Giebisch,  G.  (1958).  J.  Cellular  Comp.  Physiol.  51,  221. 

Giebisch,  G.  (1960).  Circulation  21,  879. 

Giebisch,  G.  (1961).  J.  Oen.  Physiol.  44,  6.59. 

Giebisch,  G.  and  Dornian,  P.J.  (1958).  Proc.  Soc.  Exptl.  Biol.  Med.  98,  50. 

Gilbert,  D.A.  (1963).  Biochitn.  Biophys.  Acta  77,  657. 

GilfiUan,  R.F.,  Holtman,  D.F.,  and  Ross,  R.T.  (19.56).  J.  Bacteriol.  72,  620. 

Gillespie,  J.M.  (1956).  Australian  J.  Biol.  Sci.  9,  448. 

Gillette,  J.R.,  Brodie,  B.B.,  and  La  Du,  B.N.  (1957).  J.  Pharmacol.  Exptl.  Therap. 

119,  532. 
Gilman,  A.,  Philips,  F.S.,  Koelle,  E.S.,  Allen,  R.P.,  and  St.  John,  E.  (1946).  Am.  J. 

Physiol.  147,   115. 
Gilmour,  D.  (1960).  Nature  186,  295. 

Gilmour,  D.  and  Gellert,  M.  (1961).  Arch.  Biochem.  Biophys.  93,  605. 
Gilmour,  D.  and  Griffiths,  M.  (1957).  Arch.  Biochem.  Biophys.  72,  302. 
Gil  y  Gil,  C.  (1924).  Beitr.  Pathol.  Anat.  Allgem.  Pathol.  72,  621. 
Giovanelli,  J.  and  Stumpf,  P.K.  (1957).  Plant  Physiol.  32,  498. 
GiovanelH,  J.  and  Stumpf,  P.K.  (1958).  J.  Biol.  Chem.  231,  411. 
Girerd,  R.J.,  Tenebaum,  L.E.,  Berkowitz,  J.,  Rassaert,  C.L.,  and  Green,  D.M.  (1958). 

J.  Pharmacol.  Exptl.  Therap.  1 22,  25 AT 
Giri,  K.V.,  Krishnaswamy,  P.R.,  and  Rao.  N.A.  (1958).  Biochem.  J.  70,  66. 
Giuditta,  A.  and  Strecker,  H.J.  (1960).  Biochem.  Biophys.  Res.  Commun.  2,  159. 
Glaser,  L.  and  Brown,  D.H.  (1955).  J.  Biol.  Chem.  216,  67. 
Glasziou,  K.T.  (1956).  Australian  J.  Biol.  Sci.  9,  253. 

Glazko,  A.J.  and  Ferguson,  J.H.  (1940).  Proc.  Soc.  Exptl.  Biol.  Med.  45,  43. 
Glock,  G.E.  and  McLean,  P.  (1953).  Biochem.  J.  55,  400. 
Glynn,  I.M.  (1963).  J.  Physiol.  (London)  169,  452. 
Godeaux,  J.  (1944).  Bull.  Soc.  Roy.  Sci.  Liege  13,  219. 
Godzeski,  C.  and  Stone,  R.W.  (1955).  Arch.  Biochem.  Biophys.  59,  132. 
Goebel,  W.F.  and  Perlmann,  G.E.  (1949).  J.  Exptl.  Med.  89,  479. 
Goedkoop,  J.A.  and  MacGillavry,  C.H.  (1957).  Acta  Crrjst.  10,  125. 
Goerz,  R.D.  and  Pengra,  R.M.  (1961).  J.  Bacteriol.  81,  568. 
Goffart,  M.  and  Fischer,  P.  (1948).  Arch.  Intern.  Physiol.  55,  258. 
Gold,  A.H.  and  Segal,  H.L.  (1964).  Biochemistry  3,  778. 
Goldberg,  G.  (1963).  Unpublished  work. 

Goldberg,  L.,  Martin,  L.E.,  and  Leigh,  J.  (1962).  Biochem.  J.  85,  56. 
Goldberg,  L.I.,  DaCosta,  F.M.,  and  Ozaki,  M.  (1960).  Nature  188,  502. 
Goldman,  D.S.  (1956  a).  J.  Bacteriol.  71,  732. 
Goldman,  D.S.  (1956  b).  J.  Bacteriol.  72,  401. 


REFERENCES  1013 

Goldman,  D.S.  (1959  a).  Biochim.  Biophys.  Acta  32,  80. 

Goldman,  D.S.  (1959  b).  Biochim.  Biophys.  Acta  34,  527. 

Goldschmidt,  E.P.,  Yall,  I.,  and  Koffler,  H.  (1956).  J.  Bacterial.  72,  436. 

Goldstein,  A.  and  Doherty,  M.E.  (1951).  Arch.  Biochem.  Biophys.  33,  35. 

Goldstein,  L.,  Van  Baerle,  R.R.,  and  Dearborn,  E.H.  (1957).  Enzymologia  18,  261. 

Goldstein,  M.  and  Contrera,  J.F.  (1961).  Ezperientia  17,  267. 

Goldstein,  M.  and  Contrera,  J.F.  (1962).  J.  Biol.  Chem.  237,  1898. 

Goldstein,  M.H.,  Levitt,  M.F.,  Hauser,  A.D.,  and  Polimeros,  D.  (1961).  J.  Clin.  In- 
vest. 40,  731. 

Goldstone,  A.  and  Adams,  E.  (1962).  J.  Biol.  Chem.  237,  3476. 

Gonda,  O.,  Traub,  A.,  and  Avi-Dor,  Y.  (1957).  Biochem.  J.  67,  487. 

Gonnard,  P.  (1950).  Bull.  Soc.  Chim.  Biol.  32,  535. 

Gonnert,  R.  and  Bock,  M.  (1956).  Arzneimittel-Forsch.  6,  522. 

Gooder,  H.  and  Happold,  F.C.  (1954).  Biochem.  J.  57,  369. 

Goodman,  D.S.  (1961).  J.  Biol.  Chem.  236,  2429. 

Goodman,  D.S.  and  Popjak,  G.  (1960).  J.  Lipid  Res.  1,  286. 

Goodman,  I.  and  Hiatt,  R.B.  (1964).  Biochem.  Pharmacol.  13,  871. 

Goodman,  I.,  Fonts,  J.R.,  Bresnick,  E.,  Menegas,  R.,  and  Hitchings,  G.H.  (1959). 
Science  130,  450. 

Gopinathan,  K.P.,  Sirsi,  M.,  and  Vaidyanathan,  C.S.  (1964).  Biochem.  J.  91,  277. 

Gordon,  J.J.  (1953).  Biochem.  J.  55,  812. 

Gordon,  J.J.  and  Quastel,  J.H.  (1948).  Biochem.  J.  42,  337. 

Gore,  M.B.R.  (1951).  Biochem.  J.  50,  18. 

Goth,  A.  (1959).  Am.  J.  Physiol.  197,  1056. 

Gots,  J.S.  and  Gollub,  E.G.  (1959).  Proc.  Soc.  Exptl.  Biol.  Med.  101,  641. 

Gottlieb,  D.  (1964).  Endeavour  23,  85. 

Gotto,  A.M.  and  Komberg,  H.L.  (1961).  Biochem.  J.  81,  273. 

Gourevitch,  A.,  Pursiano,  T.A.,  and  Lein,  J.  (1961).  Arch.  Biochem.  Biophys.  93,  283. 

Govorov,  N.P.  (1936).  Tr.  Vses.  Inst.  Eksperim.  Med.  1,  No.  2,  147. 

GraflBin,  A.L.,  Gray,  N.M.,  and  Bagley,  E.H.  (1952).  Bull.  Johns  Hopkins  Hosp.  91, 
137. 

Granick,  S.  (1958).  J.  Biol.  Chem.  Ill,  1101. 

Grant,  B.R.  and  Beevers,  H.  (1964).  Plant  Physiol.  39,  78. 

Grant,  P.T.  and  Sargent,  J.R.  (1960).  Biochem.  J.  76,  229. 

Grant,  P.T.  and  Sargent,  J.R.  (1961).  Biochem.  J.  81,  206. 

Grant,  W.M.  and  Kinsey,  V.E.  (1946).  J.  Biol.  Chem.  165,  485. 

Grassmann,  W.,  Klenk,  L.,  and  Peters-Mayr,  T.  (1935).  Biochem.  Z.  280,  307. 

Gray,  C.T.  (1952).  J.  Bacteriol.  63,  813. 

Gray,  S.J.  and  Felsher,  R.Z.  (1945).  Proc.  Soc.  Exptl.  Biol.  Med.  59,  287. 

Grazi,  E.,  De  Flora,  A.,  and  Pontremoli,  S.  (1960).  Biochem.  Biophys.  Res.  Commun. 
2,  121. 

Green,  A.A.  and  McElroy,  W.D.  (1955).  In  "Methods  in  Enzymology"  (S.P.  Colowick 
and  N.O.  Kaplan,  eds.).  Vol.  2,  p.  857.  Academic  Press,  New  York. 

Green,  D.E.  (1936).  Biochem.  J.  30,  2095. 

Green,  D.E.  (1960).  Ciba  Found.  Symp.,  Quinones  Electron  Transport  p.  130. 

Green,  D.E.  and  Brosteaux,  J.  (1936).  Biochem.  J.  30,  1489. 

Green,  D.E.  and  Crane,  F.L.  (1958).  Proc.  Intern.  Symp.  Enzyme  Chem.,  Tokyo  Kyoto 
1957  p.  275.  Maruzen,  Tokyo. 


1014  REFERENCES 

Green,  D.E.  and  Fleischer,  S.  (1960).  In  "Metabolic  Pathways"  (D.M.  Greenberg,  ed.) 

Vol.  1,  p.  41.  Academic  Press,  New  York. 
Green,  D.E.  and  Wakil,  S.J.  (1960).  In  "Lipide  Metabolism"  (K.  Bloch,  ed.),  p.  1. 

Wiley,  New  York. 
Green,  D.E.,  Dewan,  J.G.,  and  Leloir,  L.F.  (1937  a).  Biochem.  J.  31,  934. 
Green,  D.E.,  Needham,  D.M.,  and  Dewan,  J.G.  (1937  b).  Biochem.  J.  31,  2327. 
Green,  D.W.,  Ingram,  V.M.,  and  Perutz,  M.F.  (1954).  Proc.  Roy.  Soc.  A  225,  287. 
Green,  I.  and  Mommaerts,  W.F.H.M.  (1954).  J.  Biol.  Chem.  210,  695. 
Green,  J.W.  and  Parpart,  A.K.  (1953).  J.  Cellular  Camp.  Physiol.  42,  191. 
Green,  N.M.  and  Neurath,  H.  (1953).  J.  Biol.  Chem.  204,  379. 
Greenawalt,  J.W.,  WeibuU,  C.,  and  Low,  H.  (1962).  J.  Biol.  Chem.  237,  853. 
Greenbaum,  L.M.  and  Sherman,  R.  (1962).  J.  Biol.  Chem.  237,  1082. 
Greenberg,  D.M.  and  Winnick,  T.  (1940).  J.  Biol.  Chem.  135,  761. 
Greenberg,  D.M.,  Bagot,  A.E.,  and  Roholt,  D.A.,  Jr.  (1956).  Arch.  Biochem.  Biophys. 

62,  446. 
Greengard,  O.  and  Feigelson,  P.  (1962).  J.  Biol.  Chem.  237,  1903. 
Greengard,  P.  and  Straub,  R.W.  (1962).  J.  Physiol.  (London)  161,  414. 
Greenstein,  J.P.  (1938).  J.  Biol.  Chem.  125,  501. 
Greenstein,  J.P.  and  Edsall,  J.T.  (1940).  J.  Biol.  Chem.  133,  397. 
Greenstein,  J.P.  and  Jenrette,  W.V.  (1942).  J.  Biol.  Chem.  142,  175. 
Greenstein,  J.P.,  Thompson,  J.W.,  and  Jenrette,  W.V.  (1940).  J.  Natl.  Cancer  Inst. 

1,  367. 
Gregg,  D.C.  and  Nelson,  J.M.  (1940).  .7.  Am.  Chem.  Soc.  62,  2506. 
Gregg,  J.R.  and  Ornstein,  N.  (1953).  Biol.  Bull.  105,  466. 
Gregolin,  C.  and  Singer,  T.P.  (1963).  Biochim.  Biophys.  Acta  67,  201. 
Greif,  R.L.  (1960).  Federation  Proc.  19,  362. 
Greif,  R.L.  and  Jacobs,  G.S.  (1958).  Am.  J.  Physiol.  192,  599. 
Greif,  R.L.,  Sullivan,  W^J.,  Jacobs,  G.S.,  and  Pitts,  R.F.  (1956).  J.  Clin.  Invest.  35, 

38. 
Grein,  L.  and  Pfleiderer,  G.  (1958).  Biochem.  Z.  330,  433. 
Gremels,  H.  (1929).  Arch.  Exptl.  Pathol.  Pharmakol.  140,  205. 
Greville,  G.D.  (1936).  Biochem.  J.  30,  877. 
Greville,  G.D.  and  Mildvan,  A.S.  (1962).  Biochem.  J.  84,  21p. 
Greville,  G.D.  and  Needham,  D.M.  (1955).  Biochim.  Biophys.  Acta  16,  284. 
Greville,  G.D.  and  Tapley,  D.F.  (1960).  Biochem.  J.  76,  49p. 
Grey,  E.G.  (1924).  Proc.  Roy.  Soc.  B96,  156. 

Griffith,  G.C.,  Butt,  E.M.,  and  Walker,  J.  (1954).  Ann.  Internal  Med.  41,  501. 
Griffith,  W.H.  (1937).  Proc.  Soc.  Exptl.  Biol.  Med.  37,  279. 

Griffiths,  D.E.  and  Chaplain,  R.A.  (1962).  Biochem.  Biophys.  Res.  Commun.  8,  497. 
Griffiths,  D.E.,  Morrison,  J.F.,  and  Ennor,  A.H.  (1957).  Biochem.  J.  65,  153. 
Grisolia,  S.,  Caravaca,  J.,  and  Joyce,  B.K.  (1958).  Biochim.  Biophys.  Acta  29,  432. 
Gromet-Elhanen,  Z.   and  Avron,  M.   (1963).   Biochem.   Biophys.   Res.  Commun.  10, 

215. 
Gros,  F.  and  Naono,  S.  (1961).  I7i  "Protein  Biosynthesis"  (R.J.C.  Harris,  ed.),  p.  195. 

Academic  Press,  New  York. 
Gross,  R.E.  (1921).  Z.  Physiol.  Chem.  112,  236. 
Grossman,  L.  and  Kaplan,  N.O.  (1958  a).  J.  Biol.  Chem.  231,  717. 
Grossman,  L.  and  Kaplan,  N.O.  (1958  b).  J.  Biol.  Chem.  231,  727. 


REFERENCES  1015 

Grossowicz,  N.  and  Halpem,  Y.S.  (1956  a).  Nature  177,  623. 

Grossowicz,  N.  and  Halpern,  Y.S.  (1956  b).  Biochim.  Biophys.  Acta  20,  577. 

Groupe,  V.,  Engle,  C.G.,  GafFney,  P.E.,  and  Manaker,  R.A.  (1955).  Appl.  Microbiol. 

3,  333. 
Gruber,  CM.,  de  Berardinis,  C,  and  Erf,  L.A.  (1949).  Arch.  Intern.  Pharmacodyn. 

79,  461. 
Gruber,  W.,  Warzecha,  K.,  Pfleiderer,  G.,  and  Wieland,  T.  (1962).  Biochem.  Z.  336, 

107. 
Gruber,  W.,  Hohl,  R.,  and  Wieland,  T.  (1963).  Biochem.  Biophys.  Res.  Commun.  12, 

242. 
Grunert,  R.R.  (1960).  Arch.  Biochem..  Biophys.  86,  190. 

Grunert,  R.R.  and  Rohdenburg,  E.L.  (1960).  Arch.  Biochem.  Biophys.  86,  185. 
Guarino,  A.J.  and  Sable,  H.Z.  (1956).  Biochim.  Biophys.  Acta  20,  201. 
Gubler,  C.J.  (1958).  Federation  Proc.  17,  477. 
Gubler,  C.J.  (1961).  J.  Biol.  Chem.  236,  3112. 
Guest,  J.R.  (1960).  Biochem.  J.  76,  405. 

Guggenheim,  K.  and  Diamant,  E.J.  (1958).  Biochem.  J.  69,  56. 
Guha,  S.R.  (1962).  Enzymologia  lA,  310. 
Gulland,  J.M.  and  Farrar,  W.V.  (1944).  Nature  154,  88. 

Gumbmann,  M.  and  Tappel,  A.L.  (1962  a).  Arch.  Biochem.  Biophys.  98,  262. 
Gumbmann,  M.  and  Tappel,  A.L.  (1962  b).  Arch.  Biochem.  Biophys.  98,  502. 
Gunja,  Z.H.,  Manners,  D.J.,  and  Maung,  K.  (1961).  Biochem.  J.  81,  392. 
Gurd,  F.R.N,  and  Wilcox,  P.E.  (1956).  Advan.  Protein  Chem.  11,  311. 
Gurtner,  H.P.  (1961).  Helv.  Physiol.  Pharmacol.  Acta  Suppl.  9. 

Haarniann,  W.  (1943  a).  Biochem.  Z.  314,  1. 

Haarmann,  W.  (1943  b).  Biochem.  Z.  314,  18. 

Haas,  E.  (1944).  J.  Biol.  Chem.  155,  321. 

Haas,  H.T.A.  (1941).  Arch.  Exptl.  Pathol.  Pharmakol  197,  284. 

Haas,  W.J.,  Sizer,  I.W.,  and  Loofbourow,  J.R.  (1951).  Biochim.  Biophys.  Acta  6, 

601. 
Habeeb,  A.F.S.A.  (1960).  Can.  J.  Biochem.  Physiol.  38,  269. 
Hachisuka,  Y.,  Kato,  H.,  Asano,  N.,  and  Kuno,  T.  (1955).  J.  Bacterial.  69,  407. 
Hackett,  D.P.  (1958).  Plant  Physiol.  33,  8. 

Hackett,  D.P.,  Rice,  B.,  and  Schmid,  C.  (1960).  J.  Biol.  Chem.  235,  2140. 
Hackney,  F.M.V.  (1948).  Proc.  Linnean  Soc.  N.  S.  Wales  73,  439. 
Haft,  D.E.  and  Mirsky,  I.A.  (1952).  J.  Pharmacol.  Exptl.  Therap.  104,  340. 
Hagen,  U.  (1955).  Arch.  Exptl.  Pathol.  Pharmakol.  124,  193. 
Hagihira,  H.,  Lin,  E.C.C.,  Samiy,  A.H.,  and  Wilson,  T.H.  (1961).  Biochem.  Biophys. 

Res.  Commun.  4,  478. 
Hagihira,  H.,  Wilson,  T.H.,  and  Lin,  E.C.C.  (1963).  Biochim.  Biophys.  Acta  78,  505. 
Hagstrom,  B.E.  (1963).  Biol.  Bull.  124,  55. 
Hahn,  H.  and  Taeger,  H.  (1931).  Klin.  Wochschr.  10,  489. 
Hahn,  L.  and  Fekete,  J.  (1953).  Acta  Chem.  Scand.  7,  798. 
Hahn,  L.  and  Frank,  E.  (1953).  Acta  Chem.  Scand.  7,  806. 
Hahn,  M.  and  Remy,  E.  (1922).  Deut.  Med.  Woschschr.  48,  793. 
Haining,  J.L.,  Fukui,  T.,  and  Axelrod,  B.  (1960).  J.  Biol.  Chem.  235,  160. 
Hakala,  M.T.,  Glaid,  A.J.,  and  Schwert,  G.W\  (1953).  Federation  Proc.  12,  213. 


1016  REFERENCES 

Halasz,  P.,  Mechler,  F.,  Feher,  0.,  and  Damjanovich,  S.  (1960).  Acta  Physiol.  Acad. 

Sci.  Hung.  18,  47. 
Halenz,  D.R.,  Feng,  J.-Y.,  Hegre,  C.S.,  and  Lane,  M.D.  (1962).  J.  Biol.  Chem.  237, 

2140. 
Haley,  T.J.  (1945).  J.  Am.  Pharm.  Assoc,  Sci.  Ed.  36,  229. 
HaU,  A.N.,  Lea,  D.J.,  and  Rydon,  H.N.  (1962).  Biochem.  J.  84,  12. 
Hall,  G.A.,  Jr.  (1949).  J.  Am.  Chem.  Soc.  71,  2691. 
Halliday,  S.L.,  Sloboda,  A.,  Will,  L.W.,  and  Oleson,  J.J.  (1957).  Federation  Proc.  16, 

190. 
Hallman,  N.  (1940).  Acta  Physiol.  Scand.  1,  Suppl.  4. 
Halvorson,  H.  and  Ellias,  L.  (1958).  Biochim.  Biophys.  Acta  30,  28. 
Halvorson,  H.,  Spiegelman,  S.,  and  Hinman,  R.L.  (1955).  Arch.  Biochem.  Biophys. 

55,  512. 
Halvorson,  H.  and  Spiegelman,  S.  (1952).  J.  Bacteriol.  64,  207. 
Hamilton-Miller,  J.M.T.  and  Smith,  J.T.  (1964).  Nature  201,  999. 
Hamilton-Miller,  J.M.T.,  Smith,  J.T.,  and  Knox,  R.  (1964).  Nature  201,  867. 
Hampton,  A.  (1962).  Federation  Proc.  21,  370. 
Hampton,  A.  (1963).  J.  Biol.  Chem.  238,  3068. 

Hand,  W.C,  Edwards,  B.B.,  and  Caley,  E.R.  (1943).  J.  Lab.  Clin.  Med.  28,  1835. 
Handler,  P.  (1945).  J.  Biol.  Chem.  161,  53. 
Handler,  P.  and  Klein,  J.R.  (1942).  J.  Biol.  Chem.  143,  49. 
Handley,  C.A.  and  Lavik,  P.S.  (1950).  J.  Pharmacol.  Exptl.  Therap.  100,  115. 
Handley,  C.A.  and  Seibert,  R.A.  (1956).  J.  Pharmacol.  Exptl.  Therap.  117,  253. 
Handschumacher,  R.E.  (1960).  J.  Biol.  Chem.  235,  2917. 

Handschumacher,  R.E.  and  Pasternak,  C.A.  (1958).  Biochim,.  Biophys.  Acta  30,  451. 
Hanly,  V.F.,  Rowan,  K.S.,  and  Turner,  J.S.  (1952).  Australian  J.  Sci.  Res.  Ser.  B5,  64. 
Hannoun,  C.  (1952).  Compt.  Rend.  235,  1445. 
Hanson,  A.  (1958).  Naturwissenschaften  45,  423. 
Hanzlik,  P.J.  (1923  a).  J.  Ind.  Hyg.  4,  386. 
Hanzlik,  P.J.  (1923  b).  J.  Ind.  Hyg.  4,  448. 
Hanzlik,  P.J.  (1923  c).  J.  Pharmacol.  Exptl.  Therap.  20,  463. 
Happold,  F.C.  (1950).  Advan.  Enzymol.  10,  51. 
Harary,  I.  (1957).  Federation  Proc.  16,  192. 
Hargreaves,  A.B.  (1955).  Arch.  Biochem.  Biophys.  57,  41. 
Harley,  J.L.  and  Ap  Rees,  T.  (1959).  New  Phytologist  58,  364. 
Harper,  E.  and  Strecker,  H.J.  (1962).  J.  Neurochem.  9,  125. 
Harpur,  R.P.  and  Quastel,  J.H.  (1949).  Nature  164,  693. 
Harris,  J.  and  Hellerman,  L.  (1956).  In  "Inorganic  Nitrogen  Metabolism"  (W.D. 

McElroy  and  H.B.  Glass,  eds.),  p.  565.  Johns  Hopkins  Press,  Baltimore,  Maryland. 
Harris,  J.E.  (1957).  Am..  J.  Ophthalmol.  44,  Suppl.  p.  262. 
Harris,  J.O.,  Eisenstark,  A.,  and  Dragsdorf,  R.D.  (1954).  J.  Bacteriol.  68,  745. 
Harting,  J.  (1947).  Biol.  Bull.  93,  194. 
Hartman,  S.C.  (1963  a).  J.  Biol.  Chem.  238,  3024. 
Hartman,  S.C.  (1963  b).  J.  Biol.  Chem.  238,  3036. 

Hartman,  W.J.,  Akawie,  R.L,  and  Clark,  W.G.  (1955).  J.  Biol.  Chem.  216,  507. 
Hartmann,  K.-U.  and  Heidelberger,  C.  (1961).  J.  Biol.  Chem.  236,  3006. 
Hartshorne,  D.  and  Greenberg,  D.M.  (1964).  Arch.  Biochem..  Biophys.  105,  173. 
Harvey,  G.T.  and  Beck,  S.D.  (1953).  J.  Biol.  Chem.  201,  765. 


REFERENCES  1017 

Harvey,  S.C.  (1949).  J.  Biol.  Chem.  179,  435. 

Hashimoto,  T.,  Nakao,  M.,  and  Yoshikawa,  H.  (1960).  J.  Biochem.  (Tokyo)  48,  158. 

Haskins,  F.A.,  Tissieres,  A.,  Mitchell,  H.K.,  and  Mitchell,  M.B.  (1953).  J.  Biol.  Chem. 

200,  819. 
Haslam,  R.J.  and  Krebs,  H.A.  (1963).  Biochem.  J.  88,  566. 
Hass,  L.F.  and  Byrne,  W.L.  (1960).  J.  Am.  Chem.  Soc.  82,  947. 
HassaU,  K.A.  (1958).  Nature  181,  1273. 
Hasse,  K.  and  Schleyer,  H.  (1961).  Biochem.  Z.  334,  360. 
Hasselbach,  W.  (1953).  Z.  Naturforsch.  8,  212. 
Hatch,  M.D.  and  Stumpf,  P.K.  (1961).  J.  Biol.  Chem.  236,  2879. 
Hatch,  M.D.  and  Turner,  J.F.  (1960).  Biochem.  J.  76,  556. 
Hatch,  M.D.,  Pearson,  J.A.,  Millerd,  A.,  and  Robertson,  R.N.  (1959).  Australian  J. 

Biol.  Sci.  12,  167. 
Hatefi,  Y.,  Haavik,  A.G.,  and  Griffiths,  D.E.  (1961a).  Biochem.  Biophys.  Res.  Commun. 

4,  441. 
Hatefi,  Y.,  Haavik,  A.G.,  and  Jurtshuk,  P.  (1961b).  Biochim.  Biophys.  Acta  52,  106. 
Haugaard,  N.  (1946).  J.  Biol.  Chem.  164,  265. 
Haugaard,  N.  and  Marsh,  J.B.  (1952).  J.  Biol.  Chem.  194,  33. 
Hauge,  J.G.  (1956).  J.  Am.  Chem.  Soc.  78,  5266. 
Hauge,  J.G.  (1960).  Biochim.  Biophys.  Acta  45,  250. 
Haughton,  B.G.  and  King,  H.K.  (1958).  Biochem.  J.  70,  660. 
Haughton,  B.G.  and  King,  H.K.  (1961).  Biochem.  J.  80,  268. 
Haiirowitz,  F.  and  Tekman,  S.  (1947).  Biochim.  Biophys.  Acta  1,  484. 
Hawtrey,  A.O.  (1962).  Biochem.  J.  85,  293. 
Hayaishi,  0.  (1953).  J.  Am.  Chem.  Soc.  75,  4367. 
Hayaishi,  O.  (1954).  Federation  Proc.  13,  226. 
Hayaishi,  O.  (1955  a).  J.  Biol.  Chem.  215,  125. 
Hayaishi,  O.  (1955  b).  In  "Amino  Acid  Metabolism"  (W.D.  McElroy  and  H.B.  Glass, 

eds.),  p.  914.  Johns  Hopkins  Press,  Naltimore,  Maryland. 
Hayaishi,  O.  and  Kornberg,  A.  (1952).  J.  Biol.  Chem.  197,  717. 
Hayaishi,  0.,  Katagiri,  M.,  and  Rothberg,  S.  (1957).  J.  Biol.  Chem.  229,  905. 
Heald,  P.J.  (1953).  Biochem.  J.  55,  625. 
Heilbrunn,  L.V.  (1925).  Biol.  Bull.  49,  241. 

Heilbrunn,  L.V.  and  Wilson,  W.L.  (1955).  J.  Cellular  Comp.  Physiol.  46,  361. 
Heimann-Hollaender,  E.  and  Lichtenstein,  N.  (1954).  Nature  174,  559. 
Heinz,  F.,  Bertelsen,  K.,  and  Lamprecht,  W.  (1962).  Z.  Physiol.  Cheyn.  329,  222. 
Heinz,  R.  (1899).  Arch.  Pathol.  Anat.  Physiol.  155,  44. 
Hellerman,  L.  (1937).  Physiol.  Rev.  17,  454. 

Hellerman,  L.  (1939).  Cold  Spring  Harbor  Symp.  Quant.  Biol.  7,  165. 
Hellerman,  L.  and  Chinard,  F.P.  (1955).  In  "Methods  of  Biochemical  Analysis"  (D. 

GUck,  ed.).  Vol.  1,  p.  1.  Wiley  (Interscience),  New  York. 
Hellerman,  L.  and  Newman,  M.D.  (1932).  J.  Am.  Chem.  Soc.  54,  2859. 
Hellerman,  L.  and  Perkins,  M.E.  (1934).  J.  Biol.  Chem.  107,  241. 
Hellerman,  L.,  Chinard,  F.P.,  and  Ramsdell,  P.A.   (1941).  J.  Am.  Chem.  Soc.  63, 

2551. 
Hellerman,  L.,  Chinard,  F.P.,  and  Deitz,  V.R.  (1943).  J.  Biol.  Chem.  147,  443. 
Hellerman,  L.,  Lindsay,  A.,  and  Bovarnik,  M.R.  (1946).  J.  Biol.  Chem.  163,  553. 
HeUerman,  L.,  Schellenberg,  K.A.,  and  Reiss,  O.K.  (1958).  J.  Biol.  Chem.  233,  1468. 


1018  REFERENCES 

Hellerman,  L.,  Reiss,  O.K.,  Parmar,  S.S.,  Wein,  J.,  and  Lasser,  N.L.  (1960).  J.  Biol. 

Chem.  235,  2468. 
Hellerman,  L.,  Reiss,  O.K.,  and  Gey,  M.K.  (1962).  J.  Biol.  Chem.  lil,  997. 
Hellig,  H.  and  Popjak,  G.  (1961).  J.  Lipid  Res.  2,  235. 
Helmreich,  E.  and  Eisen,  H.N.  (1959).  J.  Biol.  Chem.  234,  1958. 
Hemker,  H.C.  and  Hiilsmann,  W.C.  (1960).  Biochim.  Biophys.  Acta  44,  175. 
Henderson,  J.F.  (1962).  J.  Biol.  Chem.  237,  2631. 
Henderson,  J.H.M.  and  Stauffer,  J.F.  (1944).  Am.  J.  Botany  31,  528. 
Hendley,  D.D.  and  Beers,  R.F.,  Jr.  (1959).  Federation  Proc.  18,  245. 
Hendley,  D.D.  and  Beers,  R.F.,  Jr.  (1961).  J.  Biol.  Chem.  236,  2050. 
Hendlin,  D.  and  Soars,  M.H.  (1951).  Federation  Proc.  10,  384. 
Hendlin,  D.,  Koditschek,  L.K.,  and  Soars,  M.H.  (1953).  J.  Bacterial.  65,  466. 
Hepler,  O.E.  and  Simonds,  J.P.  (1946).  Arch.  Pathol.  41,  42. 
Hepler,  O.E.,  Gurley,  H.,  and  Simonds,  J.P.  (1945).  Arch.  Pathol.  39,  133. 
Hepp,  P.  (1887).  Arch.  Exptl.  Pathol.  Pharmakol.  23,  91. 

Heppel,  L.A.,  Harkness,  D.R.,  and  Hilmoe,  R.J.  (1962).  J.  Biol.  Chem.  237,  841. 
Herbert,  D.,  Gordon,  H.,  Subrahmanyan,  V.,  and  Green,  D.E.  (1940).  Biochem.  J. 

34,   1108. 
Herbert,  E.  (1956).  J.  Cellular  Comp.  Physiol.  47,  11. 
Heredia,  C.F.  and  Medina,  A.  (1960).  Biochem.  J.  77,  24. 
Herken,  H.  and  Neuhoff,  V.  (1963).  Z.  Physiol.  Chem.  331,  85. 
Herner,  B.  (1944).  Acta  Physiol.  Scand.  8,  Suppl.  23. 
Herring,  P.T.  and  Hynd,  A.  (1928).  J.  Physiol.  (London)  66,  267. 
Herriott,  R.M.  (1937).  J.  Gen.  Physiol.  20,  335. 
Herriott,  R.M.  (1947).  J.  Gen.  Physiol.  31,  19. 

Herz,  R.,  Jr.,  Tapley,  D.F.,  and  Ross,  J.E.  (1961).  Biochim.  Biophys.  Acta  53,  273. 
Herzog,  R.O.  and  Betzel,  R.  (1911).  Z.  Physiol.  Chem.  74,  221. 
Hess,  R.  and  Pearse,  A.G.E.  (1963).  Biochim.  Biophys.  Acta  71,  285. 
Hess,  S.M.,  Connamacher,  R.H.,  Ozaki,  M.,  and  Udenfriend,  S.  (1961).  J.  Pharmacol. 

Exptl.  Therap.  134,  129. 
Hestrin,  S.  and  Avigad,  G.  (1958).  Biochem.  J.  69,  388. 
Heydeman,  M.T.  and  Azoulay,  E.  (1963).  Biochim.  Biophys.  Acta  77,  545. 
Heymann,  H.,  Rogach,  Z.,  and  Mayer,  R.L.  (1954).  J.  Am.  Chem.  Soc.  76,  6330. 
Heymann,  H.,  Gulick,  Z.R.,  De  Boer,  C.J.,  De  Stevens,  G.,  and  Mayer,  R.L.  (1958). 

Arch.  Biochem.  Biophys.  73,  366. 
Heymann,  H.,  Ginsberg,  T.,  Gulick,  Z.R.,  and  Mayer,  R.L.  (1959).  Proc.  Soc.  Exptl. 

Biol.  Med.  100,  279. 
Heymans,  J.F.  (1889).  Arch.  Phijsiol.  13,  168. 

Hiai,  S.,  Mori,  T.,  Hino,  S.,  and  Mori,  T.  (1957).  J.  Biochem.  {Tokyo)  44,  839. 
Hiatt,  A.J.  (1961).  Plant  Physiol.  36,  552. 

Hiatt,  H.H.,  Marks,  P.A.,  and  Shorr,  E.  (1953).  J.  Biol.  Chem.  204,  187. 
Hicks,  S.P.  (1953).  Arch.  Pathol.  55.  302. 
Hicks,  S.P.  (1955).  Am.  J.  Pathol.  31,  189. 

Hietanen,  S.  and  Sillen,  L.G.  (1952).  Acta  Chem.  Scand.  6,  747. 
Higgins,  E.S.  (1961).  Enzymolocjia  23,  176. 

Higgins,  E.S.,  Richert,  D.A.,  and  Westerfeld,  W.W.  (1956a).  Federation  Proc.  15,  274. 
Higgins,  E.S.,  Richert,  D.A.,  and  Westerfeld,  W.W.  (1956b).  Proc.  Soc.  Exptl.  Biol. 

Med.  92,  509. 


REFERENCES  1019 

Hildebrandt,  A.C.,  Riker,  A. J.,  and  Watertor,  J.L.  (1954).  Phytopathology  44,  422. 

Hill,  B.R.  (1956).  Cancer  Res.  16,  460. 

Hill,  R.L.  and  Smith,  E.L.  (1957).  J.  Biol.  Chern.  224,  209. 

Hilton,  J.L.  (1958).  Science  128,  1509. 

Hilz,  H.  and  Klempien,  E.J.  (1959).  Biochem.  Z.  331,  563. 

Hirade,  J.  (1952).  J.  Biochem.  (Tokyo)  39,  185. 

Hirade,  J.  and  Hayashi,  T.  (1953).  J.  Biochem.  (Tokyo)  40,  461. 

Hirashinia,  K.  (1934).  Folia  Pharmacol.  Japon.  19,  219. 

Hirashima,  K.  (1935).  Folia  Pharmacol.  Japon.  19,  333. 

Hird,  F.J.R.  and  Yates,  J.R.  (1961).  Biochem.  J.  80,  612. 

Hitchcock,  P.  (1946).  J.  Pharmacol.  Exptl.  Therap.  87,  55. 

Hitchings,  G.H.  (1960).  Clin.  Pharmacol.  Therap.  1,  570. 

Hoadley,  L.  (1930).  Biol.  Bull.  58,  123. 

Hoare,  J.L.  and  Speakman,  J.B.  (1963).  Mature  199,  1091. 

Hoberman,  H.D.  and  Rittenberg,  D.  (1943).  J.  Biol.  Chem.  147,  211. 

Hoch,  F.L.  and  Vallee,  B.L.  (1959).  In  ''Sulfur  in  Proteins"  (R.  Benesch  et  al.,  eds.) 
p.  245.  Academic  Press,  New  York. 

Hoch,  F.L.  and  Vallee,  B.L.  (1960).  Arch.  Biochem.  Biophys.  91,  1. 

Hoch,  G.E.,  Schneider,  K.C.,  and  Burris,  R.H.  (1960).  Biochim.  Biophys.  Acta  37,  273. 

Hochachka,  P.W.  and  Teal,  J.M.  (1964).  Biol.  Bull.  126,  274. 

Hochster,  R.M.  and  Chang,  V.M.  (1963).  Can.  J.  Biochem.  Physiol.  41,  1503. 

Hochster,  R.M.  and  Nozzolillo,  C.G.  (1960).  Can.  J.  Biochem.  Physiol.  38,  79. 

Hochster,  R.M.  and  Quastel,  J.H.  (1952).  Arch.  Biochem.  Biophys.  36,  132. 

Hochster,  R.M.  and  Quastel,  J.H.  (eds.)  (1963).  '"Metabolic  Inhibitors,"  Academic 
Press,  New  York.  Vol.  I. 

Hockenhull,  D.J.D.,  Fantes,  K.H.,  Herbert,  M.,  and  Whitehead,  B.K.  (1954a).  J. 
Gen.  Microbiol.  10,  353. 

Hockenhull,  D.J.D.,  Herbert,  M.,  Walker,  A.D.,  Wilkin,  G.D.,  and  Winder,  F.G. 
(1954  b).  Biochem.  J.  56,  73. 

Hockenhull,  D.J.D.,  A.shton,  G.C.,  Fantes,  K.H.,  and  Whitehead,  B.K.  (1954  c).  Bio- 
chim. J.  57,  93. 

Hofling,  E.,  Lieb,  H.,  and  Schoniger,  W.  (1952).  Monatsh.  Chem.  83,  60. 

Hofmann,  E.,  Konig,  S.,  and  Behne,  I.  (1962).  Naturwissenschaften  49,  40. 

Hofmann,  E.C.G.  (1955).  Biochem.  Z.  327,  273. 

Hofmann,  E.C.G.  and  Rapoport,  S.  (1957).  Biochem.  Z.  329,  437. 

Hofstee,  B.H.J.  (1949).  J.  Biol.  Chem.  179.  633. 

Hofstee,  B.H.J.  (1955).  J.  Biol.  Chern.  216,  235. 

Hofstee,  B.H.J.  (1960).  J.  Am.  Chem.  Soc.  82,  5166. 

Hofstee,  B.H.J.  (1961).  Biochem.  Biophys.  Res.  Commun.  4,  5. 

Hokfelt,  B.  and  Bydgeman,  S.  (1961).  Proc.  Soc.  Exptl.  Biol.  Med.  106,  537. 

Hokfelt,  B.  and  Hultquist,  G.  (1961).  Proc.  Soc.  Exptl.  Biol.  Med.  107,  66. 

Holden,  M.  and  Pirie,  N.W.  (1955).  Biochem.  J.  60,  53. 

Holland,  J.  and  Humphrey,  G.F.  (1953).  Australian  J.  Exptl.  Biol.  Med.  Sci.  31,  299. 

Holland,  J.F.  (1961).  Clin.  PJmrmacol.  Therap.  2,  374. 

HoUocher,  T.C.,  Jr.  and  Commoner,  B.  (1960).  Proc.  Natl.  Acad.  Sci.  U.  S.  46,  416. 

Holly,  F.W.,  Peel,  E.W.,  Mozingo,  R.,  and  Folkers,  K.  (1950).  J.  Am.  Chem.  Soc.  72, 
5416. 

Holmberg,  C.G.  (1939).  Biochem.  J.  33,  1901. 


1020  REFERENCES 

Holmes,  W.L.  (1956).  J.  Biol.  Chem.  223,  677. 

Holt,  A.  and  Wold,  F.  (1961).  J.  Biol.  Chem.  236,  3227. 

Holtkamp,  D.E.  and  Hill,  R.M.  (1951).  Arch.  Biochem.  Biophys.  34,  216. 

Holton,  F.A.  and  Colpa-Boonstra,  J.  (1960).  Biochem.  J.  76,  179. 

Holtz,  P.  and  Palm,  D.  (1964).  Pharmacol.  Rev.  16,  113. 

Holzer,  H.  (1956).  Medizinische  No.  15,  576. 

Holzer,  H.  and  Frank,  S.  (1958).  Angeiv.  Chem.  70,  570. 

Holzer,  H.  and  Holldorf,  A.  (1957).  Biochem.  Z.  329,  292. 

Holzer,  H.  and  Schneider,  S.  (1961).  Biochitn.  Biophys.  Acta  48,  71. 

Holzer,  H.,  Holzer,  E.,  and  Schultz,  G.  (1955a).  Biochem.  Z.  326,  385. 

Holzer,  H.,  Haan,  J.,  and  Schneider,  S.  (1955  b).  Biochem.  Z.  326,  451. 

Holzer,  H.,  Haan,  J.,  and  Pette,  D.  (1955  c).  Biochem.  R.  327,  195. 

Holzer,  H.,  Goedde,  H.W.,  and  Schneider,  S.  (1955  d).  Biochem.  Z.  327,  245. 

Hommes,  F.A.  (1963).  Biochim.  Biophys.  Acta  77,  183. 

Honda,  S.I.  (1955).  Plant  Physiol.  30,  402. 

Honda,  S.I.  (1957).  Plant  Physiol.  32,  23. 

Honda,  S.I.  and  Muenster,  A.-M.  (1961).  Plant  Physiol.  36,  105. 

Hopkins,  F.G.  and  Morgan,  E.J.  (1938).  Biochem.  J.  32,  611. 

Hopkins,  F.G.,  Morgan,  E.J.,  and  Lutwak-Mann,  C.  (1938).  Biochem.  J.  32,  1829. 

Hopper,  S.  and  Segal,  H.L.  (1964).  Arch.  Biochem.  Biophys.  105,  501. 

Hori,  K.  (1963).  J.  Biochem.  (Tokyo)  53,  354. 

Horio,  T.  and  Kamen,  M.D.  (1962  a).  Biochemistry  1,  144. 

Horio,  T.  and  Kamen,  M.D.  (1962  b).  Biochemistry  1,  1141. 

Horio,  T.  and  Okunuki,  K.  (1954).  J.  Biochem.  (Tokyo)  41,  717. 

Horning,  M.G.,  Martin,  D.B.,  Karmen,  A.,  and  Vagelos,  P.R.  (1960).  Biochem.  Bio- 
phys. Res.  Commun.  3,  101. 

Horowitz,  M.G.  and  Klotz,  I.M.  (1956).  Arch.  Biochem.  Biophys.  63,  77. 

Horwitt,  M.K.  (1940).  Science  92,  89. 

Horwitz,  L.  (1957).  J.  Cellular  Comp.  Physiol.  49,  437. 

Hoshino,  M.  (1959).  J.  Biochem.  (Tokyo)  46,  825. 

Hoskins,  D.D.,  Whiteley,  H.R.,  and  Mackler,  B.  (1962).  J.  Biol.  Chem.  237,  2647. 

Hosoya,  N.  and  Kawada,  N.  (1958).  J.  Biochem.  (Tokyo)  45,  363. 

Hosoya,  N.  and  Kawada,  N.  (1961).  J.  Biochem.  (Tyoko)  49,  728. 

Hosoya,  N.,  Kawada,  N.,  and  Matsumnra,  Y.  (1960).  J.  Biochem.  (Tokyo)  48,  262. 

Hosoya,  T.  (1960).  J.  Biochem.  (Tokyo)  48,  375. 

Hotchkiss,  R.D.  (1956).  Arch.  Biochem.  Biophys.  65,  302. 

Houck,  C.R.  (1942).  J.  Cellular  Comp.  Physiol.  20,  277. 

Houck,  J.C.  (1957  a).  Arch.  Biochem.  Biophys.  71,  336. 

Houck,  J.C.  (1957  b).  Biochim.  Biophys.  Acta  26,  649. 

Howe,  W.B.,  Melson,  G.L.,  Meridith,  C.H.,  Morrison,  J.R.,  Piatt,  M.H.,  and  Stro- 
minger,  J.L.  (1964).  J.  Pharmacol.  Exptl.  Therap.  143,  282. 

Huang,  H.T.  and  Niemann,  C.  (1952).  J.  Am.  Chem.  Soc.  74,  101. 

Hubbard,  J.S.  and  Durham,  N.N.  (1961).  J.  Bacteriol.  82,  361. 

Hudson,  M.T.  and  Woodward,  G.E.  (1958).  Biochim.  Biophys.  Acta  28,  127. 

Huennekens,  F.M.,  Basford,  R.E.,  and  Gabrio,  B.W.  (1955).  J.  Biol.  Chem.  213,  951. 

Huennekens,  F.M.,  Caifrey,  R.W.,  Basford,  R.E.,  and  Gabrio,  B.W.  (1^51  a,).  J.  Biol. 
Chem.  in,  261. 

Huennekens,  F.M.,  Felton,  S.P.,  and  Snell,  E.E.  (1957  b).  J.  Am.  Chem.  Soc.  79,  2258. 


REFERENCES  1021 

Huffaker,  R.C.  and  Wallace,  A.  (1961).  Biochim.  Biophys.  Acta  46,  403. 

Hughes,  A.F.W.  (1950).  Quart.  J.  Microscop.  Sci.  91,  251. 

Hughes,  C.  and  Spragg,  S.P.  (1958).  Biochem.  J.  70,  205. 

Hughes,  D.E.  and  Williamson,  D.H.  (1952).  Biochem.  J.  51,  45. 

Hughes,  W.J.,  Jr.  (1947).  J.  Am.  Ckem.  Soc.  69,  1836. 

Hughes,  W.L.,  Jr.  (1950).  Cold  Spring  Harbor  Symp.  Quant.  Biol.  14,  79. 

Hughes,  W.L.,  Jr.  (1957).  Ann.  N.  Y.  Acad.  Sci.  65,  454. 

Hughes,  W.L.,  Jr.  and  Dintzis,  H.M.  (1964).  J.  Biol.  Ckem.  239,  845. 

Hughes,  W.L.,  Jr.  and  Straessle,  R.  (1950).  J.  Am.  Chem.  Soc.  72,  452. 

Huisman,  T.H.J.  (1959).  In  "Sulfur  in  Proteins"   (R.  Benesch  et  al.,  eds.),  p.  153. 

Academic  Press,  New  York. 
Hiilsmann,  W.C.  (1961).  Nature  192,  1153. 

Hummel,  J.P.,  Anderson,  D.O.,  and  Patel,  C.  (1958).  J.  Biol.  Chem.  233,  712. 
Humphrey,  B.A.  and  Humphrey,  G.F.  (1948).  J.  Exptl.  Biol.  25,  123. 
Humphrey,  G.F.  (1947).  J.  Exptl.  Biol.  24,  352. 
Humphrey.  G.F.  (1950).  Australian  J.  Exptl.  Biol.  Med.  Sci.  28,  1. 
Humphrey,  G.F.  (1957).  Biochem.  J.  65,  546. 
Humphreys,  T.E.,  Newcomb,  E.H.,  Bokman,  A.H.,  and  Stumpf,  P.K.  (1954).  J.  Biol. 

Chem.  210,  941. 
Hunter,  A.  and  Downs,  C.E.  (1945).  J.  Biol.  Chem.  157,  427. 
Hunter,  D.,  Bomford,  R.R.,  and  Russell,  D.S.  (1940).  Quart.  J.  Med.  9,  193. 
Hunter,  F.E.,  Jr.  and  Ford,  L.  (1955).  J.  Biol.  Chem.  216,  357. 
Hunter,  F.E.,  Jr.,  Fink,  J.,  and  Hurwitz,  A.  (19.59  a).  Federation  Proc.  18,  251. 
Hunter,  F.E.,  Jr.,  Levy,  J.F.,  Fink,  J.,  Schutz,  B.,  Guerra,  F.,  and  Hurwitz,  A.  (1959b). 

J.  Biol.  Chem.  234,  2176. 
Hunter,  F.R.  (1960).  Exptl.  Parasitol.  9,  271. 
Hunter,  G.J.E.  (1953).  Biochem.  J.  55,  320. 
Hunter,  N.W.  (1958).  Physiol.  Zool.  31,  23. 
Hunter,  N.W.  and  Hunter,  R.  (1957).  Science  126,  1246. 
Hunter,  W.C.  (1929).  Ann.  Internal  Med.  2,  796. 
Hunter,  W.R.  (1949).  J.  Exptl.  'Biol.  26,  113. 
Hurwitz,  J.  (1952).  Biochim.  Biophys.  Acta  9,  496. 
Hurwitz,  J.  (1953).  J.  Biol.  Chem.  205,  935. 
Hurwitz,  J.  (1955  a).  J.  Biol.  Chem.  212,  757. 
Hurwitz,  J.  (1955  b).  J.  Biol.  Chem.  217,  513. 
Hurwitz,  J.,  Weissbach,  A.,  Horecker,  B.L.,  and  Smyrniotis,  P.Z.  (1956).  J.  Biol. 

Chem.  218,  769. 
Hurwitz,  J.,  Heppel,  L.A.,  and  Horecker,  B.L.  (1957).  J.  Biol.  Chem.  226,  525. 
Hurwitz,  L.  and  Chaffee,  E.  (1954).  Federation  Proc.  13,  369. 
Huszak,  S.  (1935).  Z.  Physiol.  Chem.  236,  66. 
Huszak,  S.  (1940).  Biochem.  Z.  303,  349. 
Huxley,  J.S.  (11928).  Am.  Naturalist  62,  363, 

Hylin,  J.W.  ahd  Matsumoto,  H.  (1964).  Toxicol.  Appl.  Pharmacol.  6,  168. 
Hynd,  A.  (1927).  Proc.  Roy.  Soc.  B101,  244. 


Ibsen,  K.H., 


oe,  E.L.,  and  McKee,  R.W.  (1958).  Biochim.  Biophys.  Acta  30,  384. 


Ibsen,  K.H.,  Coe,  E.L.,  and  McKee,  R.W.  (1962).  Cancer  Res.  22,  182. 


Ichihara,  A.,  I 


hihara,  E.A.,  and  Suda,  M.  (1960).  J.  Biochem.  (Tokyo)  48,  412. 


1022  REFERENCES 

Ichihara,  K.,  Yashimatsu,  H.,  and  Sakamoto,  Y.  (1956).  J.  Biochem.  (Tokyo)  43,  803. 

Igarasi,  H.  (1939).  J.  Agr.  Chem.  Sac.  Japan  15,  225. 

Iliffe,  J.  and  Myant,  N.B.  (1964).  Biochem.  J.  91,  369. 

Illingworth,  B.,  Jansz,  H.S.,  Brown,  D.H.,  and  Cori,  C.F.  (1958).  Proc.  Natl.  Acad. 
Sci.  U.S.  44,  1180. 

Imamoto,  F.,  Iwasa,  K.,  and  Okunuki,  K.  (1959).  J.  Biochem.  {Tokyo)  46,  141. 

Imshenetsky,  A. A.  and  Perova,  K.Z.  (1957).  Microbiology  {USSR)  {English  Transl.) 
26,  301. 

Inagaki,  T.  (1940).  J.  Biochem.  {Tokyo)  32,  57. 

Ingrahara,  R.C.  and  Visscher,  M.B.  (1936).  Am.  J.  Physiol.  114,  681. 

Ingram,  V.M.  (1955).  Biochem.  J.  59,  653. 

Irvin,  J.L.  and  Irvin,  E.M.  (1954).  J.  Biol.  Chem.  210,  45. 

Isaka,  S.  and  Aikawa,  T.  (1963).  Exptl.  Cell  Res.  30,  139. 

Ishikawa,  S.  and  Lehninger,  A.L.  (1962).  J.  Biol.  Chem..  237,  2401. 

Ishimoto,  M.  and  Fujimoto,  D.  (1961).  J.  Biochem.  {Tokyo)  50,  299. 

Ishimoto,  M.  and  Yagi,  T.  (1961).  J.  Biochem.  {Tokyo)  49,  103. 

Ishimoto,  M.,  Kondo,  Y.,  Kameyama,  T.,  Yagi,  T.,  and  Shiraki,  M.  (1958).  Proc.  In- 
tern. Syynp.  Enzyme  Chem.,  Tokyo  Kyoto  1957  p.  229.  Maruzen,  Tokyo. 

Isles,  T.E.  and  Jocelyn,  P.C.  (1963).  Biochem.  J.  88,  84. 

Ito,  E.,  Kondo,  S.,  and  Watanabe,  S.  (1955).  J.  Biochem.  {Tokyo)  42,  793. 

Ives,  D.H.,  Morse,  F.A.,  Jr.,  and  Potter,  V.R.  (1963).  J.  Biol.  Chem.  238,  1467. 

Iwatsubo,  M.  and  Labeyrie,  F.  (1962).  Biochim.  Biophys.  Acta  59,  614. 

Iwatsuka,  H.,  Kuno,  M.,  and  Maruyama,  M.  (1962).  Plant  Cell  Physiol.  {Tokyo)  3,  157. 

Izar,  G.  (1909).  Biochem.  Z.  11,  371. 

Jaattela,  A.J.  and  Paasonen,  M.K.  (1961).  Acta  Pharmacol.  Toxicol.  18,  95. 

Jackson,  D.E.  (1926  a).  J.  Pharmacol.  Exptl.  Therap.  27,  244. 

Jackson,  D.E.  (1926  b).  J.  Pharmacol.  Exptl.  Therap.  29,  471. 

Jackson,  P.C.,  Hendricks,  S.B.,  and  Vasta,  B.M.  (1962).  Plant  Physiol.  37,  8. 

Jacob,  H.S.  and  Jandl,  J.H.  (1962).  J.  Clin.  Invest.  41,  779. 

Jacobs,  E.E.,  Jacob,  M.,  Sanadi,  D.R.,  and  Bradley,  L.B.  (1956).  J.  Biol.  Chem.  113, 

147. 
Jacobs,  F.A.  (1958).  Federation  Proc.  17,  79. 

Jacobs,  F.A.  and  Hillman,  R.S.L.  (1958).  J.  Biol.  Chem.  232,  445. 
Jacobs,  F.A.,  Coen,  L.J.,  and  Hillman,  R.S.L.  (1960).  J.  Biol.  Chem.  235,  1372. 
Jacobs,  H.I.  (1950).  Ph.D.  thesis.  University  of  Southern  California. 
Jacobsohn,  K.  (1953).  Enzymologia  16,  113. 
Jacobson,  K.B.  (1959).  J.  Gen.  Physiol.  43,  323. 
Jacobson,  K.B.  (1960).  J.  Biol.  Chem.  235,  2713. 

Jacobson,  K.B.  and  Kaplan,  N.O.  (1957  a).  J.  Biophys.  Biochem.  Cytol.  3,  31. 
Jacobson,  K.B.  and  Kaplan,  N.O.  (1957  b).  J.  Biol.  Chem.  lib,  427. 
Jacoby,  G.A.  and  La  Du,  B.N.  (1964).  J.  Biol.  Chem.  239,  419. 
Jacoby,  M.  (1933).  Biochem.  Z.  161,  181. 
Jacquez,  J.A.  (1961).  Am.  J.  Physiol.  200,  1063. 
Jaenicke,  L.  and  Brode,  E.  (1961).  Biochem.  Z.  334,  342. 
Jagendorf,  A.T.  and  Avron,  M.  (1959).  Arch.  Biochem.  Biophys.  80,  246. 
Jahn,  F.  (1914).  Arch.  Exptl.  Pathol.  Pharmakol.  76,  16. 
Jakoby,  W.B.  (1954).  J.  Biol.  Chem.  207,  657. 


REFERENCES  1023 

Jakoby,  W.B.  and  Bonner,  D.M.  (1953).  J.  Biol.  Chem.  205,  709. 

James,  W.O.  and  Beavers,  H.  (1950).  New  Phijtologist  49,  353. 

Janacek,  K.  (1962).  Biochim.  Biophys.  Acta  56,  42. 

Jansen,  E.F.,  Nutting,  M.-D.F.,  Jang,  R.,  and  Balls,  A.K.  (1950).  J.  Biol.  Chem.  185, 

209. 
Jansen,  E.F.,  Curl,  A.L.,  and  Balls,  A.K.,  (1951).  J.  Biol.  Chem.  189,  671. 
Jarnefelt,  J.  (1962).  Biochim.  Biophys.  Acta  59,  643. 
Jecsai,  G.  and  Elodi,  P.  (1963).  Acta  Physiol.  Acad.  Sci.  Hung.  24,  29. 
Jedeikin,  L.A.  and  Weinhouse,  S.  (1954).  Arch.  Biochem.  Biophys.  50,  134. 
Jeffree,  G.M.  (1956).  Biochem..  J.  64,  49p. 
Jeffree,  G.M.  (1957).  Biochim.  Biophys.  Acta  24,  532. 
Jencks,  W.P.  and  Lipmann,  F.  (1957).  J.  Biol.  Chem.  225,  207. 
Jenerick,  H.P.  (1957).  J.  Cellular  Comp.  Physiol.  49,  171. 

Jenkins,  W.T.,  Yphantis,  D.A.,  and  Sizer,  I.W.  (1959).  J.  Biol.  Chem.  234,  51. 
Jensen,  E.V.,  Hospelhorn,  V.D.,  Tapley,  D.F.,  and  Huggins,  C.  (1950).  J.  Biol.  Chem. 

185,  411. 
Jermstad,  A.  and  Jensen,  K.B.  (1950).  Pharm.  Acta  Helv.  25,  209. 
Jermstad,  A.  and  Jensen,  K.B.  (1951).  Bull.  Soc.  Chivi.  Biol.  33,  258. 
Jervis,  E.L.  and  Smyth,  D.H.  (1960).  J.  Physiol.  (London)  151,  51. 
Joachimoglu,  G.  (1922).  Biochem.  Z.  130,  239. 
Jodrey,  L.H.  and  Wilbur,  K.M.  (1955).  Biol.  Bull.  108,  346. 
Johnson,  G.T.  and  Moos,  C.  ( 1956).  J.  Cellular  Comp.  Physiol.  48,  243. 
Johnson,  M.A.  and  Frear,  D.S.  (1963).  Phrjtochemistry  2,  75. 
Johnson,  M.P.  and  Bonner,  J.  (1956).  Physiol.  Plantarum  9,  102. 
Johnson,  W.,  Corte,  G.,  and  Jasmin,   R.   (1958).  Proc.  Soc.  Exptl.  Biol.  Med.   99, 

677. 
Johnson,  W.J.  (195.5).  Can.  J.  Biochem.  Physiol.  33,  107. 
Johnson,  W.J.  and  kcCoU.  J.D.  (1955).  Science  122,  834. 
Johnson,  W.J.  and  McColl,  J.D.  (1956).  Federation  Proc.  15,  284. 
Johnston,  R.L.  (1941).  J.  Lab.  Clin.  Med.  27,  303. 
Johnstone,  J.H.  and  Mitchell,  I.L.S.  (1953).  Biochem.  J.  55,  xvii. 
Johnstone,  R.M.  (1958).  Federation  Proc.  17,  250. 
Joklik,  W.K.  (1950a).  Australian  J.  Sci.  Res.  Ser.  B3,  28. 
Joklik,  W.K.  (1950  b).  Australian  J.  Exptl.  Biol.  Med.  Sci.  28,  331. 
Jolley,  R.L.,  CheldeHn,  V.H.,  and  Newburgh,  R.W.  (1958).  J.  Biol.  Chem.  233,  1289. 
Jones,  E.A.  and  Gutfreund.  H.  (1961).  Biochem.  J.  79,  608. 
Jones,  E.A.  and  Gutfreund,  H.  (1964).  Biochem.  J.  91,  Ic. 

Jones,  J.D.,  Hulme,  A.C.,  and  Wooltorton,  L.S.C.  (1964).  Phytochemistry  3,  201. 
Jones,  J.H.,  Foster,  C,  and  Henle,  W.  (1948).  Proc.  Soc.  Exptl.  Biol.  Med.  69,  454. 
Jones,  J.R.  (1946).  J.  Exptl.  Biol.  23,  298. 
Jones,  O.T.G.  (1963).  J.  Exj)tl.  Botany  14,  399. 
Joshi,  J.G.  and  Handler,  P.  (1962).  J.  Biol.  Chem.  237,  929. 
Jowett,  M.  and  Brooks,  J.  (1928).  Biochem.  J.  22,  720. 
Jowett,  M.  and  Quastel,  J.H.  (1935  a).  Biochem.  J.  29,  2143. 
Jowett,  M.  and  Quastel,  J.H.  (1935  b).  Biochem.  J.  29,  2159. 
Jowett,  M.  and  Quastel,  J.H.  (1935  c).  Biochem.  J.  29,  2181. 
Jowett,  M.  and  Quastel,  J.H.  (1937).  Biochem.  J.  31,  275. 
Joyce,  B.K.  and  Grisolia,  S.  (1938).  J.  Biol.  Chem.  233,  350. 


1024  REFERENCES 

Joyce,  C.R.B.,  Moore,  H.,  and  Weatherall,  M.  (1954).  Brit.  J.  Pharmacol,  9,  463. 
Jung,  F.  (1947).  Arch.  Exptl.  Pathol.  Pharmakol.  204,  139. 

Kagawa,  Y.,  Minakami,  S.,  and  Yoneyama,  Y.  (1959).  J.  Biochetn.  (Tokyo)  46,  771. 
Kahana,  S.E.,  Lowry,  O.H.,  Schulz,  D.W.,  Passonneau,  J.V.,  and  Crawford,  E.J. 

(1960).  J.  Biol.  Chem.  235,  2178. 
Kahn,  J.S.  and  Jagendorf,  A.T.  (1961).  J.  Biol.  Chem.  236,  940. 
Kaiser,  C.  (1960).  In  "Medicinal  Chemistry"  (A.  Burger,  ed.).  2nd  ed.,  p.  89.  Wiley 

(Interscience),  New  York. 
Kaiser,  E.  (1953).  Nature  171,  607. 

Kalckar,  H.M.,  Kjeldgaard,  N.O.,  and  Klenow,  H.  (1948).  J.  Biol.  Chem,.  174,  771. 
Kalckar,  H.M.,  Kjeldgaard,  N.O.,  and  Klenow,  H.  (1950).  Biochim.  Biophys.  Acta  5, 

586. 
Kaldor,  G.  (1960).  Proc.  Soc.  Exptl.  Biol.  Med.  105,  645. 
Kaldor,  G.  and  Gitlin,  J.  (1963).  Arch.  Biochem.  Biophys.  102,  216. 
Kallen,  R.G.  and  Lowenstein,  J.M.  (1962).  Arch.  Biochem.  Biophys.  96,  188. 
Kalnitsky,  G.  and  Tapley,  D.F.  (1958).  Biochem.  J.  70,  28. 
Kaltenbach,  J.P.  and  Kalnitsky,  G.  (1951a).  J.  Biol.  Chem.  192,  629. 
Kaltenbach,  J.P.  and  Kalnitsky,  G.  (1951b).  J.  Biol.  Chem.  192,  641. 
Kalyankar,  G.D.,  Krishnaswamy,  P.R.,  and  Sreenivasaya,  M.  (1952).  Current  Set, 

(India)  21,  220. 
Kamrin,  R.P.  and  Kamrin,  A.A.  (1961).  J.  Neurochem.  6,  219. 
Kanamori,  M.  and  Wixom,  R.L.  (1963).  J.  Biol.  Chem.  238,  998. 
Kandler,  O.  (1955).  Z.  Naturforsch.  10b,  38. 
Kann,  E.E.  and  Mills,  R.C.  (1955).  J.  Bacterial.  69,  659. 
Kantarjian,  A.D.  (1961).  Neurology  11,  639. 

Kaper,  J.M.  and  Houwing,  C.  (1962  a).  Arch.  Biochem.  Biophys.  96,  125. 
Kaper,  J.M.  and  Houwing,  C.  (1962  b).  Arch.  Biochem.  Biophys.  97,  449. 
Kaplan,  C.  (1959).  Nature  184,  1074. 

Kaplan,  E.H.,  Kennedy,  J.,  and  Davis,  J.  (1954).  Arch.  Biochem.  Biophys.  51,  47. 
Kaplan,  N.O.  (1959).  Ciba  Found.  Study  Group  2,  37. 
Kaplan,  N.O.  and  Ciotti,  M.M.  (1954).  J.  Biol.  Chem.  211,  431. 
Kaplan,  N.O.,  Colowick,  S.P.,  and  Nason,  A.  (1951).  J.  Biol.  Chem.  191,  473. 
Kaplan,  N.O.,  Goldin,  A.,  Humphreys,  S.R.,  Ciotti,  M.M.,  and  Venditti,  J.M.  (1954). 

Science  120,  437. 
Kaplan,  N.O.,  Ciotti,  M.M.,  and  Stolzenbach,  F.E.  (1956).  J.  Biol.  Chem.  221,  833. 
Kaplan,  N.O.,  Ciotti,  M.M.,  Hamolsky,  M.,  and  Bieber,  R.E.  (1960).  Science  131,  392. 
Kappner,  W.  (1961).  Protoplasma  53,  504. 

Karasek,  M.A.  and  Greenberg,  D.M.  (1957).  J.  Biol.  Chem.  Ill,  191. 
Karlsson,  J.L.  (1950).  J.  Biol.  Chem.  183,  .549. 
Karnofsky,  D.A.,  Stock,  C.C,  Ridgway,  L.P.,  and  Patterson,  P.A.  (1950).  J.  Biol. 

Chem.  182,  471. 
Karunairatnam,  M.C.  and  Levvy,  G.A.  (1949).  Biochem.  J.  44,  599. 
Kasamaki,  A.,  Sasaki,  S.,  and  Usami,  S.  (1963).  J.  Biochem.  (Tokyo)  53,  89. 
Kashket,  E.R.  and  Brodie,  A.F.  (1963  a).  Biochim.  Biophys.  Acta  78,  52. 
Kashket,  E.R.  and  Brodie,  A.F.  (1963  b).  J.  Biol.  Chem-.  238,  2564. 
Kassanis,  B.  and  Kleczkowski,  A.  (1944).  Biochem.  J.  38,  20. 
Katchalski,  E.,  Berger,  A.,  and  Neumann,  H.  (1954).  Nature  173,  998. 


REFERENCES  1025 

Kato,  I.,  Matsushima,  T.,  and  Akabori,  S.  (1960).  J.  Biochem.  (Tokyo)  48,  199. 

Katoh,  S.  and  Takamiya,  A.  (1963).  Arch.  Biochem.  Biophys.  102,  189. 

Katoh,  T.  (1963).  Plant  Cell  Physiol.  (Tokyo)  4,  13. 

Katsuya,  H.,  Kinoshita,  T.,  Doi,  Y.,  Hashizume,  T.,  and  Motoyama,  H.  (1963).  Na- 
ture 198,  497. 

Katyal,  J.M.  and  Gorin,  G.  (1959).  Arch.  Biochem.  Biophys.  82,  319. 

Katz,  A.M.  (1963).  Biochim.  Biophys.  Acta  71,  397. 

Katz,  A.M.  and  Mommaerts,  W.F.H.M.  (1962).  Biochim.  Biophys.  Acta  65,  82. 

Katz,  J.  and  Chaikoff,  I.L.  (1955).  J.  Am.  Chem.  Soc.  77,  2659. 

Katz,  S.  (1952).  J.  Am.  Chem.  Soc.  74,  2238. 

Katz,  S.  (1962).  Nature  194,  569. 

Katz,  S.  and  Santilli,  V.  (1962  a).  Federation  Proc.  21,  376. 

Katz,  S.  and  Santilli,  V.  (1962  b).  Biochim.  Biophys.  Acta  55,  621. 

Kaufman,  B.  and  Kaplan,  N.O.  (1961).  J.  Biol.  Chem.  236,  2133. 

Kaufman,  B.T.  (1964).  J.  Biol.  Chem.  239,  PC669. 

Kaufman,  B.T.  and  Kaplan,  N.O.  (1960).  Biochim.  Biophys.  Acta  39,  332. 

Kaufman,  S.  and  Neurath,  H.  (1949).  Arch.  Biochem.  21,  245. 

Kavanau,  J.L.  (1963).  Nature  198,  525. 

Kawano,  Y.  and  Kawabata,  S.  (1953).  J.  Ferment.  Technol.  (Japan)  31,  184. 

Kay,  CM.  and  Marsh,  M.M.  (1959).  Biochim..  Biophys.  Acta  35,  262. 

Kaziro,  Y.  (1959).  J.  Biochem.  (Tokyo)  46,  1523. 

Kaziro,  Y.,  Hass,  L.F.,  Boyer,  P.D.,  and  Ochoa,  S.  (1962).  J.  Biol.  Chem.  237,  1460. 

Kearney,  E.B.  (1952).  J.  Biol.  Chem.  194,  747. 

Kearney,  E.B.  (1955).  In  "Methods  in  Enzymology"  (S.P.  Colowick  and  N.O.  Kaplan, 
eds.).  Vol.  2,  p.  640.  Academic  Press,  New  York. 

Kearney,  E.B.  (1957).  J.  Biol.  Chem.  113,  363. 

Kearney,  E.B.  (1958).  Proc.  Intern.  Symp.  Enzyme  Chem.,  Tokyo  Kyoto  1957  p.  340. 
Maruzen,  Tokyo. 

Kearney,  E.B.  (1960).  J.  Biol.  Chem.  235,  865. 

Kearney,  E.B.  and  Singer,  T.P.  (1956).  J.  Biol.  Chem.  219,  963. 

Keech,  D.B.  and  Utter,  M.F.  (1963).  J.  Biol.  Chem.  238,  2609. 

Keilin,  D.  and  Hartree,  E.F.  (1936).  Proc.  Roy.  Soc.  B119,  114. 

Keilin,  D.  and  Hartree,  E.F.  (1940).  Proc.  Roy.  Soc.  B129,  277. 

Keilin,  D.  and  King,  T.E.  (1960).  Proc.  Roy.  Soc.  B152,  163. 

Keister,  D.L.  and  San  Pietro,  A.  (1963).  Arch.  Biochem.  Biophys.  103,  45. 

Keister,  D.L.,  San  Pietro,  A.,  and  Stolzenbach,  F.E.  (1960).  J.  Biol.  Chem.  235, 
2989. 

Keleti,  T.  and  Telegdi,  M.  (1959).  Acta  Physiol.  Acad.  Sci.  Hung.  16,  235. 

Keller,  D.M.  and  Lotspeich,  W.D.  (19.59  a).  J.  Biol.  Chem.  234,  987. 

Keller,  D.M.  and  Lotspeich,  W.D.  (1959  b).  J.  Biol.  Chem.  234,  991. 

Kellerman,  G.M.  (1959).  Biochim.  Biophys.  Acta  33,  101. 

Kelly,  S.  (1953).  Biol.  Bull.  104,  138. 

Kenney,  F.T.  (1959).  J.  Biol.  Chem.  234,  2707. 

Kenten,  R.H.  (1958).  Biochem.  J.  69,  439. 

Keodam,  J.C,  Steyn-Parve,  E.P.,  and  van  Rheenen,  D.L.  (1956).  Biochim.  Biophys, 
Acta  19,  181. 

Kerby,  G.P.  and  Eadie,  G.S.  (1953).  Proc.  Soc.  Exptl.  Biol.  Med.  83,  111. 

Kermack,  W.O.  and  Matheson,  N.A.  (1957).  Biochem.  J.  65,  48. 


1026  REFERENCES 

Kernan,  R.P.  (1963).  J.  Physiol.  {London)  169,  862. 

Kessler,  D.  and  Krause,  R.M.  (1963).  Proc.  Soc.  Exptl.  Biol.  Med.  114,  822. 

Kessler,  R.H.  (1960).  Clin.  Pharmacol.  Therap.  1,  723. 

Kessler,  R.H.,  Lozano,  R.,  and  Pitts,  R.F.  (1957  a).  J.  Pharmacol.  Exptl.  Therap.  121, 

432. 
Kessler,  R.H.,  Lozano,  R.,  and  Pitts,  R.F.  (1957  b).  J.  Clin.  Invest.  36,  656. 
Kessler,  R.H.,  Hierholzer,   K.,  Gurd,  R.S.,  and  Pitts,  R.F.  (1958).  Am.  J.  Physiol. 

194,  540. 
Keston,  A.S.  (1963).  Arch.  Biochem.  Biophys.  102,  306. 
Kharasch,  N.  (ed.)  (1961).  "Organic  Sulfur  Compounds,"  Vol.  1.  Macmillan  (Perga- 

mon).  New  York. 
Khomutov,  R.M.,  Karpeiskii,  M.Y.,  Severin,  E.S.,  and  Gnuchev,  N.V.  (1961).  Dokl. 

Akad.  Nauk  SSSR  140,  2. 
Kielley,  W.W.  and  Bradley,  L.B.  (1956).  J.  Biol.  Chem.  218,  653. 
Kielley,  W.W.  and  Kielley,  R.K.  (1951).  J.  Biol.  Chem.  191,  485. 
Kielley,  W.W.  and  KieUey,  R.K.  (1953).  J.  Biol.  Chem.  200,  213. 
Kiesow,  L.  (1959).  Z.  Naturforsch.  14b,  492. 
Kiesow,  L.  (1960  a).  Z.  Naturforsch.  15b,  174. 
Kiesow,  L.  (1960  b).  Z.  Naturforsch.  15b,  291. 
Kiesow,  L.  (1960  c).  Z.  Naturforsch.  15b,  293. 
Kiessling,  K.-H.  (1956).  Biochim.  Biophys.  Acta  20,  293. 
Kihara,  H.  and  Snell,  E.E.  (1957).  J.  Biol.  Chem.  226,  485. 
Kilbourne,  E.D.  (1959).  Nature  183,  271. 

Kilsheimer,  G.S.  and  Axelrod,  B.  (1957).  J.  Biol.  Chem.  227,  879. 
Kilsheimer,  G.S.  and  Axeb-od,  B.  (1958).  Nature  182,  1733. 
Kimmel,  J.R.  and  Smith,  E.L.  (1954).  J.  Biol.  Chem.  207,  515. 
Kimmel,  J.R.  and  Smith,  E.L.  (1957).  Advan.  Enzymol.  19,  267. 
Kimura,  H.  and  Kummerow,  F.A.  (1963).  Arch.  Biochem..  Biophys.  102,  86. 
Kimura,  T.  and  Tobari,  J.  (1963).  Biochim.  Biophys.  Acta  73,  399. 
Kimura,  Y.  and  Niwa,  T.  (1953).  Nature  171,  881. 
King,  T.E.  and  Cheldelin,  V.H.  (1954).  J.  Biol.  Chem.  208,  821. 
King,  T.E.  and  Cheldelin,  V.H.  (1956).  J.  Biol.  Chem.  220,  177. 
King,  T.E.  and  Howard,  R.L.  (1962).  J.  Biol.  Chem.  237,  1686. 
King,  T.E.,  Kawasaki,  E.H.,  and  Cheldelin,  V.H.  (1956).  J.  Bacteriol.  72,  418. 
Kini,  M.M.  and  Quastel,  J.H.  (1959).  Nature  184,  252. 
Kinsky,  S.C.  and  McElroy,  W.D.  (1958).  Arch.  Biochem.  Biophys.  73,  466. 
Kipnis,  D.M.  (1958).  Federation  Proc.  17,  254. 
Kipnis,  D.M.  and  Cori,  C.F.  (1959).  J.  Biol.  Chem.  234,  171. 
Kipnis,  D.M.  and  Cori,  C.F.  (1960).  J.  Biol.  Chem.  235,  3070. 
Kirkham,  D.S.  and  Flood,  A.E.  (1963).  J.  Gen.  Microbiol.  32,  123. 
Kirkman,  H.N.  (1962).  J.  Biol.  Chem.  237,  2364. 
Kishimoto,  U.  (1961).  Comp.  Biochem.  Physiol.  2,  81. 

Kistiakowsky,  G.B.  and  Shaw,  W.H.R.  (1953).  J.  Am.  Chem.  Soc.  75,  866. 
Kistiakowsky,  G.B.,  Mangelsdorf,  P.C,  Rosenberg,  A.J.,  and  Shaw,  W.H.R.  (1952). 

J.  Am.  Chem.  Soc.  74,  5015. 
Kistner,  S.  (1960).  Acta  Chem.  Scand.  14,  1389. 
Kit,  S.  and  Greenberg,  D.M.  (1951).  Cancer  Pes.  11,  495. 
Klein,  A.O.  and  Vishniac,  W.  (1961).  J.  Biol.  Chem.  236,  2544. 


REFERENCES  1027 

Klein,  J.R.  (1940).  J.  Biol.  Chem.  134,  43. 

Klein,  J.R.  (1953).  J.  Biol.  Chem.  205,  725. 

Klein,  J.R.  (1956).  Federation  Proc.  15,  290. 

Klein,  J.R.  (1957).  Arch.  Biochem.  Biophys.  67,  423. 

Klein,  J.R.  (1960).  Biochim.  Biophys.  Acta  37,  534. 

Klein,  J.R.  and  Kamin,  H.  (1941).  J.  Biol.  Chem.  138,  507. 

Klein,  J.R.  and  Olsen,  N.S.  (1947).  J.  Biol.  Chem.  170,  151. 

Klein,  M.,  Brewer,  J.H.,  Perez,  J.E.,  and  Day,  B.  (1948).  J.  Immunol.  59,  135. 

Klein,  W.  (1957).  Z.  Physiol.  Chem.  307,  254. 

Kleinfeld,  M.,  Stein,  E.,  and  Murphy,  B.  (1964).  Am.  J.  Physiol.  206,  975. 

Kleinzeller,  A.  (1961).  In  "Membrane  Transport  and  Metabolism"  (A.  Kleinzeller  and 

A.  Kotyk,  eds.),  p.  527.  Academic  Press,  New  York. 
Kleinzeller,  A.,  Kotyk,  A.,  and  Kovac,  L.  (1959).  Nature  183,  1402. 
Klemperer,  H.G.  (1955).  Biochem.  J.  60,  122. 
Klemperer,  H.G.  (1957).  Biochem.  J.  67,  381. 

Klimek,  J.W.,  Cavallito,  C.J.,  and  Bailey,  J.H.  (1948).  J.  Bacteriol.  55,  139. 
Kline,  D.L.  and  DeLuca,  H.A.  (1956).  Can.  J.  Biochem.  Physiol.  34,  429. 
Klingenberg,  M.  and  Biicher,  T.  (1961).  Biochem.  Z.  334,  1. 
Klingenberg,  M.  and  Schollmeyer,  P.  (1960).  Biochem.  Z.  333,  335. 
Klingman,  J.D.  and  Handler,  P.  (1958).  J.  Biol.  Chem.  I'il,  369. 
Klotz,  I.M.  and  Klotz,  T.A.  (1959).  In  "Sulfur  in  Proteins"  (R.  Benesch  e<  aZ.,  eds.), 

p.  127.  Academic  Press,  New  York. 
Klotz,  I.M.  and  Tietze,  F.  (1947).  J.  Biol.  Chem.  168,  399. 
Klotz,  I.M.,  Ayers,  J.,  Ho,  J.Y.C.,  Horowitz,  M.G.,  and  Heiney,  R.E.  (1958).  J.  Am. 

Chem.  Soc.  80,  2132. 
Knopfmacher,  H.P.  and  Salle,  A.J.  (1941).  J.  Oen.  Physiol.  24,  377. 
Kobayashi,  Y.  (1957).  Arch.  Biochem.  Biophys.  71,  352. 
Koch,  R.  (1881).  Mitt.  Gesundh.  1,  234. 

Kocholaty,  W.  and  Krejci,  L.E.  (1948).  Arch.  Biochem.  18,  1. 
Kocourek,  J.,  Ticha,  M.,  Jiracek,  V.,  and  Kostif,  J.  (1963).  Biochim.  Biophys.  Acta 

71,  497. 
Koedam,  J.C.  (1958).  Biochim.  Biophys.  Acta  29,  333. 
Koedam,  J.C,  Steyn-Parve,  E.P.,  and  van  Rheenen,  D.L.  (1956).  Biochim.  Biophys. 

Acta  19,  181. 
Koeppe,  O.J.,  Boyer,  P.D.,  and  Stulberg,  M.P.  (19.56).  J.  Biol.  Chem.  219,  569. 
Koike,  K.  and  Okui,  S.  (1964).  J.  Biochem.  [Tokyo)  55,  195. 
Koishi,  T.  (1959  a).  Japan.  J.  Pharmacol.  8,  101. 
Koishi,  T.  (1959  b).  Japan.  J.  Pharmacol.  8,  124. 

Koivusalo,  M.  and  Luukkainen,  T.  (1959).  Acta  Physiol.  Scand.  45,  283. 
Kolthoff,  I.M.  and  Anastasi,  A.  (1958).  J.  Am.  Chem.  Soc.  80,  4248. 
KolthofF,  I.M.,  Stricks,  W.,  and  Morren,  L.  (1954).  Anal.  Chem.  26,  366. 
Kominz,  D.R.  (1961).  Biochim.  Biophys.  Acta  51,  456. 
Kono,  M.  and  Quastel,  J.H.  (1962).  Biochem.  J.  85,  24. 
Kono,  T.  and  Colowick,  S.P.  (1961).  Arch.  Biochem.  Biophys.  93,  514. 
Korey,  S.  (1950).  Biochim.  Biophys.  Acta  4,  58. 
Korn,  E.D.  (1962).  J.  Biol.  Chem.  237,  3423. 
Kornberg,  H.L.  and  Krebs,  H.A.  (1957).  Nature  179,  988. 
Kornguth,  S.E.  and  Stahmann,  M.A.  (1960).  Arch.  Biochem.  Biophys.  91,  32. 


1028  REFERENCES 

Kostytschew,  S.  and  Berg,  V.  (1930).  Z.  Physiol.  Chem.  188,  133. 

Kotaka,  S.  (1963).  J.  Gen.  Physiol.  46,  1087. 

Koukol,  J.  and  Conn,  E.E.  (1961).  J.  Biol.  Chem.  236,  2692. 

Kovac,  L.  (1960).  Enzymologia  11,  27. 

Kovachevich,  R.  and  Wood,  W.A.  (1955).  J.  Biol.  Chem.  213,  745. 

Krahe,  E.  (1924).  Klin.  Wochschr.  3,  70. 

Krahl,  M.E.  and  Clowes,  G.H.A.  (1940).  J.  Gen.  Physiol.  23,  413. 

Kramer,  M.  and  Arrigoni-Martelli,  E.  (1960).  Arch.  Ezptl.  Pathol.  Pharmalcol.  238,  43. 

Kramer,  P.J.  (1951).  Plant  Physiol.  26,  30. 

Krane,  S.M.  and  Crane,  R.K.  (1959).  J.  Biol.  Chem.  234,  211. 

Krasna,  A.I.  (1961).  J.  Biol.  Chem.  236,  749. 

Krasna,  A.I.  and  Rittenberg,  D.  (1954).  Proc.  Natl.  Acad.  Sci.  U.S.  40,  225. 

Krasna,  A.I.,  Riklis,  E.,  and  Rittenberg,  D.  (1960).  J.  Biol.  Chem.  235,  2717. 

Krebs,  E.G.  and  Norris,  E.R.  (1949).  Arch.  Biochem.  24,  49. 

Krebs,  E.G.,  Graves,  D.J.,  and  Fischer,  E.H.  (1959).  J.  Biol.  Chem.  234,  2867. 

Krebs,  H.A.  (1930).  Biochem.  Z.  220,  289. 

Krebs,  H.A.  (1935).  Biochem.  J.  29,  1951. 

Krebs,  H.A.  and  Eggleston,  L.V.  (1940).  Biochem.  J.  34,  442. 

Krebs,  H.A.  and  Eggleston,  L.V.  (1944).  Biochem.  J.  38,  426. 

Krebs,  H.A.  and  Eggleston,  L.V.  (1948).  Biochim.  Biophys.  Acta  1,  319. 

Krebs,  H.A.  and  Johnson,  W.A.  (1948).  Tabulae  Biol.  19,  100. 

Krebs,  H.A.  and  Lowenstein,  J.M.  (1960).  In  "Metabolic  Pathways"  (D.M.  Greenberg 

ed.).  Vol.  1,  p.  129.  Academic  Press,  New  York. 
Krebs,  H.A.,  Salvin,  E.,  and  Johnson,  W.A.  (1938).  Biochem.  J.  32,  113. 
Krebs,  H.A.,  Gurin,  S.,  and  Eggleston,  L.V.  (1952).  Biochem.  J.  51,  614. 
Kreke,  C.W.,  Kroger,  M.H.,  and  Cook,  E.S.  (1949).  J.  Biol.  Chem.  180,  565. 
Kreke,  C.W.,  Schaefer,  M.A.,  Seibert,  M.A.,  and  Cook,  E.S.  (1950).  J.  Biol.  Chem. 

185,  469. 
Krisch,  K.  (1963).  Biochem.  Z.  337,  546. 

Kriszat,  G.  and  Runnstrom,  J.  (1952).  Exptl.  Cell  Res.  3,  500. 
Krueger,  A.P.  and  Baldwin,  D.M.  (1933).  J.  Gen.  Physiol.  17,  129. 
Krueger,  A.P.  and  Baldwin,  D.M.  (1934).  J.  Gen.  Physiol.  17,  499. 
Krueger,  R.  (1950).  Experientia  6,  106. 
Krueger,  R.C.  (1955).  Arch.  Biochem.  Biophys.  57,  52. 
Krusius,  F.-E.  (1940).  Acta  Physiol.  Scand.  Suppl.  3. 
Kubowitz,  F.  and  Ott,  P.  (1941).  Naturwissenschaften  39,  590. 
Kuby,  S.A.,  Noda,  L.,  and  Lardy,  H.A.  (1954).  J.  Biol.  Chem.  210,  65. 
Kuhn,  R.  (1923).  Z.  Physiol.  Chem.  129,  57. 
Kuhn,  R.  and  Beinert,  H.  (1943).  Ber.  76,  904. 
Kuhn,  R.  and  Beinert,  H.  (1947).  Ber.  80,  101. 
Kuhn,  R.  and  Desnuelle,  P.  (1938).  Z.  Physiol.  Chem.  251,  14. 
Kuhn,  R.  and  Franke,  W.  (1935).  Ber.  68,  1528. 

Kuhn,  R.,  Katz,  H.,  and  Franke,  W.  (1934).  Naturwissenschaften  11,  808. 
Kull,  F.C.,  Grimm,  M.R.,  and  Mayer,  R.L.  (1954).  Proc.  Soc.  Exptl.  Biol.  Med.  86, 

330. 
Kun,  E.,  Dechary,  J.M.,  and  Pitot,  H.C.  (1954).  J.  Biol.  Chem.  210,  269. 
Kun,  E.,  Grassetti,  D.R.,  Fanshier,  D.W.,  and  Featherstone,  R.M.  (1958).  Biochem. 

Pharmacol.  1,  207. 


REFERENCES  1029 

Kun,  E.,  Fanshier,  D.W.,  and  Grassetti,  D.R.  (1960).  J.  Biol.  Chem.  235,  416. 

Kunz,  H.A.  (1956).  Helv.  Physiol.  Pharmacol.  Acta  14,  411. 

Kunz,  W.  (1963).  Z.  Physiol.  Chem.  334,  128. 

Kuo,  M.-H.  and  Blumenthal,  H.J.  (1961).  Biochim.  Biophys.  Acta  52,  13. 

Kupiecki,  F.P.  and  Coon,  M.J.  (1959).  J.  Biol.  Chem.  234,  2428. 

Kupiecki,  F.P.  and  Virtanen,  A.I.  (1960).  Acta  Chem.  Scand.  14,  1913. 

Kuratomi,  K.  (1959).  J.  Biochem.  (Tokyo)  46,  453. 

Kuratomi,  K.  and  Fukunaga,  K.  (1963).  Biochim.  Biophys.  Acta  78,  617. 

Kurtz,  A.N.  and  Niemann,  C.  (1961).  Biochim.  Biophys.  Acta  50,  520. 

Kurup,  C.K.R.  and  Vaidyanathan,  C.S.  (1963).  Biochem.  J.  88,  239. 

Kuschinsky,  G.  and  Liillmann,  H.  (1954).  Arch.  Exptl.  Pathol.  Pharmakol.  11^,  299. 

Kuschinsky,  G.  and  Turba,  F.  (1950).  Naturwissenschaften  37,  425. 

Kuschinsky,  G.  and  Turba,  F.  (1951).  Biochim.  Biophys.  Acta  6,  426. 

Kuschinsky,  G.,  Mercker,  H.,  Westermann,  E.,  and  Liillmann,  H.  (1953).  Arch.  Exptl. 
Pathol.  Pharmakol.  217,  401. 

Kusunose,  M.,  Nagai,  S.,  Kusunose,  E.,  and  Yamamura,  Y.  (1956).  J.  Bacteriol.  72, 
754. 

Kusunose,  E.,  Kusunose,  M.,  Kowa,  Y.,  and  Yamamura,  Y.  (1960).  J.  Biochem.  (To- 
kyo) 47,  689. 

Kutscha,  W.  (1961).  Arch.  Ges.  Physiol.  273,  409. 

Kutscher,  W.  and  Hasenfuss,  M.  (1940).  Z.  Physiol.  Chem.  265,  181. 

Kutscher,  W.  and  Sarreither,  W.  (1940).  Z.  Physiol.  Chem.  265,  152. 

Kuttner,  R.  and  Wagreich,  H.  (1953).  Arch.  Biochem.  Biophys.  43,  80. 

Kvam,  D.C.  and  Parks,  R.E.,  Jr.  (1960).  J.  Biol.  Chem.  235,  2893. 

Kvamme,  E.  (1957).  Acta  Chem.  Scand.  11,  1091. 

Kvamme,  E.  (1958  a).  Acta  Physiol.  Scand.  42,  204. 

Kvamme,  E.  (1958  b).  Acta  Physiol.  Scand.  42,  219. 

Kvamme,  E.  (1958  c).  Acta  Physiol.  Scand.  A2,  231. 

Kvamme,  E.  (1958d).  Acta  Physiol.  Scand.  42,  239. 

Labbe,  R.F.  and  Hubbard,  N.  (1961).  Biochim.  Biophys.  Acta  52,  130. 

Labeyrie,  F.  and  Stachiewicz,  E.  (1961).  Biochim.  Biophys.  Acta  52,  136. 

Ladd,  J.N.  (1956).  Australian  J.  Biol.  Sci.  9,  92. 

Ladd,  J.N.  and  Nossal,  P.M.  (1954).  Australian  J.  Exptl.  Biol.  Med.  Sci.  32,  523. 

La  Du,  B.N.  and  Zannoni,  V.G.  (1955).  J.  Biol.  Chem.  217,  777. 

Lagnado,  J.R.  and  Sourkes,  T.L.  (1956).  Can.  J.  Biochem.  Physiol.  34,  1185. 

Laki,  K.  (1935).  Z.  Physiol.  Chem.  236,  31. 

Lakshmanan,  M.R.,  Vaidyanathan,  C.S.,  and  Cama,  H.R.  (1964).  Biochem.  J.  90,  569. 

Lambie,  A.T.  and  Robson,  J.S.  (1960).  Clin.  Sci.  20,  123. 

Lambooy,  J.P.  (1950).  J.  Am.  Chem.  Soc.  72,  5225. 

Lambooy,  J.P.  (1955).  Natl.  Vitamin  Found.  Nutr.  Symp.  Sec.  11,  p.  12. 

Landau,  B.R.,  Laszlo,  J.,  Stengle,  J.,  and  Burk,  D.  (1958).  J.  Natl.  Cancer  Inst.  21, 

485. 
Landau,  J.V.,  Marsland,  D.,  and  Zimmerman,  A.M.  (1954).  Anat.  Record.  120,  789. 
Landon,  E.J.  and  Carter,  C.E.  (1960).  J.  Biol.  Chem.  235,  819. 
Landon,  E.J.  and  Norris,  J.L.  (1963).  Biochim.  Biophys.  Acta  71,  266. 
Lands,  W.E.M.  and  Niemann,  C.   (1959).  J.  Am.  Chem.  Soc.  81,  2204. 
Lane,  M.  and  Brindley,  CO.  (1964).  Proc.  Soc.  Exptl.  Biol.  Med.  116,  57. 


1030  REFERENCES 

Lane,  M.,  Fahey,  J.L.,  Sullivan,  R.D.,  and  Zubrod,  C.G.  (1958).  J.  Pharmacol.  Ezptl. 

Therap.  122,  315. 
Lane,  M.,  Petering,  H.G.,  and  Brindley,  CO.  (1959).  J.  Natl.  Cancer  Inst.  22,  349. 
Lang,  C.A.  and  Nason,  A.  (1959).  J.  Biol.  Chem.  234,  1874. 
Lang,  K.  and  Bassler,  K.H.  (1953).  Biochem.  Z.  323,  456. 
Langdon,  R.G.  and  Weakley,  D.R.  (1957).  Federation  Proc.  16,  208. 
Lange,  C.F.,  Jr.  and  Kohn,  P.  (1961  a).  J.  Biol.  Chem.  236,  1. 
Lange,  C.F.,  Jr.  and  Kohn,  P.  (1961b).  Cancer  Pes.  21,  1055. 
Langer,  L.J.  and  Engel,  L.L.  (1958).  J.  Biol.  Chem.  233,  583. 
Lara,  F.J.S.  (1952).  J.  Bacteriol.  64,  279. 

Lardy,  H.A.  and  Philipps,  P.H.  (1943  a).  J.  Biol.  Chem.  148,  333. 
Lardy,  H.A.  and  Phillips,  P.H.  (1943  b).  J.  Biol.  Chem.  148,  343. 
Lardy,  H.A.  and  Phillips,  P.H.  (1945).  Arch.  Biochem.  6,  53. 
Lardy,  H.A.  and  Wellman,  H.  (1953).  J.  Biol.  Chem.  201,  357. 
Lardy,  H.A.,  Wiebelhaus,  V.D.,  and  Mann,  K.M.  (1950).  J.  Biol.  Chem.  187,  325. 
Laris,  P.O.,  Ewers,  A.,  and  Novinger,  G.  (1962).  J.  Cellular  Comp.  Physiol.  59,  145. 
Lamer,  J.  and  Gillespie,  R.E.  (1956).  J.  Biol.  Chem.  223,  709. 
Lamer,  J.  and  Schliselfeld,  L.H.  (1956).  Biochim.  Biophys.  Acta  20,  53. 
Larner,  J.,  Jackson,  W.T.,  Graves,  D.J.,  and  Stamer,  J.R.  (1956).  Arch.  Biochem. 

Biophys.  60,  352. 
Larrabee,  M.G.  and  Bronk,  D.W.  (1952).  Cold  Spring  Harbor  Symp.  Quant.  Biol.  17, 

245. 
Larson,  F.C.  and  Albright,  E.G.  (1961).  J.  Clin.  Invest.  40,  1132. 
Lascelles,  J.  (1956).  Biochem.  J.  62,  78. 

Lassen,  U.V.  and  Overgaard-Hansen,  K.  (1962).  Biochim.  Biophys.  Acta  57,  111. 
Laszio,  J.  Landau,  B.R.,  Wight,  K.,  and  Burk,  D.  (1958).  J.  Natl.  Cancer  Inst.  21,  475. 
Laszlo,  J.,  Humphreys,  S.R.,  and  Goldin,  A.  (1960).  J.  Natl.  Cancer  Inst.  24,  267. 
Laszio,  J.,  Harlan,  W.R.,  Klein,  R.F.,  Kirshner,  N.,  Estes,  E.H.,  Jr.,  and  BogdonofF, 

M.D.  (1961).  J.  Clin.  Invest.  40,  171. 
Lathe,  G.H.  and  Walker,  M.  (1958).  Biochem.  J.  70,  705. 
Laties,  G.G.  (1949a).  Arch.  Biochem.  20,  284. 
Laties,  G.G.  (1949  b).  Arch.  Biochem.  22,  8. 
Laties,  G.G.  (1953).  Phtjsiol.  Plantarum  6,  215. 
Laties,  G.G.  (1959  a).  Arch.  Biochem.  Biophys.  79,  364. 
Laties,  G.G.  (1959  b).  Arch.  Biochem.  Biophys.  79,  378. 
Laties,  G.G.  (1964).  Plant  Physiol.  39,  391. 

Latuasan,  H.E.  and  Berends,  W.  (1961).  Biochim.  Biophys.  Acta  52,  502. 
Launoy,  L.  and  Levaditi,  C.  (1913).  Compt.  Rend.  Soc.  Biol.  72,  653. 
Lazdunski,  M.  and  Ouellet,  L.  (1962).  Can.  J.  Biochem.  Physiol.  40,  1619. 
Lazzarini,  R.A.  and  Atkinson,  D.E.  (1961).  J.  Biol.  Chem.  236,  3330. 
Lazzarini,  R.A.  and  San  Pietro,  A.  (1964).  Arch.  Biochem.  Biophys.  106,  6. 
Leach,  S.J.  (1960).  Australian  J.  Chem.  13,  520. 
Lebrun,  J.  and  Delaunay,  A.  (1951).  Ann.  Inst.  Pasteur  80,  524. 
Leder,  I.G.  (1957).  J.  Biol.  Chem.  225,  125. 
Ledoux,  L.  (1953).  Biochim.  Biophys.  Acta  11,  517. 
Ledoux,  L.  (1954).  Biochim..  Biophys.  Acta  13,  121. 
Lee,  C.-P.,  Azzone,  G.F.,  and  Ernster,  L.  (1964).  Nature  201,  152. 
Lee,  H.A.,  Jr.  and  Abeles,  R.H.  (1963).  J.  Biol.  Chem.  238,  2367. 


REFERENCES  1031 

Lee,  J.S.  and  Lifson,  N.  (1951).  J.  Biol.  Chem.  193,  253. 

Lee,  Y.-P.  (1957).  J.  Biol.  Chem.  227,  999. 

Lee,  Y.-P.  (1960  a).  Biochim.  Biophys.  Acta  43,  18. 

Lee,  Y.-P.  (1960  b).  Biochim.  Biophys.  Acta  43,  25. 

Lees,  H.  and  Simpson,  J.R.  (1955).  Biochem.  J.  59,  i. 

Lees,  H.  and  Simpson,  J.R.  (1957).  Biochem.  J.  65,  297. 

LeFevre,  C.G.  and  LeFevre,  R.J.W.  (1937).  J.  Chem.  Soc.  p.  1088. 

LeFevre,  P.G.  (1947).  Biol.  Bull.  93,  224. 

LeFevre,  P.G.  (1948).  J.  Gen.  Physiol.  31,  505. 

LeFevre,  P.G.  (1954).  Symp.  Soc.  Exptl.  Biol.  8,  118. 

LeFevre,  P.G.  and  Davies,  R.I.  (1951).  J.  Gen.  Physiol.  34,  515. 

LeFevre,  P.G.  and  Marshall,  J.K.  (1958).  Am.  J.  Physiol.  194,  333. 

Legge,  J.W.  and  Turner,  A.W.  (1954).  Australian  J.  Biol.  Sci.  7,  496. 

Lehman,  I.R.  and  Nason,  A.  (1956).  J.  Biol.  Chem.  Ill,  497. 

Lehman,  I.R.,  Roussos,  G.G.,  and  Pratt,  E.A.  (1962  a).  J.  Biol.  Chem.  237,  819. 

Lehman,  I.R.,  Roussos,  G.G.,  and  Pratt,  E.A.  (1962  b).  J.  Biol.  Chem.  237.  829. 

Lehman,  R.A.,  King,  E.E.,  and  Taube,  H.  (1950).  J.  Pharmacol.  Exptl.  Therap.  99, 

149. 
Lehmann,  J.  (1938).  Skand.  Arch.  Physiol.  80,  237. 
Lehninger,  A.L.  (1946  a).  J.  Biol.  Chem.  164,  291. 
Lehninger,  A.L.  (1946  b).  J.  Biol.  Chem.  165,  131. 
Lehninger,  A.L.  (1949).  J.  Biol.  Chem.  178,  625. 
Lehninger,  A.L.  (1951).  In  "Phosphorus  MetaboHsm"  (W.D.  McElroy  and  H.B.  Glass, 

eds.).  Vol.  1,  p.  344.  Johns  Hopkins  Press,  Baltimore,  Maryland. 
Lehninger,  A.L.  and  Kennedy,  E.P.  (1948).  J.  Biol.  Chem.  173,  753. 
Lehinger,  A.L.,  Wadkins,  C.L.,  Cooper,  C,  Devlin,  T.M.,  and  Gamble,  J.L.,  Jr.  (1958). 

Science  128,  450. 
Leloir,  L.F.  and  Goldemberg,  S.H.  (1960).  J.  Biol.  Chem.  235,  919. 
Leloir,  L.F.,  Olavarria,  J.M.,  Goldemberg,  S.H.,  and  Carminatti,  H.  (1959).  Arch. 

Biochem.  Biophys.  81,  508. 
Lembeck,  F.  and  Resch,  H.  (1960).  Arch.  Exptl.  Pathol.  Pharmakol.  240,  210. 
Lenney,  J.F.  (1956).  J.  Biol.  Chem.  ll^,  919. 
Lenti,  C.  and  Grillo,  M.A.  (1955).  Z.  Physiol.  Chem.  302,  253. 
Lenz,  G.R.  and  Martell,  A.E.  (1964).  Biochemistry  3,  745. 
Lenzi,  F.  and  Caniggia,  A.  (1953).  Folia  Cardiol.  12,  No.  3  and  4. 
Lerner,  A.B.,  Fitzpatrick,  T.B.,  Calkins,  E.,  and  Summerson,  W.H.  (1951).  J.  Biol. 

Chem.  191,  799. 
Leshe,  J.,  Butler,  L.G.,  and  Gorin,  G.  (1962  a).  Arch.  Biochem.  Biophys.  99,  86. 
Leshe,  J.,  WilUams,  D.L.,  and  Gorin,  G.  (1962  b).  Anal.  Biochem.  3,  257. 
Lester,  G.  and  Bonner,  D.M.  (1952).  J.  Bacteriol.  63,  759. 
Letnansky,  K.  and  Seehch,  F.  (1960).  Z.  Krehsforsch.  64,  1. 

Leuschner,  F.,  Heim,  F.,  and  Poertzel,  H.  (1957).  Arch.  Intern.  Pharmacodyn.  110,  177. 
Levenberg,  B.,  Melnick,  I.,  and  Buchanan,  J.M.  (1957).  J.  Biol.  Chem.  225,  163. 
Levey,  H.A.  and  Szego,  CM.  (1955).  Am.  J.  Physiol.  182,  507. 
Levey,  S.  and  Sheinfeld,  S.  (1954).  Gastroenterology  T7,  625. 
Levvy,  G.A.  (1952).  Biochem.  J.  52,  464. 
Levvy,  G.A.  (1954).  Biochem.  J.  58,  462. 
Levvy,  G.A.  and  Marsh,  C.A.  (1952).  Biochem.  J.  52,  690. 


1032  REFERENCES 

Levvy,  G.A.  and  Marsh,  C.A.  (1960).  In  "The  Enzymes"  (P.D.  Boyer,  H.  Lardy,  and 
K.  Myrback,  eds.).  Vol.  4,  p.  397.  Academic  Press,  New  York. 

Levvy,  G.A.  and  Worgan,  J.T.  (1955).  Biochem.  J.  59,  451. 

Levvy,  G.A.,  Hay,  A. J.,  and  Marsh,  C.A.  (1957).  Biochem.  J.  65,  203. 

Levvy,  G.A.,  McAllan,  A.,  and  Hay,  A.J.  (1962).  Biochem.  J.  82,  225. 

Levy,  H.M.,  Sharon,  N.,  Ryan,  E.M.,  and  Koshland,  D.E.,  Jr.  (1962).  Biochim.  Bio- 
phys.  Acta  56,  118. 

Levy,  H.R.  and  Talalay,  P.  (1959).  J.  Biol.  Chem.  234,  2014. 

Levy,  R.I.,  Weiner,  I.M.,  and  Mudge,  G.H.  (1958).  J.  Clin.  Invest.  37,  1016. 

Lewin,  J.  (1951).  J.  Am.  Chem.  Soc.  73,  3906. 

Lewis,  S.E.  and  Slater,  E.G.  (1954).  Biochem.  J.  58,  207. 

Li,  C.H.  (1945).  J.  Am.  Chem.  Soc.  67,  1065. 

Li,  T.-K.,  Ulmer,  D.D.,  and  Vallee,  B.L.  (1962).  Biochemistry  1,  114. 

Li,  Y.,  Li,  S.C.,  and  Shetlar,  M.R.  (1963).  Arch.  Biochem.  Biophys.  103,  436. 

Lichtenstein,  N.,  Ross,  H.E.,  and  Cohen,  P.P.  (1953).  Nature  171,  45. 

Liebecq,  C.  and  Peters,  R.A.  (1949).  Biochim.  Biophys.  Acta  3,  215. 

Lieben,  F.  and  Bauminger,  B.  (1933a).  Biochem.  Z.  261,  374. 

Lieben,  F.  and  Bauminger,  B.  (1933  b).  Biochem.  Z.  261,  387. 

Lieber,  E.  and  Smith,  G.B.L.  (1939).  Chem.  Rev.  25,  213. 

Lieberman,  I.  (1956).  J.  Biol.  Chem.  223,  327. 

Lieberman,  M.  (1961).  Plant  Physiol.  36,  804. 

Lieberman,  M.  and  Biale,  J.B.  (19.56  a).  Plant  Physiol.  31,  420. 

Lieberman,  M.  and  Biale,  J.B.  (1956  b).  Plant  Physiol.  31,  425. 

Liener,  I.E.  (1961).  Biochim..  Biophys.  Acta  53,  332. 

Lifson,  N.  and  Stolen,  J.A.  (1950).  Proc.  Soc.  Exptl.  Biol.  Med.  74,  451. 

LindeU,  S.E.,  Nilsson,  K.,  Rocs,  B.-E.,  and  Westling,  H.  (1960).  Brit.  J.  Pharmacol. 
15,  351. 

Linderholm,  H.  (1952).  Acta  Physiol.  Scand.  27,  Suppl.  97. 

Linderstrom-Lang,  K.  and  Jacobsen,  C.F.  (1941).  J.  Biol.  Chem.  137,  443. 

Lindner,  R.C.,  Kirkpatrick,  H.C.,  and  Weeks,  T.E.  (1959).  Phytopathology  49,  802. 

Lindstrom,  E.S.,  Burris,  R.H.,  and  Wilson,  P.W.  (1949).  J.  Bacterial.  58,  313. 

Lineweaver,  H.  (1933).  J.  Biol.  Chem.  99,  575. 

Ling,  G.  and  Gerard,  R.W.  (1949).  J.  Cellular  Comp.  Physiol.  34,  413. 

Link,  G.K.K.,  Klein,  R.M.,  and  Barron,  E.S.G.  (1952).  J.  Exptl.  Botany  3,  216. 

Lipke,  H.  and  Kearns,  C.W^  (1959).  J.  Biol.  Chem.  234,  2129. 

Lipsett,  M.N.  (1964).  J.  Biol.  Chem.  239,  1250. 

Lister,  A.J.  (1956).  J.  Gen.  Microbiol.  14,  478. 

Litchfield,  J.H.  and  Ordal,  Z.J.  (1958).  Can.  J.  Microbiol.  4,  205. 

Livermore,  A.H.  and  Sealock,  R.R.  (1947).  J.  Biol.  Chem.  167,  699. 

Lloyd,  D.R.  and  Prince,  R.H.  (1961).  Proc.  Chem.  Soc.  p.  464. 

Lockwood,  W.H.  and  Benson,  A.  (1960).  Biochem.  J.  75,  372. 

Lodin,  Z.,  Janacek,  K.,  and  Miiller,  J.  (1963).  J.  Cellular  Comp.  Physiol.  62,  215. 

Loefer,  J.B.  (1951).  Cancer  Res.  11,  481. 

Loevenhart,  A.S.  and  Grove,  W.E.  (1909).  J.  Pharmacol.  Exptl.  Therap.  1,  289. 

Loevenhart,  A.S.  and  Grove,  W.E.  (1911).  J.  Pharmacol.  Exptl.  Therap.  3,  101. 

London,  M.,  McHugh,  R.,  and  Hudson,  P.B.  (1958).  Arch.  Biochem.  Biophys.  73,  72. 

Long,  W.K.  and  Farah,  A.  (1946).  J.  Pharmacol.  Exptl.  Therap.  88,  388. 

Loomis,  W.F.  and  Lipmann,  F.  (1948).  J.  Biol.  Chem.  173,  807. 


REFERENCES  1033 

Lorand,  L.  and  Rule,  N.G.  (1961).  Nature  190,  722. 

Lord,  C.F.  and  Husa,  W.J.  (1954).  J.  Am.  Pharm.  Assoc,  Sci.  Ed.  43,  438. 

Lorenz,  B.,  Lorenz,  R.,  and  Zollner,  N.  (1960).  Z.  Naturforsch.  15b,  62. 

Lotspeich,  W.D.  and  Peters,  R.A.  (1951).  Biochem.  J.  49,  704. 

Lotspeich,  W.D.  and  Woronkow,  S.  (1964).  Am.  J.  Physiol.  206,  391. 

Loureiro,  J.A.  de  and  Lito,  E.  (1946).  J.  Hyg.  44,  463. 

Loveless,  L.E.,  Spoeri,  E.,  and  Weisman,  T.H.  (1954).  J.  Bacteriol.  68,  637. 

Lovenberg,  W.,  Barchas,  J.,  Weissbach,  H.,  and  Udenfriend,  S.  (1963).  Arch.  Bio- 
chem. Biophys.  103,  9. 

Levtrup,  S.  and  Svennerholm,  L.  (1963).  Exptl.  Cell.  Res.  29,  298. 

Low,  B.W.  (1952).  J.  Am.  Chem.  Soc.  74,  4830. 

Low,  B.W.  and  Weichel,  E.J.  (1951).  J.  Am.  Chem.  Soc.  73,  3911. 

Low,  H.  (1959  a).  Biochim.  Biophys.  Acta  32,  1. 

Low,  H.  (1959  b).  Exptl.  Cell.  Res.  16,  456. 

Low,  H.,  Siekevitz,  P.,  Ernster,  L.,  and  Lindberg,  O.  (1958).  Biochim.  Biophys.  Acta 
29,  392. 

Low,  H.,  Lindberg,  O.,  and  Ernster,  L.  (1959).  Biochim.  Biophys.  Acta  35,  284. 

Lowry,  O.H.  (1943).  Biol.  Symp.  10,  233. 

Lowry,  O.K.,  Bessey,  O.A.,  and  Crawford,  E.J.  (1949  a).  J.  Biol.  Chem.  180,  389. 

Lowry,  O.H.,  Bessey,  O.A.,  and  Crawford,  E.J.  (1949  b).  J.  Biol.  Chem.  180,  399. 

Liick,  H.  (1957).  Biochem.  Z.  328,  411. 

Ludwig,  B.J.  and  Nelson,  J.M.  (1939).  J.  Am.  Chem.  Soc.  61,  2601. 

Luh,  B.S.  and  Phaff,  H.J.  (1954).  Arch.  Biochem.  Biophys.  48,  23. 

Lukomskaia,  I.S.  and  Rozenfel'd,  E.L.  (1958).  Biochemistry  (USSfi)  (English  Transl.) 
23,  244. 

Luteraan,  P.J.  (1953).  Compt.  Rend.  236,  2531. 

Lutwak-Mann,  C.  (1947).  Biochem.  J.  41,  19. 

Luzzato,  R.  and  Satta,  G.  (1910).  Arch.  Farmacol.  Sper.  8,  554. 

Lwoff,  A.  and  Audureau,  A.  (1947).  An7i.  Inst.  Pasteur  73,  517. 

Lwoff,  A.,  Audereau,  A.,  and  Cailleau,  R.  (1947).  Compt.  Rend.  11^,  303. 

Lynen,  F.  (1943).  Ann.  554,  40. 

Lynen,  F.,  Reichert,  E.,  and  Rueff,  L.  (1951).  Ann.  574,  1. 

Lyon,  I.  and  Geyer,  R.P.  (1954).  J.  Biol.  Chem.  208,  529. 

McCaUa,  T.M.  and  Foltz,  V.D.  (1941).  Trans.  Kansas  Acad.  Sci.  44,  46. 
McComb,  R.B.  and  Yushok,  W.D.  (1959).  Biochim.  Biophys.  Acta  34,  515. 
McComb,  R.B.  and  Yushok,  W.D.  (1964  a).  Cancer  Res.  24,  193. 
McComb,  R.B.  and  Yushok,  W.D.  (1964  b).  Cancer  Res.  24,  198. 
McCormick,  D.B.  (1962).  J.  Biol.  Chem.  237,  959. 
McCormick,  D.B.  (1964).  Nature  201,  295. 

McCormick,  D.B.  and  Snell,  E.E.  (1961).  J.  Biol.  Chem.  236,  2085. 
McCormick,  D.B.,  Gregory,  M.E.,  and  Snell,  E.E.  (1961).  J.  Biol.  Chem.  236,  2076. 
McCormick,  N.G.,  Ordal,  E.J.,  and  Whiteley,  H.R.  (1962).  J.  Bacteriol.  83,  887. 
McCrea,  F.D.  and  Meek,  W.J.  (1929).  J.  Pharmacol.  Exptl.  Therap.  36,  295. 
McCurdy,  H.D.,  Jr.  and  Cantino,  E.G.  (1960).  Plant  Physiol.  35,  463. 
McDaniel,  E.G.,  Hundley,  J.M.,  and  Sebrell,  W.H.  (1955).  J.  Nutr.  55,  623. 
McDonald,  J.K.,  Anderson,  J.A.,  Cheldelin,  V.H.,  and  King,  T.E.  (1963).  Biochim. 
Biophys.  Acta  73,  533. 


1034  REFERENCES 

Macdonald,  K.  (1961).  Biochem.  J.  80,  154. 

MacDonnell,  L.R.,  Silva,  R.B.,  Feeney,  R.E.  (1951).  Arch.  Biochem..  Biophys.  32,  288. 

Macdowall,  F.D.H.  (1949).  Plant  Physiol.  24,  462. 

McEwen,  B.S.,  Allfrey,  V.G.,  and  Mirsky,  A.E.  (1963  a).  J.  Biol.  Chem.  238,  2571. 

McEwen,  B.S.,  Allfrey,  V.G.,  and  Mirsky,  A.E.  (1963  b).  J.  Biol.  Chem.  238,  2579. 

McFadden,  B.A.  and  Atkinson,  D.E.  (1957).  Arch.  Biochem.  Biophys.  66,  16. 

McFadden,  B.A.  and  Howes,  W.V.  (1963).  J.  Biol.  Chem.  238,  1737. 

Macfarlane,  E.W.E.  (1951).  Growth  15,  241. 

Macfarlane,  E.W.E.  (1953).  Ezptl.  Cell  Ees.  5,  375. 

Macfarlane,  E.W.E.  and  Nadeau,  L.V.  (1948).  Proc.  Soc.  Exptl.  Biol.  Med.  67,  118. 

Macfarlane,  W.V.  and  Meares,  J.D.  (1958).  J.  Physiol.  {London)  142,  78. 

McGarrahan,  J.F.  and  Maley,  F.  (1962).  J.  Biol.  Chem.  237,  2458. 

McGowan,  J.C,  Brian,  P.W.,  and  Hemming,  H.G.  (1948).  Ann.  Appl.  Biol.  35,  25. 

McGuire,  J.S.,  Jr.,  Hollis,  V.W.,  Jr.,  and  Tomkins,  G.M.  (1960).  J.  Biol.  Chem.  235, 

3112. 
Machlis,  L.  (1944).  Am.  J.  Botany  31,  183. 
Macht,  D.I.  (1931  a).  Am.  J.  Botany  18,  598. 
Macht,  D.I.  (1931b).  Proc.  Soc.  Exptl.  Biol.  Med.  28,  687. 
Mcllwain,  H.  (1940).  Brit.  J.  Exptl.  Physiol.  21,  136. 
Mcllwain,  H.  (1941).  Brit.  J.  Exptl.  Physiol.  11,  148. 
Mcllwain,  H.  (1945).  Biochem.  J.  39,  329. 
Mcllwain,  H.  (1950).  Biochem.  J.  46,  612. 
Mcllwain,  H.  (1959).  Biochem..  J.  71,  281. 
Mcllwain,  H.  and  Rodnight,  R.  (1949).  Biochem.  J.  44,  470. 
Mackay,  D.  and  Shepherd,  D.M.  (1962).  Biochim.  Biophys.  Acta  59,  553. 
McKinney,  G.R.  and  Gemma,  F.E.  (1958).  J.  Pharmacol.  Exptl.  Therap.  122,  51A. 
McKinney,  G.R.,  Smith,  L.C.,  and  Martin,  S.P.  (1955).  J.  Pharmacol.  Exptl.  Therap. 

113,  37. 
Mackler,  B.  (1961).  Biochim.  Biophys.  Acta  50,  141. 

Mackler,  B.,  Mahler,  H.R.,  and  Green,  D.E.  (1954).  J.  Biol.  Chem.  210,  149. 
MacLennan,  D.H.  and  Beevers,  H.  (1964).  Phytochemistry  3,  109. 
MacLeod,  J.  (1951).  J.  Gen.  Physiol.  34,  705. 

MacLeod,  R.A.  and  Snell,  E.E.  (1950).  Ann.  N.  Y.  Acad.  Sci.  52,  1249. 
MacLeod,  R.M.,  Farkas,  W.,  Fridovich,  I.,  and  Handler,  P.  (1961).  J.  Biol.  Chem. 

236,  1841. 
McMahon,  R.E.  and  Mills,  J.  (1961).  J.  Med.  Pharm.  Chem.  4,  211. 
MacNider,  W.  deB.  (1918  a).  J.  Pharmacol.  Exptl.  Therap.  11,  186. 
MacNider,  W.  deB.  (1918  b).  Proc.  Soc.  Exptl.  Biol.  Med.  16,  82. 
MacNider,  W.  deB.  (1941).  J.  Pharmacol.  Exptl.  Therap.  73,  186. 
McShan,  W.H.  and  Meyer,  R.K.  (1946).  Arch.  Biochem.  9,  165. 
McShan,  W.H.,  Kramer,  S.,  and  Schlegel,  V.  (1954).  Biol.  Bull.  106,  341. 
Madinaveitia,  J.  (1946).  Biochem.  J.  40,  373. 
Madsen,  N.B.  (1956).  J.  Biol.  Chem.  113,  1067. 
Madsen,  N.B.  (1960).  Can.  J.  Biochem.  Physiol.  38,  481. 
Madsen,  N.B.  and  Cori,  C.F.  (1955).  Biochim.  Biophys.  Acta  18,  156. 
Madsen,  N.B.  and  Cori,  C.F.  (1956).  J.  Biol.  Chem.  223,  1055. 
Madsen,  N.B.  and  Gurd,  F.R.N.  (1956).  J.  Biol.  Chem.  223,  1075. 
Maengwyn-Davies,  G.D.  and  Friedenwald,  J.S.  (1954).  Arch.  Biochem.  Biophys.   53,29. 


REFERENCES  1035 

Mager,  J.  and  Magasanik,  B.  (1960).  J.  Biol.  Chem.  235,  1474. 

Mahadevan,  S.,  Murthy,  S.K.,  and  Ganguly,  J.  (1962).  Biochem.  J.  85,  326. 

Mahler,  H.R.  (1954).  J.  Biol.  Chem.  206,  13. 

Mahler,  H.R.  (1955).  In  "Methods  in  Enzymology"  (S.P.  Colowick  and  N.O.  Kaplan, 

eds.).  Vol.  2,  p.  688.  Academic  Press,  New  York. 
Mahler,  H.R.  (1956).  Advan.  Enzyi7wl.  17,  233. 
Mahler,  H.R.,  Mackler,  B.,  Green,  D.E.,  and  Bock,  R.M.  (1954).  J.  Biol.  Chem.  210, 

465. 
Mahler,  H.R.,  Fairhurst,  A.S.,  and  Mackler,  B.  (1955a).  J.  Am.  Chem.  Soc.  77,  1514. 
Mahler,  H.R.,  Hiibscher,  G.,  and  Baum,  H.M.  (1955  b).  J.  Biol.  Chem.  216,  625. 
Mahler,  H.R.,  Baum,  H.M.,  and  Hiibscher,  G.  (1956).  Science  124,  705. 
Mahler,  H.R.,  Raw,  I.,  Molinari,  R.,  and  do  Amaral,  D.F.  (1958).  J.  Biol.  Chem.  233, 

230. 
Maio,  J.J.  and  Rickenberg,  H.V.  (1962).  Exptl.  Cell.  Res.  11.  31. 
Maitra,  P.K.  and  Estabrook,  R.W.  (1962).  Biochem.  Biophys.  Res.  Commun.  9,  1. 
Maitra,  P.K.  and  Roy,  S.C.  (1959).  J.  Biol.  Chem.  234,  2497. 
Maitra,  P.K.  and  Roy,  S.C.  (1961).  Biochem.  J.  79,  446. 
Maitre,  L.  and  Staehelin,  M.  (1963).  Experientia  19,  573. 
Maizels,  M.  (1951).  J.  Physiol.  [London)  112,  59. 

Maizels,  M.  and  Remington,  M.  (1958).  J.  Physiol.  {London)  143,  283. 
Maizels,  M.,  Remington,  M.,  and  Truscoe,  R.  (19.58).  J.  Physiol.  (London)  140,  80. 
Makino,  K.  and  Koike,  M.  (1954  a).  Nature  174,  1056. 
Makino,  K.  and  Koike,  M.  (1954  b).  Enzymologia  17,  157. 
Malachowski,  R.  and  Wanczura,  T.  (1933).  Bull.  Intern.  Acad.  Polon.  Sci.,  Classe  Sci. 

Math.  Nat.  p.  547. 
Maley,  F.  (1958).  Federation  Proc.  17,  267. 
Maley,  F.  and  Lardy,  H.A.  (1955).  J.  Biol.  Chem.  214,  765. 
Maley,  F.  and  Ochoa,  S.  (1958).  J.  Biol.  Chem.  233,  1538. 
Maley,  G.F.  and  Maley,  F.  (1959).  J.  Biol.  Chem.  234,  2975. 
Maley,  G.F.  and  Maley,  F.  (1964).  J.  Biol.  Chem.  239,  1168. 
Malkin,  A.  and  Denstedt,  O.F.  (1956).  Can.  J.  Biochem.  Physiol.  34,  141. 
Mallus,  E.  (1931).  Clin.  Pediat.  13,  265. 

Malmstrom,  B.G.  (1962).  Arch.  Biochem.  Biophys.  Suppl.  1,  p.  247. 
Malvin,  R.L.  (1956).  Federation  Proc.  15,  125. 
Mancilla,  R.  and  Agosin,  M.  (1960).  Exptl.  Parsitol.  10,  43. 
Mandel,  H.G.  (1955).  In  "Antimetabolites  and  Cancer"  (C.P.  Rhoads,  ed.),  p.  199, 

Am.  Assoc.  Advance.  Sci.,  Washington,  D.C. 
Mandelstam,  J.  (1961).  Biochem.  J.  79,  489. 

Mann,  F.D.  and  Mann,  F.C.  (1939).  Am.  J.  Digest.  Diseases  6,  322. 
Mann,  P.J.G.  and  Quastel,  J.H.  (1941).  Biochem.  J.  35,  502. 
Manning,  D.T.  and  Niemann.  C.  (1958).  J.  Am.  Chem.  Soc.  80,  1478. 
Manning,  G.B.  and  Campbell,  L.L.,  Jr.  (1957).  Can.  J.  Microbiol.  3,  1001. 
Manning,  G.B.  and  Campbell,  L.L.,  Jr.  (1961).  J.  Biol.  Chem.  236,  2952. 
Mano,  Y.  and  Tanaka,  R.  (1960).  J.  Biochem.  (Tokyo)  47,  401. 
Mansour,  T.E.  (1963).  J.  Biol.  Chem.  238,  2285. 
Mapson,  L.W.  (1946).  Biochem.  J.  40,  240. 
Mapson,  L.W.  and  Breslow,  E.  (1958).  Biochem.  J.  68,  395. 
Mapson,  L.W.  and  Isherwood,  F.A.  (1963).  Biochem.  J.  86,  173. 


1036  REFERENCES 

Marcus,  G.J.  and  Manery,  J.F.  (1963).  Cab.  J.  Biochem.  Physiol.  41,  2529. 

Marcus,  P.I.  and  Talalay,  P.  (1955).  Proc.  Roy.  Soc.  B144,  116. 

Mardashev,  S.R.  and  Semina,  L.A.  (1961).  Biochemistry  (USSR)  (English  Transl.)  26, 

27. 
Marinetti,  G.V.,  Erbland,  J.,  Albrecht,  M.,  and  Stotz,  E.  (1957).  Biochim.  Biophys. 

Acta  26,  130. 
Marini,  M.  and  Levey,  S.  (1955).  Proc.  Soc.  Ezptl.  Biol.  Med.  88,  611. 
Marks,  P.A.,  Szeinberg,  A.,  and  Banks,  J.  (1961).  J.  Biol.  Chem.  236,  10. 
Markwardt,  F.  (1953  a).  Naturwissenschaften  40,  341. 
Markwardt,  F.  (1953  b).  Z.  Physiol.  Chem.  295,  264. 
Marre,  E.  and  Servettaz,  0.  (1958).  Arch.  Biochem.  Biophys.  75,  309. 
Marre,  E.,  Forti,  G.,  and  Pece,  G.  (1956).  Atti  Accad.  Naz.  Lincei  Rend.,  Classe  Sci. 

Fis.,  Mat.  Nat.  [8]  20,  646. 
Marsh,  C.A.  and  Levvy,  G.A.  (1957).  Biochem.  J.  66,  4p. 
Marsh,  C.A.  and  Levvy,  G.A.  (1958).  Biochem.  J.  68,  610. 

Marshall,  M.,  Metzenberg,  R.L.,  and  Cohen,  P.P.  (1961).  J.  Biol.  Chem.  236,  2229. 
Marshall,  P.B.  (1948).  Brit.  J.  Pharmacol.  3,  1. 
Martell,  A.E.  and  Calvin,  M.  (1952).  "Chemistry  of  the  Metal  Chelating  Compounds." 

Prentice-Hall,  Englewood  Cliffs,  New  Jersey. 
Martin,  G.J.  (1951).  "Biological  Antagonism."  McGraw-Hill  (Blakiston),  New  York. 
Martin,  G.J.  and  Beiler,  J.M.  (1947).  Arch.  Biochem.  15,  201. 
Martin,  G.J.  and  Beiler,  J.M.  (1948).  J.  Am.  Pharm.  Assoc,  Sci.  Ed.  37,  32. 
Martin,  G.J.,  Tolman,  L.,  and  Brendel,  R.  (1947).  Arch.  Biochem.  15,  324. 
Martin,  G.J.,  Brendel,  R.,  Graff,  M.,  and  Beiler,  J.M.  (1950).  J.  Am.  Pharm.  Assoc, 

Sci.  Ed.  39,  338. 
Martin,  R.G.  (1963).  J.  Biol.  Chem.  238,  257. 
Martin,  S.M.  (1954).  Can.  J.  Microbiol.  1,  6. 
Martin,  S.M.  (1955).  Can.  J.  Microbiol.  1,  644. 
Martonosi,  A.  and  Gouvea,  M.A.  (1961).  J.  Biol.  Chem.  236,  1345. 
Martonosi,  A.  and  Meyer,  H.  (1964).  J.  Biol.  Chem.  239,  640. 
Maschmann,  E.  (1937  a).  Biochem.  Z.  294,  1. 
Maschmann,  E.  (1937  b).  Biochem.  Z.  295,  1. 

Maschmann,  E.  and  Helmert,  E.  (1933  a).  Z.  Physiol.  Chem.  219,  99. 
Maschmann,  E.  and  Helmert,  E.  (1933  b).  Z.  Physiol.  Chem.  220,  199. 
Mason,  H.L.  (1930).  J.  Biol.  Chem.  86,  623. 
Mason,  M.  (1959).  J.  Biol.  Chem.  234,  2770. 

Mason,  S.F.  (1957).  Ciba  Found.  Symp.,  Chem.  Biol.  Purines  p.  60. 
Massart,  L.,  Vandendriessche,  L.,  and  Dufait,  R.  (1940).  Enzymologia  8,  204. 
Massey,  V.  (1953  a).  Biochem.  J.  53,  67. 
Massey,  V.  (1953  b).  Biochem.  J.  55,  172. 
Massey,  V.  (1958).  Biochim.  Biophys.  Acta  30,  500. 

Massey,  V.  and  Rogers,  W.P.  (1950).  Australian  J.  Sci.  Res.  Ser.  B3,  251. 
Massey,  V.  and  Singer,  T.P.  (1957).  J.  Biol.  Chem.  228,  263. 
Massini,  P.  (1957).  Acta  Botan.  Neerl.  6,  434. 
Masuda,  T.  (1955).  Phnrm.  Bull.  (Tokyo)  3,  434. 

Masuoka,  D.T.,  Berman,  D.A.,  and  Saunders,  P.R.  (1952).  Am.  J.  Physiol.  170,  301. 
Mathews,  A. P.  (1904).  Am.  J.  Physiol.  10,  290. 
Mathews,  C.K.  and  Cohen,  S.S.  (1963).  J.  Biol.  Chem.  238,  367. 


REFERENCES  1037 

Matkovics,  B.,  Foldeak,  S.,  and  Konsanszky,  A.  (1961).  Enzymologia  23,  314. 

Matsuo,  Y.  and  Greenberg,  D.M.  (1959).  J.  Biol.  Chem.  234,  507. 

Matthew,  M.  and  Neuberger,  A.  (1963).  Biochem.  J.  87,  601. 

Matthews,  R.E.F.  (1958).  Pharmacol.  Rev.  10,  359. 

Matthews,  R.E.F.  and  Proctor,  C.H.  (1956).  J.  Gen.  Microbiol.  14,  366. 

Mauzerall,  D.  and  Granick,  S.  (1958).  J.  Biol.  Chem.  232,  1141. 

Maver,  M.E.  and  Thompson,  J.W.  (1946).  J.  Natl.  Cancer  Inst.  6,  355. 

Maw,  G.A.  (1959).  Biochem.  J.  72,  602. 

Maxwell,  E.S.  (1957).  J.  Biol.  Chem.  229,  139. 

Maxwell,  E.S.,  Kalckar,  H.M.,  and  Strominger,  J.L.  (1956).  Arch.  Biochem.  Biophys. 

65,  2. 
Mayer,  A.M.  (1959).  Enzymologia  20,  313. 

Mazelis,  M.  and  Crevehng,  R.K.  (1963).  J.  Biol.  Chem.  238,  3358. 
Mazehs,  M.  and  Stumpf,  P.K.  (1955).  Plant  Physiol.  30,  237. 
Mazia,  D.  (1959).  Harvey  Lectures  53,  130. 
Mazur,  A.  (1955).  In  "Methods  in  Enzymology"  (S.P.  Colowick  and  N.O.  Kaplan, 

eds.).  Vol.  1,  p.  651.  Academic  Press,  New  York. 
Medina,  A.  and  Nicholas,  D.J.D.  (1957  a).  Biochem.  J.  66,  573. 
Medina,  A.  and  Nicholas,  D.J.D.  (1957  b).  Biochim.  Biophys.  Acta  25,  138. 
Mehl,  J.W.  (1944).  Science  99,  518. 
Meier,  R.  (1926).  Biochem.  Z.  174,  384. 
Meier,  R.  (1933).  Z.  Ges.  Exptl.  Med.  87,  283. 
Meister,  A.  (1950).  J.  Biol.  Chem.  184,  117. 

Mello  Ayres,  G.C.  and  Lara,  F.J.S.  (1962).  Biochim.  Biophys.  Acta  62,  435. 
Mendel,  B.  (1937).  Am.  J.  Cancer  30,  549. 
Mendel,  B.  and  Strelitz,  F.  (1937).  Nature  140,  771. 
Mendelsohn,  M.L.  (1955).  Am.  J.  Physiol.  180,  599. 
Mendez,  R.  (1946).  Science  104,  5. 

Mendez,  R.  and  Peralta,  B.  (1947).  J.  Pharmacol.  Exptl.  Therap.  90,  128. 
Mendez,  R.  and  Pisanty,  J.  (1949).  Arch.  Inst.  Cardiol.  Mex.  19,  709. 
Mendicino,  J.  and  Vasarhely,  F.  (1963).  J.  Biol.  Chem.  238,  3528. 
Meneghetti,  E.  (1922)'.  Biochem.  Z.  131,  38. 

Menon,  G.K.K.  and  Stern,  J.R.  (1960).  J.  Biol.  Chem.  235,  3393. 
Menon,  G.K.K. ,  Friedman,  D.L.,  and  Stern,  J.R.  (1960).  Biochim.  Biophys.  Acta  AA, 

375. 
Merrett,  M.J.  and  Syrett,  P.J.  (1960).  Physiol.  Plantarum  13,  237. 
Mescon,  H.  and  Flesch,  P.  (1952).  J.  Invest.  Dermatol.  18,  261. 
Messing,  R.A.  (1961).  Enzymologia  23,  52. 
Metzenberg,  R.L.  (1963).  Arch.  Biochem.  Biophys.  100,  503. 
Metzger,  R.P.,  Wilcox,  S.S.,  and  Wick,  A.N.  (1964).  J.  Biol.  Chem.  239,  1769. 
Meyer,  D.K.  (1952).  Federation  Proc.  11,  107. 
Meyer,  H.  (1960).  Arch.  Mikrohiol.  37,  18. 

Meyer,  H.,  Preiss,  B.,  and  Bauer,  S.  (1960).  Biochem.  J.  76,  27. 
Meyer,  J.R.  (1948).  Science  108,  188. 
Meyer,  V.  and  Wachter,  W.  (1892).  Ber.  25,  2632. 
Meyerhof,  O.  and  Burk,  D.  (1928).  Z.  Physik.  Chem.  A139,  117. 
Michaelis,  L.  and  Rona,  P.  (1914).  Biochem.  Z.  60,  62. 
Michaelis,  L.  and  Pechstein,  H.  (1914).  Biochem.  Z.  60,  79. 


1038  REFERENCES 

Mickelson,  M.N.  and  Schuler,  M.N.  (1953).  J.  Bacteriol.  65,  297. 

Middlebrook,  M.  and  Szent-Gyorgyi,  A.  (1955).  Biochim.  Biophys.  Acta  18,  407. 

Mii,  S.  and  Ochoa,  S.  (1957).  Biochim.  Biophys.  Acta  26,  445. 

Mildvan,  A.S.  and  Greville,  G.D.  (1962).  Biochem.  J.  82,  22p. 

Millar,  D.B.S.  and  Schwert,  G.W.  (1963).  J.  Biol.  Chem.  238,  3249. 

Millbank,  J.W.  (1957).  J.  Exptl.  Botany  8,  96. 

Miller,  A.  and  Waelsch,  B.  (1957).  J.  Biol.  Chem.  228,  365. 

Miller,  J.J.  and  Halpern,  C.  (1956).  Can.  J.  Microbiol.  2,  519. 

Miller,  O.N.  and  Olson,  R.E.  (1954).  Arch.  Biochem.  Biophys.  50,  257. 

Miller,  T.B.  and  Farah,  A.E.  (1962  a).  J.  Pharmacol.  Exptl.  Therap.  135,  102. 

Miller,  T.B.  and  Farah,  A.E.  (1962  b).  J.  Pharmacol.  Exptl.  Therap.  136,  10. 

Miller,  V.L.,  Klavano,  P.A.,  and  Csonka,  E.  (1960).  Toxicol.  Appl.  Pharmacol.  2,  344. 

Miller,  V.L.,  Klavano,  P.A.,  Jerstad,  A.C.,  and  Csonka,  E.  (1961).  Toxicol.  Appl. 

Pharmacol.  3,  459. 
Millerd,  A.  (1951).  Proc.  Linnean  Soc.  N.  S.  Wales  76,  123. 
Millington,  P.F.  and  Finean,  J.B.  (1958).  J.  Ultrastruct.  Res.  2,  215. 
Millington,  P.F.  and  Finean,  J.B.  (1961).  J.  Ultrastruct.  Res.  5,  470. 
Mills,  G.T.,  Paul,  J.,  and  Smith,  E.E.B.  (1953).  Biochem.  J.  53,  232. 
Mills,  R.R.  and  Cochran,  D.G.  (1963).  Biochim.  Biophys.  Acta  73,  213. 
Milstein,  C.  (1961).  Biochem.  J.  79,  591. 
Minakami,  S.,  Cremona,  T.,  Ringler,  R.L.,  and  Singer,  T.P.  (1963).  J.  Biol.  Chem. 

238,   1529. 
Minatoya,  H.,  Luduena,  P.P.,  and  von  Euler,  L.  (1960).  Federation  Proc.  19,  65. 
Mirsky,  A.E.  (1941).  J.  Gen.  Physiol.  24,  709. 
Mirsky,  A.E.  and  Anson,  M.L.  (1935).  J.  Gen.  Physiol.  18,  307. 
Mirsky,  A.E.  and  Anson,  M.L.  (1936  a).  J.  Gen.  Physiol.  19,  439. 
Mirsky,  A.E.  and  Anson,  M.L.  (1936  b).  J.  Gen.  Physiol.  19,  451. 
Misaka,  E.  and  Nakanishi,  K.  (1963).  J.  Biochem.  {Tokyo)  53,  465. 
Misra,  U.K.  and  French,  D.  (1960).  Biochem.  J.  77,  Ip. 
Mitchell,  I.L.S.  and  Bayne,  S.  (1952).  Biochem.  J.  50,  xxvii. 
Mitchell,  P.  (1953).  J.  Gen.  Microbiol.  9,  273. 
Mitchell,  P.  (1954).  Symp.  Soc.  Exptl.  Biol.  8,  254. 

Mitchell,  R.A.,  Kang,  H.H.,  and  Henderson,  L.M.  (1963).  J.  Biol.  Chem.  238,  1151. 
Mittermayer,  C.  and  Breuer,  H.  (1963).  Biochim.  Biophys.  Acta  77,  191. 
Miura,  N.  (1934).  Japan.  J.  Med.  Sci.  IV.  Pharmacol.  8,  No.  1. 
Miyachi,  S.  (1960).  Plant  Cell  Physiol.  (Tokyo)  1,  117. 
Mize,  C.E.  and  Langdon,  R.G.  (1962).  J.  Biol.  Chem.  237,  1589. 
Mizushima,  S.  and  Izaki,  K.  (1958).  Proc.  Intern.  Symp.  Enzyme  Chem.,  Tokyo  Kyoto 

1957  p.  307.  Maruzen,  Tokyo. 
Modell,  W.  and  Krop,  S.  (1944).  Proc.  Soc.  Exptl.  Biol.  Med.  55,  80. 
Modignani,  R.L.  and  Foa,  P.P.  (1963).  Proc.  Soc.  Exptl.  Biol.  Med.  112,  106. 
Mokrasch,  L.C.  and  McGilvery,  R.W.  (1956).  J.  Biol.  Chem.  221,  909. 
Mohnari,  R.  and  Lara,  F.J.S.  (1960).  Biochem.  J.  75,  57. 

Molnar,  D.M.,  Burris,  R.H.,  and  Wilson,  P.W.  (1948).  J.  Am.  Chem.  Soc.  70,  1713. 
Momose,  G.  (1914).  Biochem.  Z.  61,  312. 
Momose,  G.  (1925).  J.  Biochem.  (Tokyo)  4,  441. 
Monroy,  A.  and  Runnstrom,  J.  (1952).  Exptl.  Cell.  Res.  3,  10. 
Montgomery,  CM.  and  Bamberger,  J.W.  (1955).  Science  122,  967. 


REFERENCES  1039 

Montgomery,  CM.  and  Webb,  J.L.  (1956  a).  J.  Biol.  Chem.  221,  347. 

Montgomery,  CM.  and  Webb,  J.L.  (1956  b).  Unpublished  work. 

Mookerjea,  S.  and  Sadhu,  D.P.  (1955).  Arch.  Biochem.  Biophys.  58,  232. 

Moore,  B.W.  (1959).  Arch.  Biochem.  Biophys.  80,  458. 

Moos,  C,  Alpert,  N.R.,  and  Myers,  T.C  (1960).  Arch.  Biochem.  Biophys.  88,  183. 

Mora,  P.T.  (1962).  J.  Biol.  Chem.  237,  3210. 

Mora,  P.T.  and  Young,  E.G.  (1959).  Arch.  Biochem.  Biophys.  82,  6. 

Morawetz,  H.  and  Sage,  H.  (1955).  Arch.  Biochem.  Biophys.  56,  103. 

Morgan,  E.J.  and  Friedmann,  E.  (1938  a).  Biochem.  J.  32,  733. 

Morgan,  E.J.  and  Friedmann,  E.  (1938  b).  Biochem.  J.  32,  862. 

Morgan,  E.J.  and  Friedmann,  E.  (1938  c).  Biochem.  J.  32,  2296. 

Morgan,  H.E.  and  Park,  CR.  (1958).  Federation  Proc.  17,  278. 

Morgan,  H.E.,  Henderson,  M.J.,  Regen,  D.M.,  and  Park,  CR.  (1961  a).  J.  Biol.  Chem. 

236,  253. 
Morgan,  H.E.,  Cadenas,  E.,  Regen,  D.M.,  and  Park,  CR.  (1961b).  J.  Biol.  Chem.  236, 

262. 
Morgan,  H.R.  (1954).  J.  Exptl.  Med.  99,  451. 
Morihara,  K.  (1963).  Biochim.  Biophys.  Acta  73,  113. 

Morino,  Y.,  Wada,  H.,  Morisue,  T.,  Sakamoto,  Y.,  and  Ichihara,  K.  (1960).  J.  Bio- 
chem. {Tokyo)  48,  18. 
Morisue,  T.,  Morino,  Y.,  Sakamoto,  Y.,  and  Ichihara,  K.  (1960).  J.  Biochem.  {Tokyo) 

48,  28. 
Moriyama,  H.  and  Ohashi,  S.  (1941).  Arch.  Ges.  Virusforsch.  2,  205. 
Morris,  M.P.,  Pagan,  C,  and  Warmke,  H.E.  (1954).  Science  119,  322. 
Morrison,  J.F.  (1950).  Australian  J.  Exptl.  Biol.  Med.  Sci.  28,  311. 
Morrison,  J.F.,  Griffiths,  D.E.,  and  Ennor,  A.H.  (1957).  Biochem.  J.  65,  143. 
Morton,  H.E.,  Kocholaty,  W.,  Junowicz-Kocholaty,  R.,  and  Kelner,  A.  (1945).  J. 

Bacterial.  50,  579. 
Morton,  R.K.  (1955).  Biochem.  J.  61,  232. 
Moses,  V.  (1955).  J.  Gen.  Microbiol.  13,  235. 
Moulder,  J.W.,  McCormack,  B.R.S.,  and  Itatani,  M.K.  (1953).  J.  Infect.  Diseases  93, 

140. 
Moulder,  J.W.,  Novosel,  D.L.,  and  Officer,  J.E.  (1963).  J.  Bacteriol.  85,  707. 
Mounter,  L.A.  (1954).  J.  Biol.  Chem.  290,  813. 
Mounter,  L.A.  and  Chanutin,  A.  (1953).  J.  Biol.  Chem.  204,  837. 
Mounter,  L.A.  and  Whittaker,  V.P.  (1953).  Biochem.  J.  53,  167. 
Moussatche,  H.  and  Prouvost-Danon,  A.  (1957).  Naturwissenschaften  44,  637. 
Moussatche,  H.  and  Prouvost-Danon,  A.  (1958).  Arch.  Biochem.  Biophys.  77,  108. 
Moyed,  H.S.  (1960).  J.  Biol.  Chem.  235,  1098. 
Moyed,  H.S.  and  O'Kane,  D.J.  (1956).  J.  Biol.  Chem.  218,  831. 

Moyer,  J.H.,  Handley,  C.A.,  and  Seibert,  R.A.  (1957).  Am.  N.  Y.  Acad.  Sci.  65,  511. 
Mozen,  M.M.  (1955).  Ph.  D.  Thesis,  University  of  Wisconsin,  Madison. 
Mozen,  M.M.,  Burris,  R.H.,  Lundbom,  S.,  and  Virtanen,  A.I.  (1955).  Acta  Chem. 

Scand.  9,  1232. 
Mudd,  J.B.  and  McManus,  T.T.  (1962).  J.  Biol.  Chem.  237,  2057. 
Mudd,  J.B.  and  McManus,  T.T.  (1964).  Plant  Physiol.  39,  115. 
Mudd,  J.B.  and  Stumpf,  P.K.  (1961).  J.  Biol.  Chem.  236,  2602. 
Mudge,  G.H.  (1951).  Am.  J.  Physiol.  167,  206. 


1040  REFERENCES 

Mudge,  G.H.  and  Weiner,  I.M.  (1958).  Ann.  N.  Y.  Acad.  Sci.  71,  344. 

Mueller,  J.F.  and  Vilter,  R.W.  (1950).  J.  Clin.  Invest.  29,  193. 

Mueller,  J.F.,  lacono,  J.,  Will.  J.J.,  and  Repasky,  W.  (1959).  J.  Lab.  Clin.  Med.  54, 

929. 
Muir,  C,  de  Kock,  L.L.,  de  Kock,  P.C.,  and  Inkson,  R.H.E.  (1959).  Experientia  15, 

354. 
Miiller,  A.F.  and  Leuthardt,  F.  (1950).  Helv.  Chim.  Acta  33,  268. 
Muller,  F.,  Schoeller,  W.,  and  Schrauth,  W.  (1911).  Biochem.  Z.  33,  381. 
Miiller,  G.,  Iwainsky,  H.,  and  Taufel,  K.  (1960).  Z.  Naturforsch.  15b,  180. 
Miiller,  E.  and  Miiller-Rodloff,  I.  (1935).  Ann.  Chem.  521,  81. 
Mullins,  L.J.  (1960).  J.  Gen.  Physiol.  43,  Suppl.  1,  105. 
Murachi,  T.  and  Neurath,  H.  (1960).  J.  Biol.  Chem.  235,  99. 
Murachi,  T.  and  Tashiro,  M.  (1958).  Biochim..  Biophys.  Acta  29,  645. 
Murayaraa,  M.  (1958).  J.  Biol.  Chem.  230,  163. 
Murray,  D.R.P.  (1929).  Biochem.  J.  23,  292. 
Murray,  D.R.P.  and  King,  C.G.  (1930).  Biochem.  J.  lA,  190. 
Muscholl,  E.  (1958).  Arch.  Exptl.  Pathol.  Pharmakol.  235,  23. 
Muscholl,  E.  and  Maitre,  L.  (1963).  Experientia  19,  658. 

Mushett,  C.W.,  Stebbins,  R.B.,  and  Barton,  M.N.  (1947).  Am.  J.  Med.  Sci.  213,  509. 
Mustakallio,  K.K.  and  Telkka,  A.  (1953).  Science  118,  320. 

Muus,  J.,  Brockett,  F.P.,  and  Connelly,  C.C.  (1956).  Arch.  Biochem.  Biophys.  65,  268. 
Myers,  D.K.  and  Slater,  E.G.  (1957  a).  Nature  179,  363. 
Myers,  D.K.  and  Slater,  E.G.  (1957  b).  Biochem.  J.  67,  572. 
Myrback,  K.  (1926).  Z.  Physiol.  Chem.  158,  160. 
Myrback,  K.  (1957  a).  Arch.  Biochem.  Biophys.  69,  138. 
Myrback,  K.  (1957  b).  Festschr.  Arthur  Stoll  p.  551. 
Myrback,  K.  and  Persson,  B.  (1953  a).  Arkiv  Kemi  5,  177. 
Myrback,  K.  and  Persson,  B.  (1953  b).  Arkiv  Kemi  5,  477. 

Naber,  E.G.,  Cravens,  W.W.,  Baumann,  C.A.,  and  Bird,  H.R.  (1954).  J.  Nutr.  54,  579. 

Nachlas,  M.M.,  Margulies,  S.I.,  and  Seligman,  A.M.  (1960).  J.  Biol.  Chem.  235,  499. 

Nachmansohn,  D.  and  Lederer,  E.  (1939).  Bull.  Soc.  Chim.  Biol.  21,  797. 

Nachmansohn,  D.  and  Machado,  A.L.  (1943).  J.  Neurophysiol.  6,  397. 

Nadai,  Y.  (1958).  J.  Biochem.  (Tokyo)  45,  1011. 

Nadkarni,  S.R.  and  Sohonie,  K.  (1963).  Enzymologia  25,  337. 

Naganna,  B.  and  Menon,  V.K.N.  (1948).  J.  Biol.  Chem..  174,  501. 

Naik,  K.G.  and  Patel,  R.P.  (1932).  J.  Indian  Chem.  Soc.  9,  533. 

Nair,  P.M.  and  Vining,  L.C.  (1964).  Arch.  Biochem.  Biophys.  106,  422. 

Nakada,  H.I.  (1964).  J.  Biol.  Chem.  239,  468. 

Nakada,  H.I.  and  Weinhouse,  S.  (1950).  J.  Biol.  Chem.  187,  663. 

Nakada,  H.I.  and  Wick,  A.N.  (1956).  J.  Biol.  Chem.  Ill,  671. 

Nakada,  H.I.,  Wolfe,  J.B.,  and  Wick,  A.N.  (1957).  J.  Biol.  Chem.  lib,  145. 

Nakamura,  H.  and  Greenberg,  D.M.  (1961).  Arch.  Biochem.  Biophys.  93,  153. 

Nakamura,  K.  (1960).  Biochim.  Biophys.  Acta  45,  554. 

Nakamura,  K.  and  Bernheim,  F.  (1961).  Biochim..  Biophys.  Acta  50,  147. 

Nakamura,  M.  (1954).  J.  Biochem.  [Tokyo)  41,  67. 

Nakamura,  M.  and  Baker,  E.E.  (1956).  Am.  J.  Hyg.  64,  12. 

Nakata,  H.M.  and  Halvorson,  H.O.  (1960).  J.  Bacteriol.  80,  801. 


REFERENCES  1041 

Nakayama,  T.  (1959).  J.  Biochem.  {Tokyo)  46,  1217. 

Nakayama,  T.  (1961a).  J.  Biochem.  (Tokyo)  49,  158. 

Nakayama,  T.  (1961b).  J.  Biochem.  (Tokyo)  49,  240. 

Nanninga,  L.B.  (1958).  Pacific  Slope  Biochem.  Conf.  (Univ.  Calif.,  Los  Angeles)  48. 

Naoi-Tada,  M.,  Sato-Asano,  K.,  and  Egami,  F.  (1959).  J.  Biochem.  (Tokyo)  46,  757. 

Narrod,  S.A.  and  Jakoby,  W.B.  (1964).  J.  Biol.  Chem.  239,  2189. 

Nath,  R.  and  Greenberg,  D.M.  (1962).  Biochemistry  1,  435. 

Nathans,  D.,  Tapley,  D.F.,  and  Ross,  J.E.  (1960).  Biochim.  Biophys.  Acta  41,  271. 

Navazio,  F.,  Siliprandi,  N.,  and  Rossi-Fanelli,  A.  (1956).  Biochim.  Biophys.  Acta  19, 

274. 
Neame,  K.D.  (1961).  Nature  192,  173. 
Neame,  K.D.  (1964).  J.  Neurochem.  11,  67. 
Neilands,  J.B.  (1952).  J.  Biol.  Chem.  199,  373. 
Neilands,  J.B.  (1954).  J.  Biol.  Chem.  208,  225. 

Neims,  A.H.  and  Hellerman,  L.  (1962).  J.  Biol.  Chem.  237,  PC976. 
Nelson,  C.A.  and  Hummel,  J.P.  (1961).  J.  Biol.  Chem.  236,  3173. 
Nelson,  E.K.  and  Hasselbring,  H.  (1931).  J.  Am.  Chem.  Soc.  53,  1040. 
Nelson,  L.  (1959).  Biol.  Bull.  117,  327. 
Nelson,  M.  (1964).  Comp.  Biochem.  Physiol.  12,  37. 
Nelson,  M.M.  (1955).  In  "Antimetabolites  and  Cancer"  (C.P.  Rhoads,  ed.),  p.  107. 

Am.  Assoc.  Advance.  Sci.,  Washington,  D.C. 
Neufeld,  E.F.,  Kaplan,  N.O.,  and  Colowick,  S.P.  (1955).  Biochim.  Biophys.  Acta  17, 

525. 
Neuhaus,  F.C.  and  Byrne,  W.L.  (1959).  J.  Biol.  Chem.  234,  113. 
Neuhaus,  F.C.  and  Byrne,  W.L.  (1960).  J.  Biol.  Chem.  235,  2019. 
Neuhaus,  F.C.  and  Lynch,  J.L.  (1964).  Biochemistry  3,  471. 
Neuhaus,  H.  (1910).  Arch.  Intern.  Pharmacodyn.  20,  393. 
Neurath,  H.  and  Gladner,  J.A.  (1951).  J.  Biol.  Chem.  188,  407. 
Neurath,  H.  and  Schwert,  G.W.  (1950).  Chem.  Rev.  46,  69. 
Newey,  H.  and  Smyth,  D.H.  (1964).  J.  Physiol.  (London)  170,  328. 
Newmark,  M.Z.  and  Wenger,  B.S.  (1960).  Arch.  Biochem.  Biophys.  89,  110. 
Ngai,  S.H.,  Ginsburg,  S.,  and  Katz,  R.L.  (1961).  Biochem.  Pharmacol.  7,  256. 
Nichol.  C.A.  (1954).  Cancer  Res.  14,  522. 

Nichol,  C.A.  and  Welch,  A.D.  (1950).  Proc.  Soc.  Exptl.  Biol.  Med.  74,  403. 
Nicholas,  D.J.D.,  Medina,  A.,  and  Jones,  O.T.G.  (1960).  Biochim.  Biophys.  Acta  37, 

468. 
Nickerson,  M.  and  Nomaguchi,  G.M.  (1950).  Am.  J.  Physiol.  163,  484. 
Nickerson,  W.J.  and  Chadwick,  J.B.  (1946).  Arch.  Biochem.  10,  81. 
Niedergang-Kamien,  E.  and  Leopold,  A.C.  (1957).  Physiol.  Plantarum  10,  29. 
Niederpruem,  D.J.  and  Hackett,  D.P.  (1961).  Plant  Physiol.  36,  79. 
Niemer,  H.  and  Kohler,  A.  (1957).  Z.  Physiol.  Chem.  308,  58. 

Nigam,  V.N.,  Davidson,  H.M.,  and  Fishman,  W.H.  (1959).  J.  Biol.  Chem.  234,  1550. 
Nihei,  T.,  Noda,  L.,  and  Morales,  M.F.  (1961).  J.  Biol.  Chem.  236,  3203. 
Nimmo-Smith,  R.H.  and  Standen,  O.D.,  (1963).  Exptl.  Parasitol.  13,  305. 
Ninomiya,  H.  and  Suzuoki,  Z.  (1952).  J.  Biochem.  (Tokyo)  39,  321. 
Nirenberg,  M.W.  and  Hogg,  J.F.  (1957).  Federation  Proc.  16,  227. 
Nu-enberg,  M.W.  and  Hogg,  J.F.  (1958).  Cancer  Res.  18,  518. 
Nishi,  A.  (1958).  J.  Biochem.  (Tokyo)  45,  991. 


1042  REFERENCES 

Nishi,  A.  (1960).  J.  Biochem.  (Tokyo)  47,  47. 

Nishimura,  J.S.  and  Greenberg,  D.M.  (1961).  J.  Biol.  Chem.  236,  2684. 

Nishimura,  S.  (1960).  Biochim.  Biophys.  Acta  45,  15. 

Nistratova,  S.N.  and  Turpaev,  T.M.  (1959).  Biochemistry  (USSR)  (English  Transl.) 

lA,  155. 
Noda,  H.  and  Maruyama,  K.  (1960).  Biochim.  Biophys.  Acta  41,  393. 
Noda,  L.  (1958).  J.  Biol.  Chem.  1^1,  237. 

Noda,  L.,  Nihei,  T.,  and  Morales,  M.F.  (1960).  J.  Biol.  Chem.  235,  2830. 
Noe,  F.E.  (1960).  Ind.  Med.  Surg.  29,  559. 

Noguchi,  H.  and  Akatsu,  S.  (1917).  J.  Pharmacol.  Exptl.  Therap.  9,  363. 
Nohara,  H.  and  Ogata,  K.  (1959).  Biochim.  Biophys.  Acta  31,  142. 
Noltmann,  E.  and  Bruns,  F.H.  (1959).  Biochem.  Z.  331,  436. 
Nordlie,  R.C.  and  Lardy,  H.A.  (1961).  Biochim.  Biophys.  Acta  53,  309. 
Nosoh,  Y.  (1961).  J.  Biochem.  (Tokyo)  50,  450. 
Nossal,  P.M.  (1952).  Biochem.  J.  50,  349. 

Novikoff,  A.B.,  Hecht,  L.,  Podber,  E.,  and  Ryan,  J.  (1952).  J.  Biol.  Chem.  194,  153. 
Novoa,  W.B.  and  Schwert,  G.W.  (1961).  J.  Biol.  Chem.  236,  2150. 
'^fovoa,  W.B.,  Winer,  A.D.,  Glaid,  A. J.,  and  Schwert,  G.W.  (1959).  J.  Biol.  Chem.  234, 

1143. 
Nukada,  T.,  Matsuoka,  M.,  and  Imaizumi,  R.  (1962).  Japan.  J.  Pharmacol.  12,  57. 
Nygaard,  A.P.  (1955).  Acta  Chem.  Scand.  9,  1048. 
Nygaard,  A.P.  (1956).  Acta  Chem.  Scand.  10,  397. 
Nygaard,  A.P.  (1961a).  Ann.  N.  Y.  Acad.  Sci.  94,  774. 
Nygaard,  A.P.  (1961b).  J.  Biol.  Chem.  236,  920. 
Nygaard,  A.P.  (1961c).  J.  Biol.  Chem.  236,  2779. 
Nyhan,  W.L.  and  Busch,  H.  (1957).  Cancer  Pes.  17,  227. 

Gates,  J.A.,  Gillespie,  L.,  Udenfriend,  S.,  and  Sjoerdsma,  A.  (1961).  Science  131,  1890. 

Ochoa,  S.  (1944).  J.  Biol.  Chem.  155,  87. 

Ochoa,  S.  and  Weisz-Tabori,  E.  (1948).  J.  Biol.  Chem.  174,  123. 

O'Connor,  R.J.  and  Halvorson,  H.O.  (1960).  Arch.  Biochem.  Biophys.  91,  290. 

O'Connor,  R.J.  and  Halvorson,  H.O.  (1961a).  Biochim.  Biophys.  Acta  48,  47. 

O'Connor,  R.J.  and  Halvorson,  H.O.  (1961b).  J.  Bacterial.  82,  706. 

Ogata,  K.,  Ogata,""  M.,  Mochizuki,  Y.,  and  Nishiyama,  T.  (1956).  J.  Biochem.  (Tokyo) 

43,  653. 
Ogilvie,  R.F.  (1932).  J.  Pathol.  Bacteriol.  35,  743. 
Oginsky,  E.L.  and  Gehrig,  R.F.  (1952).  J.  Biol.  Chem.  198,  799. 
Ohno,  M.  and  Okunuki,  K.  (1962).  J.  Biochem.  (Tokyo)  51,  313. 
Ohr,  E.A.  (1960).  Federation  Proc.  19,  248. 
Ohta,  J.  (1954).  J.  Biochem.  (Tokyo)  41,  489. 
Okazaki,  R.  and  Romberg,  A.  (1964).  J.  Biol.  Chem.  239,  275. 
Okunuki,  K.  (1939).  Acta  Phijtochim.  (Japan)  11,  27. 
Okunuki,  K.  and  Sekuzu,  I.  (1955).  J.  Biochem.  (Tokyo)  42,  397. 
Oliphant,  J.F.  (1942).  Physiol.  Zool.  15,  443. 
Olivard,  J.  and  Snell,  E.E.  (1955).  J.  Biol.  Chem.  213,  203. 
OHver,  I.T.  (1961).  Biochim.  Biophys.  Acta  52,  75. 
Olson,  E.J.  and  Park,  J.H.  (1964).  J.  Biol.  Chem.  239,  2316. 
Olson,  J. A.  (1959).  J.  Biol.  Chem.  234,  5. 


REFERENCES  1043 

Olson,  J. A.  and  Anfinsen,  C.B.  (1953).  J.  Biol.  Chem.  202,  841. 

Olson,  J.M.,  Duysens,  L.N.M.,  and  Kronenberg,  G.H.M.  (1959).  Biochim.  Biophys. 

Acta  36,  125. 
Onrust,  H.,  Van  der  Linden,  A.C.,  and  Jansen,  B.C.P.  (1952).  Enzymologia  15,  351. 
Onrust,  H.,  Van  der  Linden,  A.C.,  and  Jansen,  B.C.P.  (1954).  Enzymologia  16,  289. 
Ordin,  L.  and  Jacobson,  L.  (1955).  Plant  Physiol.  30,  21. 
Oren,  R.,  Farnham,  A.E.,  Saito,  K.,  Milofsky,  E.,  and  Karnovsky,  M.L.  (1963).  J. 

Cell  Biol.  17,  487. 
Orloff,  J.  and  Berliner,  R.W.  (1961).  Ann.  Rev.  Pharmacol.  1,  287. 
Ornstein,  N.  and  Gregg,  J.R.  (1952).  Biol.  Bull.  103,  407. 
Orten,  J.M.  and  Smith,  A.H.  (1937).  J.  Biol.  Chem.  117,  555. 
Osborn,  M.J.,  Freeman,  M.,  and  Huennekens,  F.M.  (1958).  Proc.  Soc.  Exptl.  Biol.  Med. 

97,  429. 
Ota,  S.,  Wachi,  T.,  Uchida,  M.,  Sakamoto,  Y.,  and  Ichihara,  K.  (1956).  J.  Biochem,. 

{Tokyo)  43,  611. 
Ota,  S.,  Fu,  T.-H.,  and  Hirohata,  R.  (1961).  J.  Biochem.  (Tokyo)  49,  532. 
Otsuka,  S.  (1961).  J.  Biochem.  {Tokyo)  50,  85. 
Ott,  J.L.  (1960).  J.  Bacteriol.  80,  355. 
Ott,  W.H.  (1946).  Proc.  Soc.  Exptl.  Biol.  Med.  61,  125. 
Ott,  W.H.  (1947).  Proc.  Soc.  Exptl.  Biol.  Med.  66,  215. 

Ottolenghi,  P.  and  Denstedt,  O.F.  (1958).  Can.  J.  Biochem.  Physiol.  36,  1075. 
Overell,  B.T.  (1952).  Australian  J.  Sci.  15,  28. 

Owens,  H.S.,  Goodban,  A.E.,'and  Stark,  J.B.  (1953).  Anal.  Chem.  25,  1507. 
Owens,  R.G.  (1953  a).  Contrib.  Boyce  Thompson  Inst.  17,  221. 
Owens,  R.G.  (1953  b).  Contrib.  Boyce  Thompson  Inst.  17,  273. 
Oxender,  D.L.  and  Christensen,  H.N.  (1963).  J.  Biol.  Chem.  238,  3686. 
Ozaki,  K.  and  Wetter,  L.R.  (1960).  Can.  J.  Biochem.  Physiol.  38,  125. 
Ozaki,  K.  and  Wetter,  L.R.  (1961).  Can.  J.  Biochem.  Physiol.  39,  843. 

Packer,  L.  (1958).  Exptl.  Cell.  Res.  15,  551. 

Packer,  L.,  Papa,  M.J.,  Rust,  J.H.,  Jr.,  and  Ajl,  S.J.  (1959).  Biochim.  Biophys.  Acta 
35,  354. 

Padykula,  H.A.  and  Herman,  E.  (1955).  J.  Histochem.  Cytochem.  3,  170. 

PaflF,  G.H.,  Sugiura,  H.T.,  Bocher,  C.A.,  and  Roth,  J.S.  (1952).  Anat.  Record  114,  499. 

Palmer,  G.  (1962).  Biochim.  Biophys.  Acta  64,  135. 

Palmer,  G.  and  Massey,  V.  (1962).  Biochim.  Biophys.  Acta  58,  349. 

Pantlitschko,  M.  and  Kaiser,  E.  (1951).  Biochem.  Z.  322,  137. 

Pantlitschko,  M.  and  Seelich,  F.  (1960).  Biochem.  Z.  333,  78. 
/)Papaconstantinou,  J.  and  Colowick,  S.P.  (1957).  Federation  Proc.  16,  230. 
^vPapaconstantinou,  J.  and  Colowick,  S.P.  (1961a).  J.  Biol.  Chem.  236,  278. 

Papaconstantinou,  J.  and  Colowick,  S.P.  (1961b).  J.  Biol.  Chem.  236,  285. 

Pardee,  A.B.  and  Potter,  V.R.  (1948).  J.  Biol.  Chem.  176,  1085. 

Pardee,  A.B.  and  Potter,  V.R.  (1949).  J.  Biol.  Chem.  178,  241. 

Pardee,  A.B.  and  Prestidge,  L.S.  (1958).  Biochim.  Biophys.  Acta  TJ,  330. 

Park,  R.B.,  Pon,  N.G.,  Louwrier,  K.P.,  and  Calvin,  M.  (1960).  Biochim.  Biophys.  Acta 
Al,  27. 

Parker,  A.J.  and  Kharasch,  N.  (1959).  Chem.  Rev.  59,  583. 

Parker,  C.A.  (1954).  Nature  173,  780. 


1044  REFERENCES 

Parker,  C.A.  and  Dilworth,  M.J.  (1963).  Biochim.  Biophys.  Acta  69,  152. 

Parker,  C.A.  and  Scutt,  P.B.  (1958).  Biochim.  Biophys.  Acta  29,  662. 

Parker,  C.A.  and  Scutt,  P.B.  (1960).  Biochim.  Biophys.  Acta  38,  230. 

Parks,  R.E.,  Jr.,  Ben-Gershom,  R.,  and  Lardy,  H.A.  (1957).  J.  Biol.  Chem.  227,  231. 

Parr,  C.W.  (1956).  Nature  178,  1401. 

Parr,  C.W.  (1957).  Biochem.  J.  65,  34p. 

Partridge,  A.D.  and  Rich,  A.E.  (1962).  Phytopathology  52,  1000. 

Passonneau,  J.V.  and  Lowtv,  O.H.  (1963).  Biochem.  Biophys.  Res.  Commun.  13,  372. 

Passow,  H.  and  Rothstein,  A.  (1960).  J.  Gen.  Physiol.  43,  621. 

Pasternak,  C.A.  and  Handschumacher,  R.E.  (1959).  J.  Biol.  Chem.  234,  2992. 

Pattabiraman,  T.N.  and  Bachhawat,  B.K.  (1961).  Biochim.  Biophys.  Acta  54,  273. 

Patterson,  W.B.  and  Stetten,  D.,  Jr.  (1949).  Science  109,  256. 

Pauling,  L.  (1946).  Chem.  Eng.  News  24,  1375. 

Pauling,  L.  (1948).  Nature  161,  707. 

Payes,  B.  and  Laties,  G.G.  (1963).  Biochem.  Biophys.  Res.  Commun.  10,  460. 

Pazur,  J.H.  and  Kleppe,  K.  (1964).  Biochemistry  3,  578. 

Pearson,  D.E.  and  Levine,  M.  (1952).  J.  Org.  Chem.  17,  1351. 

Pearson,  R.G.  and  Mills,  J.M.  (1950).  J.  Am.  Chem.  Sac.  72,  1692. 

Pease,  D.C.  (1941).  J.  Exptl.  Zool.  86,  381. 

Peck,  H.D.,  Jr.,  Smith,  O.H.,  and  Gest,  H.  (1957).  Biochim.  Biophys.  Acta  25,  142. 

Pedersen,  T.A.  (1963).  Physiol.  Plantarum  16,  167. 

Penefsky,  H.S.,  Pullman,  M.E.,  Datta,  A.,  and  Racker,  E.  (1960).  J.  Biol.  Chem.  235, 

3330. 
Penn,  N.  and  Mackler,  B.  (1958).  Biochim.  Biophys.  Acta  27,  539. 
Pennington,  R.J.  (1956).  Biocheyn.  J.  64,  6p. 
Pennington,  R.J.  (1957).  Biochem.  J.  65,  534. 
Pennington,  R.J.  (1961).  Biochem.  J.  80,  649. 
Pennington,  R.J.  and  Appleton,  J.M.  (1958).  Biochem.  J.  69,  119. 
Pentschew,  A.  and  Kassowitz,  H.  (1932).  Arch.  Exptl.  Pathol.  Pharmakol.  164,  667. 
Pereira,  A.S.R.  and  Linskens,  H.F.  (1963),  Acta  Botan.  Neerl.  12,  302. 
Perez,  J.E.,  Baralt-Perez,  J.,  and  Klein,  M.  (1949).  J.  Immunol.  62,  405. 
Perkins,  D.J.  (1952).  Biochem.  J.  51,  487. 
Perkins,  D.J.  (1953).  Biochem.  J.  55,  649. 
Perkins,  D.J.  (1958).  Biochem.  J.  69,  35p. 
Perkins,  D.J.  (1961).  Biochem.  J.  80,  668. 
Perry,  J.J.  and  Evans,  J.B.  (1960).  J.  BacteHol.  79,  113. 
Pershin,  G.N.  and  Shcherbakova,  L.I.  (1958).  Biochemistry  {USSR)  {English  Transl.) 

21,  155. 
Person,  P.  and  Fine,  A.  (1959).  Arch.  Biochem.  Biophys.  84,  123. 
Petering,  H.G.  and  Schmitt,  J. A.  (1950).  J.  Am.  Chem.  Soc.  72,  2995. 
Peters,  J.M.  and  Greenberg,  D.M.  (1959).  Biochim.  Biophys.  Acta  32,  273. 
Peters,  R.A.  and  Wakelin,  R.W.  (1946).  Biochem.  J.  40,  513. 
Peterson,  E.A.  and  Greenberg,  D.M.  (1952).  J.  Biol.  Chem.  194,  359. 
Petrucci,  D.  (1954  a).  Riv.  Biol.  {Perugia)  46,  241. 
Petrucci,  D.  (1954  b).  Boll.  Soc.  Ital.  Biol.  Sper.  30,  282. 

Pfleiderer,  G.,  Hohnholz-Merz,  E.,  and  Gerlach,  D.  (1962).  Biochem.  Z.  336,  371. 
Phares,  E.F.,  Mosbach,  E.H.,  Denison,  F.W.,  Jr.,  Carson,  S.F.,  Long,  M.V,,  and  Gwin, 

B.A.  (1952).  Anal.  Chem.  24,  660. 


REFERENCES  1045 

Philips,  F.S.,  Oilman,  A.,  Koelle,  E.S.,  and  Allen,  R.P.  (1947).  J.  Biol.  Chem.  167, 
209. 

Phillips,  A.H.  and  Langdon,  R.G.  (1962).  J.  Biol.  Chem.  237,  2652. 

Philpot,  J.St.L.  and  Small,  P.A.  (1938).  Biochem.  J.  32,  542. 

Pierpoint,  W.S.  (1959).  Biochem.  J.  71,  518. 

Pietschmann,  H.,  Lachnit,  V.,  and  Reinhardt,  F.  (1962).  Z.  Ges.  Exptl.  Med.  136, 
108. 

Pihl,  A.  and  Lange,  R.  (1962).  J.  Biol.  Chem.  237,  1356. 

Pihl,  A.,  Lange,  R.,  and  Evang,  A.  (1961).  Acta  Chem.  Scand.  15,  1271. 

Piloty,  0.  and  Schwerin,  E.G.  (1901a).  Ber.  34,  1863. 

Piloty,  O.  and  Schwerin,  E.G.  (1901b).  Ber.  34,  1870. 

Piloty,  O.  and  Schwerin,  E.G.  (1901c).  Ber.  34,  2354. 

Piloty,  O.  and  Vogel,  W.  (1903).  Ber.  36,  1283. 

Pine,  M.J.  (1960).  J.  Bacteriol.  79,  827. 

Pisanty,  J.  (1948).  Arch.  Inst.  Cardiol.  Mex.  18,  111. 

Pitts,  R.F.  (1958).  Am.  J.  Med.  24,  745. 

Pitts,  R.F.  (1959).  "The  Physiological  Easis  of  Diuretic  Therapy."  Thomas,  Spring- 
field, Illinois. 

Plant,  G.W.E.  (1957).  Arch.  Biochem.  Biophys.  69,  320. 

Plant,  G.W.E.  and  Lardy,  H.A.  (1949).  J.  Biol.  Chem.  180,  13. 

Pletscher,  A.,  Burkard,  W.P.,  and  Gey,  K.F.  (1964).  Biochem.  Pharmacol.  13,  385. 
^Plummer,  D.T.  and  Wilkinson,  J.H.  (1963).  Biochem.  J.  87,  423. 

Pogell,  E.M.  (1958).  J.  Biol.  Chem.  131,  761. 

Pogrund,  R.S.  and  Clark,  W.G.  (1956).  Federation  Proc.  15,  145. 

Pohl,  J.  (1896).  Arch.  Exptl.  Pathol.  Pharmakol.  37,  413. 

Pohle,  W.  and  Matthies,  H.  (1959).  Arch.  Exptl.  Pathol.  Phermakol.  236,  253. 

Polatnick,  J.  and  Bachrach,  H.L.  (1960).  Proc.  Soc.  Exptl.  Biol.  Med.  105,  601. 

Polls,  E.D.  and  Meyerhof,  O.  (1947).  J.  Biol.  Chem.  169,  389. 

Pollock,  M.R.  and  Knox,  R.  (1943).  Biochem.  J.  37,  476. 

Polonovski,  M.,  Schapira,  G.,  and  Gonnard,  P.  (1946).  Bull.  Soc.  Chim.  Biol.  28, 
735. 

Polyanovskii,  O.L.  (1962).  Biochemistry  (USSR)  (English  Transl.)  11,  623. 

Ponz,  F.  and  Llinas,  J.M.  (1963).  Nature  197,  696. 

Popjak,  G.  and  Tietz,  A.  (1955).  Biochem.  J.  60,  147. 

Popjak,  G.,  Gosselin,  L.,  Gore,  I.Y.,  and  Gould,  R.G.  (1958).  Biochem.  J.  69,  238. 

Porter,  C.C.  and  Hellerman,  L.  (1939).  J.  Am.  Chem.  Soc.  61,  754. 

Porter,  C.C.  and  Hellerman,  L.  (1944).  J.  Am.  Chem.  Soc.  66,  1652. 

Porter,  C.C,  Clark,  I.,  and  Silber,  R.H.  (1947).  J.  Biol.  Chem.  167,  573. 

Porter,  C.C,  Totaro,  J.A.,  and  Leiby,  CM.  (1961).  J.  Pharmacol.  Exptl.  Therap.  134, 
139. 

Portzehl,  H.  (1952).  Z.  Naturforsch.  7b,  1. 

Post,  R.L.,  Morgan,  H.E.,  and  Park,  C.R.  (1961).  J.  Biol.  Chem.  236,  269. 

Postgate,  J.R.  (1956).  J.  Gen.  Microbiol.  14,  545. 

Potter,  V.R.  (1940).  J.  Biol.  Chem.  134,  417. 

Potter,  V.R.  (1951).  Proc.  Soc.  Exptl.  Biol.  Med.  76,  41. 

Potter,  V.R.  and  DuEois,  K.P.  (1943).  J.  Gen.  Physiol.  26,  391. 

Potter,  V.R.  and  Elvehjem,  CA.  (1937).  J.  Biol.  Chem.  117,  341. 

Pott«r,  V.R.  and  Schneider,  W.C  (1942).  J.  Biol.  Chem.  142,  543. 


1046  REFERENCES 

Potter,  V.R.,  Busch,  H.,  and  Reif,  A.E.  (1952).  Proc.  2nd  Natl.  Cancer  Conf.  p.  1294. 

Price,  C.A.  (1953).  Arch.  Biochem.  Biophys.  47,  314. 

Pubols,  M.H.,  Zhanley,  J.C,  and  Axelrod,  B.  (1963).  Plant  Physiol.  38,  457. 

Pugh,  D.,  Leaback,  D.H.,  and  Walker,  P.G.  (1957).  Biochem.  J.  65,  464. 

PuUman,  M.E.,  Penefsky,  H.S.,  Datta,  A.,  and  Racker,  E.  (1960).  J.  Biol.  Chem.  235, 
3322. 

Pupilli,  G.  (1930).  Biochim.  Terap.  Sper.  17,  113. 

Purpura,  D.P.,  Berl,  S.,  Gonzalez-Monteagudo,  O.,  and  Wyatt,  A.  (1960).  In  "Inhi- 
bition in  the  Nervous  System  and  y-Aminobutyric  Acid"  (E.  Roberts,  ed.),  p.  331. 
Macmillan  (Pergamon),  New  York. 

Putnam,  F.W.  (1953).  In  "The  Proteins"  (H.  Neurath  and  K.  Bailey,  eds.),  Vol.  1, 
Pt.  B,  p.  807.  Academic  Press,  New  York. 

Putter,  J.  (1961).  Z.  Krehsforsch.  64,  101. 

Pyefinch,  K.A.  and  Mott,  J.C.  (1948).  J.  Exptl.  Biol.  25,  276. 

Quastel,  J.H.  and  Bickis,  I.J.  (1959).  Nature  183,  281. 
Quastel,  J.H.  and  Cantero,  A.  (1953).  Nature  171,  252. 
Quastel,  J.H.  and  Scholefield,  P.G.  (1955).  J.  Biol.  Chem.  214,  245. 
Quastel,  J.H.  and  Wheatley,  A.H.M.  (1931).  Biochem.  J.  25,  117. 
Quastel,  J.H.  and  Wheatley,  A.H.M.  (1935).  Biochem.  J.  29,  2773. 
Quastel,  J.H.  and  Whetham,  M.D.  (1925).  Biochem.  J.  19,  520. 
Quastel,  J.H.  and  Wooldridge,  W.R.  (1928).  Biochem.  J.  11,  689. 
Quesnel,  V.C.J.  (1956).  Exptl.  Cell.  Res.  10,  575. 
Quinn,  J.R.  and  Pearson,  A.M.  (1964).  Nature  201,  928. 

Raaflaub,  J.  (1953).  Helv.  Physiol.  Pharmacol.  Acta  11,  142. 

Rabin,  R.S.  and  Klein,  R.M.  (1957).  Arch.  Biochem.  Biophys.  70,  11. 

Rabinovitch,  M.  and  Barron,  E.S.G.  (1955).  Biochim.  Biophys.  Acta  18,  316. 

Rabinovitch,  M.  and  Dohi,  S.R.  (1957).  Arch.  Biochem.  Biophys.  70,  239. 

Rabinovitz,  M.  and  McGrath,  H.  (1959).  J.  Biol.  Chem.  234,  2091. 

Rabinovitz,  M.,  Olson,  M.E.,  and  Greenberg,  D.M.  (1954).  J.  Biol.  Chem.  210,  837. 

Rabinovitz,  M.,  Olson,  M.E.,  and  Greenberg,  D.M.  (1957).  Cancer  Res.  17,  885. 

Rabinowitz,  J.C.  and  Snell,  E.E.  (1953  a).  Arch.  Biochem.  Biophys.  43,  399. 

Rabinowitz,  J.C.  and  Snell,  E.E.  (1953  b).  Arch.  Biochem.  Biophys.  43,  408. 

Racker,  E.  and  Schroeder,  E.A.R.  (1958).  Arch.  Biochem.  Biophys.  74,  326. 

Racker,  E.,  Klybas,  V.,  and  Schramm,  M.  (1959).  J.  Biol.  Chem.  234,  2510. 

Rafter,  G.W.  (1957).  Arch.  Biochem.  Biophys.  67,  267. 

Rafter,  G.W.  (1958).  J.  Biol.  Chem.  230,  643. 

Rafter,  G.W.  and  Colowick,  S.P.  (1957).  Arch.  Biochem.  Biophys.  66,  190. 

Raghupathy,  E.,  Moudgal,  N.R.,  and  Sarma,  P.S.  (1958).  Enzymologia  19,  303. 

Ragland,  J.B.  and  Liverman,  J.L.  (1958).  Arch.  Biochem.  Biophys.  76,  496. 

Rahatekar,  H.I.  and  Rao,  M.R.R.  (1963).  Enzymologia  25,  292. 

Rahn,  O.  and  Barnes,  M.N.  (1933).  J.  Gen.  Physiol.  16,  579. 

Raiziss,  G.W.,  Severac,  M.,  and  Moetsch,  J.C.  (1926).  J.  Pharmacol.  Exptl.  Therap. 

26,  447. 
Rajagopalan,  K.V.  and  Handler,  P.  (1964).  J.  Biol.  Chem.  239,  2027. 
Rajagopalan,  K.V.,  Sundaram,  T.K.,  and  Sarma,  P.S.  (1960).  Biochem..  J.  74,  355. 
Rajagopalan,  K.V.,  Fridovich,  I.,  and  Handler,  P.  (1962).  J.  Biol.  Chem.  237,  922. 


REFERENCES  1047 

Rail,  J.E.,  Roche,  J.,  Michel,  R.,  Michel,  O.,  and  Varrone,  S.  (1962).  Biochem.  Biophys. 
Res.  Commun.  7,  111. 

Ram,  D.,  Kalner,  H.S.,  and  Bloch-Frankenthal,  L.  (1963).  Cancer  Res.  23,  600. 

Ramachandran,  S.  and  Gottlieb,  D.  (1963).  Biochim.  Biophys.  Acta  69,  74. 

Ramakrishnan,  C.V.  Steel,  R.,  and  Lentz,  C.P.  (1955).  Arch.  Biochem.  Biophys.  55, 
270. 

Ramasarma,  T.  and  Giri,  K.V.  (1956).  Arch.  Biochem.  Biophys.  62,  91. 

Ranzi,  S.  (1955).  "Symposium  on  the  Fine  Structure  of  Cells,"  p.  229.  Wiley  (Inter- 
science),  New  York. 

Rao,  D.R.  and  Oesper,  P.  (1961).  Biochem.  J.  81,  405. 

Rao,  M.R.R.  and  Altekar,  W.W.  (1961).  Biochem.  Biophys.  Res.  Commun.  4,  101. 

Rao,  M.R.R.  and  Gunsalus,  I.C.  (1955).  Federation  Proc.  14,  267. 

Rao,  N.A.,  Cama,  H.R.,  Kumar,  S.A.,  and  Vaidyanathan,  C.S.  (1960).  J.  Biol.  Chem. 
235,  3353. 

Rapkine,  L.  (1938).  Biochem.  J.  32,  1729. 

Rapoport,  S.  and  Luebering,  J.  (1951).  J.  Biol.  Chem.  189,  683. 

Rapoport,  S.,  Luebering,  J.,  and  Wagner,  R.H.  (1955).  Z.  Physiol.  Chem.  302,  105. 

Rapp,  G.W.  and  Sliwinski,  R.A.  (1956).  Arch.  Biochem.  Biophys.  60,  379. 

Raska,  S.B.  (1946).  J.  Am.  Med.  Assoc.  131,  1093. 

Rassweiler,  C.F.  and  Adams,  R.  (1924).  J.  Am.  Chem.  Sac.  46,  2758.  \ 

Rathbun,  R.C.  and  Shideman,  F.E.  (1951).  Federation  Proc.  10,  330. 

Ratner,  S.  (1955).  In  "Amino  Acid  Metabolism"  (E.D.  McElroy  and  H.B.  Glass,  eds.), 
p.  231.  Johns  Hopkins  Press,  Baltimore,  Maryland. 

Ratner,  S.  and  Pappas,  A.  (1949).  J.  Biol.  Chem.  179,  1199. 

Raw,  I.,  Nogueira,  O.C,  and  Filho,  J.B.M.  (1961).  Enzymologia  23,  123. 

Razzell,  W.E.  (1961).  J.  Biol.  Chem.  236,  3031. 

Razzell,  W.E.  and  Khorana,  H.G.  (1959).  J.  Biol.  Chem.  234,  2105. 

Read,  C.P.  (1952).  Exptl.  Parasitol  1,  353. 

Read,  C.P.  (1956).  Exptl.  Parasitol.  5,  325. 

Recknagel,  R.O.  and  Potter,  V.R.  (1951).  J.  Biol.  Chem.  191,  263. 

Redetzki,  H.M.  and  Alvarez-O'Bourke,  F.  (1962).  J.  Pharmacol.  Exptl.  Therap.  1 37, 173. 

Redfearn,  E.R.  (1960).  Ciha  Found.  Symp.,  Quinones  Electron  Transport  p.  346. 

Rees,  K.R.  (1953).  Biochem.  J.  55,  478. 

Rees,  K.R.  (1954).  Biochem.  J.  58,  196. 

Reichard,  P.  (1957).  Acta  Chem.  Scand.  11,  523. 

Reichard,  P.  and  HanshofF,  G.  (1956).  Acta  Chem..  Scand.  10,  548. 

Reichmann,  M.E.  and  Hatt,  D.L.  (1961).  Biochim.  Biophys.  Acta  49,  153. 

Reinbothe,  I.H.  (1957).  Pharmazie  12,  732. 

Reiner,  L.  and  Gellhorn,  A.  (1956).  J.  Pharmacol.  Exptl.  Therap.  117,  52. 

Reisberg,  R.B.  (1954).  Biochim.  Biophys.  Acta  14,  442. 

Reiss,  O.K.  and  Hellerman,  L.  (1958).  J.  Biol.  Chem.  231,  557. 

Remy,  C.N.  (1961).  J.  Biol.  Chem.  236,  2999. 

Rendina,  G.  and  Mills,  R.C.  (1957).  J.  Bacteriol.  74,  456. 

Rennels,  E.G.  and  Ruskin,  A.  (1954).  Proc.  Soc.  Exptl.  Biol.  Med.  85,  309. 

Repaske,  R.  (1954).  J.  Bacteriol.  68,  555. 

Repaske,  R.  and  Josten,  J.J.  (1958).  J.  Biol.  Chem.  233,  466. 

Repaske,  R.  and  Wilson,  P.W.  (1952).  J.  Am.  Chem.  Soc.  74,  3101. 

Resnick,  H.,  Carter,  J.R.,  and  Kalnitsky,  G.  (1959).  J.  Biol.  Chem.  234,  1711. 


1048  REFERENCES 

Revel,  H.R.  and  Racker,  E.  (1960).  Biochim.  Biophys.  Acta  43,  465. 

Reznikoff,  P.  (1926).  J.  Gen.  Physiol.  10,  9. 

Rhoads,  C.P.  (ed.)  (1955).  "Antimetabolites  and  Cancer"  Am.  Assoc.  Advance.  Sci., 

Washington,  D.C. 
Rhoads,  W.A.  and  Wallace,  A.  (1957).  Plant  Physiol.  32,  xxii. 
Rice,  L.I.  and  Berman,  D.A.  (1961).  Am.  J.  Physiol.  200,  727. 
Richards,  O.C.  and  Rutter,  W.J.  (1961).  J.  Biol.  Chem.  236,  3185. 
Richter,  D.  (1934).  Biochem.  J.  28,  901. 
Rieser,  P.  (1961).  Exptl.  Cell.  Res.  24,  165. 
Rigbi,  M.  and  Sela,  M.  (1964).  Biochemistry  3,  629. 
Riggs,  A.  (1959).  In  "Sulfur  in  Proteins"  (R.  Benesch  et  al.,  eds.),  p.  173.  Academic 

Press,  New  York. 
Riggs,  A.F.  (1952).  J.  Gen.  Physiol.  36,  1. 

Riggs,  A.F.  and  Wolbach,  R.A.  (1956).  J.  Gen.  Physiol.  39,  585. 
Riggs,  T.R.  and  Walker,  L.M.  (1963).  J.  Biol.  Chem.  238,  2663. 
Riggs,  T.R.,  Walker,  L.M.,  and  Christensen,  H.N.  (1958).  J.  Biol.  Chem.  233,  1479. 
Riklis,  E.  and  Quastel,  J.H.  (1958).  Can.  J.  Biochem.  Physiol.  36,  363. 
Rimington,  C.  and  Tooth,  B.E.  (1961).  J.  Biochem.  (Tokyo)  49,  456. 
Rindi,  G.  and  Ferrari,  G.  (1959).  Nature  183,  608. 
Rindi,  G.  and  Perri,  V.  (1961).  Biochem.  J.  80,  214. 
Rindi,  G.,  Perri,  V.,  and  Ventura,  U.  (1959).  Nature  183,  1126. 
Rindi,  G.,  Perri,  V.,  and  De  Caro,  L.  (1961).  Experientia  17,  546. 
Robbins,  W.J.  (1941).  Proc.  Natl.  Acad.  Sci.  U.  S.  27,  419. 
Robert,  B.,  Robert,  M.,  and  Robert,  L.  (1952).  Bull.  Soc.  Chim..  Biol.  34,  183. 
Roberts,  E.  (1953).  J.  Biol.  Chem.  202,  359. 

Roberts,  E.  and  Simonsen,  D.G.  (1963).  Biochem.  Pharmacol  12,  113. 
Roberts,  E.,  Younger,  F.,  and  Frankel,  S.  (1951).  J.  Biol.  Chem.  191,  277. 
Robertson,  M.E.  (1943).  Nature  151,  365. 

Robertson,  W.  van  B.,  Ropes,  M.W.,  and  Bauer,  W.  (1940).  J.  Biol.  Chem.  133,  261. 
Robinson,  B.  and  Shepherd,  D.M.  (1961).  Biochim.  Biophys.  Acta  53,  431. 
Robinson,  J.C.,  Keay,  L.,  Molinari,  R.,  and  Sizer,  I.W.  (1962).  J.  Biol.  Chem.  237, 

2001. 
Robinson,  J.D.,  Brady,  R.O.,  and  Bradley,  R.M.  (1963).  J.  Lipid  Res.  4,  144. 
Robinson,  J.R.  (1956).  J.  Physiol.  {London)  134,  216. 
Robinson,  W.G.  and  Coon,  M.J.  (1957).  J.  Biol.  Chem.  225,  511. 
Robson,  J.  and  Fenn,  P.  (1961).  Nature  189,  501. 
Rocca,  E.  and  Ghiretti,  F.  (1958).  Arch.  Biochem.  Biophys.  77,  336. 
Roche,  J.,  Glahn,  P.-E.,  Manchol,  P.,  and  Thoai,  N.  (1959).  Biochim.  Biophys.  Acta 

35,  111. 
Rogers,  E.F.  (1962).  Ann.  N.  Y.  Acad.  Sci.  98,  412. 

Rogers,  H.J.  and  Spensley,  P.C.  (1954).  Biochim.  Biophys.  Acta  13,  293. 
Rogers,  K.S.,  Thompson,  T.E.,  and  Hellerman,  L.  (1962).  Biochim.  Biophys.  Acta  64, 

202. 
Rogers,  K.S.,  Geiger,  P.J.,  Thompson,  T.E.,  and  Hellerman,  L.  (1963).  J.  Biol.  Chem. 

238,  481. 
Rogers,  L.L.  and  Shive,  W.  (1947).  J.  Biol.  Chem.  169,  57. 
Rohnson,  G.N.  (1954).  J.  Gen.  Microbiol.  11,  412. 
Romberger,  J.A.  and  Norton,  G.  (1961).  Plant  Physiol.  36,  20. 


REFERENCES  1049 

Rosa,  N.  and  Zalik,  S.  (1963).  Can.  J.  Biochem.  Physiol.  41,  533. 

Eoscoe,  H.E.  and  Nelson,  W.L.  (1964).  J.  Biol.  Ckem.  239,  8. 

Rose,  I.A.,  Grunberg-Manago,  M.,  Korey,  S.R.,  and  Ochoa,  S.  (1954).  J.  Biol.  Chem. 

211,  737. 
Rose,  W.C.  (1924).  J.  Pharmacol.  Exptl.  Therap.  24,  147. 
RoseU-Perez,  M.  and  Lamer,  J.  (1964).  Biochemistry  3,  773. 
Rosen,  F.,  Milholland,  R.J.,  and  Nichol,  C.A.  (1960).  In  "Inhibition  in  the  Nervous 

System  and  y-Aminobutyric  Acid"  (E.  Roberts,  ed.),  p.  338.  Macmillan  (Pergamon), 

New  York. 
Rosen,  S.W.  and  Klotz,  I.M.  (1957).  Arch.  Biochem.  Biophys.  67,  161. 
Rosenberg,  A.J.  (1948).  Compt.  Bend.  Soc.  Biol.  142,  443. 

Rosenberg,  L.E.,  Downing,  S.J.,  and  Segal,  S.  (1962).  J.  Biol.  Chem.  237,  2265. 
Rosenberg,  T.  and  Wilbrandt,  W.  (1962).  Exptl.  Cell  Res.  T7 ,  100. 
Rosenthal,  A.  and  Geyer,  R.P.  (1959).  Federation  Proc.  18,  544. 
Ross,  W.F.  and  Green,  L.S.  (1941).  J.  Biol.  Chem.  137,  105. 
Roth,  J.S.  (1953  a).  Nature  171,  127. 
Roth,  J.S.  (1953  b).  Arch.  Biochem.  Biophys.  44,  265. 
Roth,  J.S.  (1956).  Biochim.  Biophys.  Acta  21,  34. 
Roth,  J.S.  (1958).  J.  Biol.  Chem.  231,  1097. 

Rothschild,  A.M.  and  Schayer,  R.W.  (1959).  Biochim.  Biophys.  Acta  34,  392. 
Rothschild,  H.A.  and  Barron,  E.S.G.  (1954).  J.  Biol.  Chem.  209,  511. 
Rothschild,  H.A.,  Cori,  O.,  and  Barron,  E.S.G.  (1954).  J.  Biol.  Chem.  208,  41. 
Rothstein,  A.  (1959).  Federation  Proc.  18,  1026. 
Rothstein,  A.  (1964).  Personal  communication. 

Rothstein,  A.  and  Bruce,  M.  (1958).  J.  Cellular  Comp.  Physiol.  51,  439. 
Rothstein,  A.  and  Hayes,  A.D.  (1960).  J.  Pharmacol.  Exptl.  Therap.  130,  166. 
Roussos,  G.G.  and  Nason,  A.  (1960).  J.  Biol.  Chem.  235,  2997. 
Rowe,  A.W.  and  Weill,  C.E.  (1962).  Biochim.  Biophys.  Acta  65,  245. 
Rowland,  R.L.  (1952).  J.  Am.  Chem.  Soc.  74,  5482. 
Roy,  A.B.  (1953).  Biochem.  J.  55,  653. 
Roy,  A.B.  (1954a).  Biochim.  Biophys.  Acta  15,  300. 
Roy,  A.B.  (1954  b).  Biochem.  J.  57,  465. 
Roy,  A.B.  (1956).  Biochem.  J.  64,  651. 
Rubin,  B.A.  and  Ladygina,  M.E.  (1960).  Biochemistry  (USSR)  (English  Transl.)  25, 

617. 
Rubin,  R.J.,  Reynard,  A.,  and  Handschumacher,  R.E.  (1964).  Cancer  Res.  24,  1002. 
Rubino,  R.  and  Varela  (1923).  Klin.  Wochschr.  2,  484. 
Rubinstein,  D.,  Ottolenghi,  P.,  and  Denstedt,  O.F.  (1956).  Can.  J.  Biochem.  Physiol. 

34,  222. 
RufFo,  A.  and  Adinolfi,  A.  (1963).  Biochem.  J.  89,  50p. 
Ruffo,  A.  and  D'Abramo,  F.  (1952).  Boll.  Soc.  Ital.  Biol.  Sper.  28,  850. 
RufFo,  A.,  Testa,  E.,  Adinolfi,  A.,  and  Pelizza,  G.  (1962  a).  Biochem.  J.  85,  588. 
Ruffo,  A.,  Adinolfi,  A.,  Budillon,  G.,  and  Capobianco,  G.  (1962  b).  Biochem.  J.  85,  593. 
Rulon,  O.  (1948).  Physiol.  Zool.  21,  100. 

Rumpf,  P.,  d'Incan,  E.,  and  Schaal,  R.  (1955).  Bull.  Soc.  Chim.  France  p.  122. 
Runnstrom,  J.  and  Kriszat,  G.  (1952).  Exptl.  Cell.  Res.  3,  497. 
Ruska,  H.  (1947).  Arch.  Exptl.  Pathol.  Pharmakol.  204,  576. 
Ruskin,  A.  and  Jolmson,  J.E.  (1949).  Proc.  Soc.  Exptl.  Biol.  Med.  72,  572. 


1050  REFERENCES 

Ruskin,  A.  and  Ruskin,  B.  (1953).  J.  Pharmacol.  Exptl.  Therap.  107,  325. 

Russell,  P.  (1955).  Nature  176,  1123. 

Rutter,  W.J.  and  Lardy,  H.A.  (1958).  J.  Biol.  Chem.  233,  374. 

Ryan,  C.A.  and  King,  T.E.  (1962  a).  Biochim.  Biophys.  Acta  62,  269. 

Ryan,  C.A.  and  King,  T.E.  (1962b).  Arch.  Biochem.  Biophys.  85,  450. 

Ryan,  K.J.  and  Engel,  L.L.  (1957).  J.  Biol.  Chem.  225,  103. 

Ryley,  J.F.  (1955  a).  Biochem.  J.  59,  353. 

Ryley,  J.F.  (1955  b).  Biochem.  J.  59,  361. 

Ryzhkov,  V.L.  and  Marchenko,  N.K.  (1954).  DoH.  Akad.  Nauk  SSSR  98,  1033. 

Ryzhkov,  V.L.  and  Marchenko,  N.K.  (1955).  Bev.  Appl.  Mycol.  34,  617. 

Sabato,  G.  di  and  Fonnesu,  A.  (1959).  Biochim.  Biophys.  Acta  35,  358. 

Sachs,  G.  (1921).  Ber.  54,  1849. 

Sachs,  H.W.  and  Dulskas,  A.  (1956).  Arch.  Pathol.  Anat.  Physiol.  329,  466. 

Sacktor,  B.  (1953).  J.  Gen.  Physiol.  36,  371. 

Sacktor,  B.  and  Cochran,  D.G.  (1957).  J.  Biol.  Cher^.  226,  241. 

Sacktor,  B.,  Thomas,  G.M.,  Moser,  J.C,  and  Bloch,  D.I.  (1953).  Biol.  Bull.  105,  166. 

Sacktor,  B.,  Cummins,  J.I.,  and  Packer,  L.  (1959).  Federation  Proc.  18,  314. 

SafFran,  M.  and  Prado,  J.L.  (1949).  J.  Biol.  Chem.  180,  1301. 

Sahasrabudhe,  M.B.,  Nerurkar,  M.K.,  Narurkar,  M.V.,  Tilak,  B.D.,  and  Bhavsar, 

M.D.  (1960).  Brit.  J.  Cancer  14,  457. 
Sakai,  H.  (1962).  J.  Gen.  Physiol.  45,  411. 
Sakami,  W.  (1955).  In  "Amino  Acid  Metabolism"  (W.D.  McElroy  and  H.B.  Glass, 

eds.),  p.  658.  Johns  Hopkins  Press,  Baltimore,  Maryland. 
Sakata,  K.,  Hayano,  S.,  and  Sloviter,  H.A.  (1963).  Am.  J.  Physiol.  204,  1127. 
Salant,  W.  and  Brodman,  K.  (1929  a).  J.  Pharmacol.  Exptl.  Therap.  36,  195. 
Salant,  W.  and  Brodman,  K.  (1929  b).  J.  Pharmacol.  Exptl.  Therap.  37,  55. 
Salant,  W.  and  Brodman,  K.  (1929  c).  J.  Pharmacol.  Exptl.  Therap.  37,  121. 
Salant,  W.  and  Kleitman,  N.  (1922).  J.  Pharmacol.  Exptl.  Therap.  19,  315. 
Salant,  W.  and  Nadler,  J.E.  (1927).  J.  Lab.  Clin.  Med.  13,  117. 
Salant,  W.  and  Nagler,  H.  (1930).  Proc.  Soc.  Exptl.  Biol.  Med.  27,  859. 
Salant,  W.  and  Nagler,  H.  (1931).  J.  Pharmacol.  Exptl.  Therap.  41,  407. 
Salle,  A.J.  (1943).  J.  Pharmacol.  Exptl.  Therap.  79,  271. 
Salle,  A.J.  and  Ginoza,  Y.W.  (1943).  Proc.  Soc.  Exptl.  Biol.  Med.  54,  85. 
Salle,  A.J.  and  Lazarus,  A.S.  (1936).  Proc.  Soc.  Exptl.  Biol.  Med.  34,  371. 
Salles,  J.B.V.  and  Ochoa,  S.  (1950).  J.  Biol.  Chem.  187,  849. 
Saltman,  P.  (1953).  J.  Biol.  Chem.  200,  145. 

Saltman,  P.,  Fiskin,  R.D.,  and  Bellinger,  S.B.  (1955).  Federation  Proc.  14,  274. 
Saltman,  P.,  Alex,  T.,  and  McCornack,  B.  (1959).  Arch.  Biochem.  Biophys.  83,  538. 
Samborski,  D.J.  and  Forsyth,  F.R.  (1960).  Can.  J.  Botany  38,  467. 
Sanabria,  A.  (1963).  Brit.  J.  Pharmacol.  20,  352. 

Sanadi,  D.R.,  Littlefield,  J.W.,  and  Bock,  R.M.  (1952).  J.  Biol.  Chem.  197,  851. 
Sanborn,  R.C.  and  Williams,  CM.  (1950).  J.  Gen.  Physiol.  33,  579. 
Sankar,  D.V.S.  (1958).  Proc.  Soc.  Exptl.  Biol.  Med.  98,  198. 
Sanner,  T.  and  Pihl,  A.  (1962).  Biochim.  Biophys.  Acta  62,  171. 
Sanner,  T.  and  Pihl,  A.  (1963).  J.  Biol.  Chem.  238,  165. 
San  Pietro,  A.  and  Lang,  H.M.  (1958).  J.  Biol.  Chem.  231,  211. 
Sanwal,  B.D.  (1961).  Arch.  Biochem.  Biophys.  93,  377. 


KEFERENCES  1051 

Sanwal,  B.D.  and  Zink,  M.W.  (1961).  Arch.  Biochem.  Biophys.  94,  430. 

Saroff,  H.A.  and  Mark,  H.J.  (1953).  J.  Am.  Chem.  Soc.  75,  1420. 

Sasaki,  A.  (1940).  J.  Biochem.  (Tokyo)  32,  211. 

Sasaki,  S.  (1962).  J.  Biochem.  (Tokyo)  51,  335. 

Sato,  R.,  Takemori,  S.,  and  Ebata,  M.  (1956).  J.  Biochem.  (Tokyo)  43,  623. 

Sauer,  G.  and  Rapoport,  S.  (1959).  Biochem.  Z.  331,  534. 

Sauermann,  G.  (1964).  Arch.  Biochem.  Biophys.  104,  208. 

Sawai,  T.  (1960).  J.  Biochem.  (Tokyo)  48,  382. 

Sayre,  F.W.  and  Roberts,  E.  (1958).  J.  Biol.  Chem..  233,  1128. 

Saz,  A.K.,  Lowery,  D.L.,  and  Jackson,  L.J.  (1961).  J.  Bacterial.  82,  298. 

Saz,  H.J.  and  Krampitz,  L.O.  (1955).  J.  Bacteriol.  69,  288. 

Scala,  R.A.  and  Lambooy,  J.P.  (1958).  Arch.  Biochem.  Biophys.  78,  10. 

Scarano,  E.,  Bonaduce,  L.,  and  de  Petrocellis,  B.  (1960).  J.  Biol.  Chem..  235,  3556. 

Scarano,  E.,  Bonaduce,  L.,  and  de  Petrocellis,  B.  (1962).  J.  Biol.  Chem.  237,  3742. 

Schachter,  D.  and  Rosen,  S.M.  (1959).  Am.  J.  Physiol.  196,  357. 

Schachter,  D.  and  Taggart,  J.V.  (1954).  J.  Biol.  Chem.  211,  271. 

Schachter,  D.,  Dowdle,  E.B.,  and  Schenker,  H.  (1960).  Am.  J.  Physiol.  198,  275. 

Schachter,  H.,  Halliday,  K.A.,  and  Dixon,  G.H.  (1963).  J.  Biol.  Chem.  238,  PC3134. 

Schanker,  L.S.  and  Tocco,  D.J.  (1962).  Biochim.  Biophys.  Acta  56,  469. 

Schapira,  G.  (1946).  Compt.  Rend.  Soc.  Biol.  140,  173. 

Scharff,  T.G.  (1961).  Arch.  Biochem.  Biophys.  95,  329. 

Schatz,  V.B.  (1960).  In  "Medicinal  Chemistry"  (A.  Burger,  ed.),  2nd  ed.,  p.  72.  Wiley 

(Interscience),  New  York. 
Schauer,  R.  and  Hillmann,  G.  (1961).  Z.  Physiol.  Chem.  325,  9. 
Schayer,  R.W.,  Smiley,  R.L.,  and  Kennedy,  J.  (1954).  J.  Biol.  Chem.  206,  461. 
Schellenberg,  K.A.  and  Hellerman,  L.  (1958).  J.  Biol.  Chem.  231,  547. 
Schimmel,  N.H.,   Matteucci,  W.V.,  and   Roger,    W.P.  (1956).  Antibiot.    Med.  Clin. 

Therapy  3,  136. 
Schlegel,  D.E.  (1957).  Virology  4,  135. 

Schmidt,  G.  and  Thannhauser,  S.J.  (1943).  J.  Biol.  Chem.  149,  369. 
Schneider,  G.  and  Schmidt,  S.  (1959).  Z.  Physiol.  Chem.  315,  20. 
Schneiderman,  H.A.,  Ketchel,  M.,  and  Williams,  CM.  (1953).  Biol.  Bull.  105,  188. 
Scholefield,  P.G.  (1955).  Biochem.  J.  59,  177. 
Scholefield,  P.G.  (1961).  Can.  J.  Biochem.  Physiol.  39,  1717. 
Schom,  R.,  Menke,  K.-H.,  and  Negelein,  E.  (1961).  Z.  Physiol.  Chem.  323,  155. 
Schonbaum,  E.,  Birmingham,  M.K.,  and  SafFran,  M.  (1956).  Can.  J.  Biochem.  Physiol. 

34,  527. 
Schopfer,  W.H.  and  Keller,  V.  (1951).  Bull.  Soc.  Chim.  Biol.  33,  1253. 
Schorcher,  C.  and  Loblich,  H.J.  (1960).  Arch.  Pathol.  Anat.  Physiol.  333,  587. 
Schrecker,  A.W.  and  Huennekens,  F.M.  (1964).  Biochem.  Pharmacol.  13,  731. 
Schrodt,  G.R.,  Dizon,  F.Y.,  Peskoe,  L.Y.,  Howell,  R.S.,  Robie,  C.H.,  and  Stevenson, 

T.D.  (1960).  Am.  Practitioner  Dig.  Treat.  11,  750. 
Schroeder,  E.F.,  Woodward,  G.E.,  and  Piatt,  M.E.  (1933a).  J.  Biol.  Chem.  100,  525. 
Schroeder,  E.F.,  Woodward,  G.E.,  and  Piatt,  M.E.  (1933  b).  J.  Biol.  Chem.  101,  133. 
Schueler,  F.W.  (1960).  "Chemobiodynamics  and  Drug  Design."  McGraw-Hill,  New 

York. 
Schiller,  H.  (1932).  Biochem.  Z.  255,  474. 
Schuler,  W.  (1952).  Experientia  8,  230. 


1052  REFERENCES 

Schultz,  S.G.  and  Zaiusky,  R.  (1964).  J.  Gen.  Physiol.  47,  1043. 

Schulz,  H.  (1888).  Arch.  Ges.  Physiol.  42,  517. 

Schumann,  E.L.,  Paquette,  L.A.,  Heinzelman,  R.V.,  Wallach,  D.P.,  Da  Vanzo,  J.P., 

and  Greig,  M.E.  (1962).  J.  Med.  Pharm.  Chem.  5,  464. 
Schiitte,  H.R.  and  Niirnberger,  H.  (1959).  Z.  Physiol.  Chem.  315,  13. 
Schweet,  R.S.  and  Allen,  E.H.  (1958).  J.  Biol.  Chem.  233,  1104. 
Schweisfurth,  R.  and  Schwartz,  W.  (1959).  Acta  Biol.  Med.  Ger.  2,  54. 
Scott,  R.L.  and  Gamble,  J.L.,  Jr.  (1961).  J.  Biol.  Chem.  236,  570. 
Scott,  T.W.,  White,  I.G.,  and  Annison,  E.F.  (1962).  Biochem.  J.  83,  392. 
Seal,  U.S.  and  Binkley,  F.  (1957).  J.  Biol.  Chem.  128,  193. 
Sealock,  R.R.  and  Goodland,  R.L.  (1944).  J.  Am.  Chem.  Soc.  66,  507. 
Sealock,  R.R.  and  Livermore,  A.H.  (1949).  J.  Biol.  Chem.  177,  553. 
Sealock,  R.R.  and  White,  H.S.  (1949).  J.  Biol.  Chem,.  181,  393. 
Seaman,  G.R.  (1949).  Biol.  Bull.  96,  257. 
Seaman,  G.R.  (1952).  Arch.  Biochem.  Biophys.  39,  241. 
Seaman,  G.R.  (1954).  Arch.  Biochem.  Biophys.  48,  424. 
Seaman,  G.R.  (1956).  Exptl.  Parasitol.  5,  138. 

Seaman,  G.R.  and  Houlihan,  R.K.  (1950).  Arch.  Biochem.  26,  436. 
Seaman,  G.R.  and  Naschke,  M.D.  (1955).  J.  Biol.  Chem.  217,  1. 
Seeley,  H.W.  (1955).  In  "  Methods  in  Enzymology"  (S.P.  Colowick  and  N.O.  Kaplan, 

eds.).  Vol.  1,  p.  624.  Academic  Press,  New  York. 
Seelich,  F.  and  Letnansky,  K.  (1960).  Z.  Krebsforsch.  63,  236. 
Seelich,  F.  and  Letnansky,  K.  (1961).  Naturivissenschaften  48,  408. 
Seevers,  M.H.,  Shideman,  F.E.,  W^oods,  L.A.,  Weeks,  J.R.,  and  Kruse,  W.T.  (1950). 

J.  Pharmacol.  Exptl.  Therap.  99,  69. 
Segal,  H.L.  (1959).  J.  Am.  Chem.  Soc.  81,  4047. 
Segal,  H.L.  and  Boyer,  P.D.  (1953).  J.  Biol.  Chem.  204,  265. 
Segal,  H.L.  and  Brenner,  B.M.  (1960).  J.  Biol.  Chem.  235,  471. 
Segal,  H.L.,  Beattie,  D.S.,  and  Hopper,  S.  (1962).  J.  Biol.  Chem.  237,  1914. 
Seibert,  M.A.,  Kreke,  C.W.,  and  Cook,  E.S.  (1950).  Science  112,  649. 
Sekuzu,  I.,  Jurtshuk,  P.,  Jr.,  and  Green,  D.E.  (1963).  J.  Biol.  Chem.  238,  975. 
Sela,  M.  (1962).  J.  Biol.  Chem..  237,  418. 
Sela,  M.  and  Steiner,  L.A.  (1963).  Biochemistry  2,  416. 

Selim,  A.S.S.M.  and  Greenberg,  D.M.  (1960).  Biochim.  Biophys.  Acta  42,  211. 
Sellinger,  O.Z.  and  Weiler,  P.,  Jr.  (1963).  Biochem.  Pharmacol.  12,  989. 
Sen,  D.K.  (1959).  Nature  184,  459. 
SentheShanmuganathan,  S.  (1960).  Biochem.  J.  77,  619. 
Serif,  G.S.  and  Stewart,  C.J.  (1958).  Federation  Proc.  17,  309. 
Serif,  G.S.  and  Wick,  A.N.  (1958).  J.  Biol.  Chem.  233,  559. 
Serif,  G.S.,  Stewart,  C.J.,  Nakada,  H.L,  and  Wick,  A.N.  (1958).  Proc.  Soc.  Exptl.  Biol. 

Med.  99,  720. 
Sery,  T.W.  and  Furgiuele,  F.P.  (1961).  Am.  J.  Ophthalmol.  51,  42. 
Severens,  J.M.  and  Tanner,  F.W.  (1945).  J.  Bacteriol.  49,  383. 
Sguros,  P.L.  and  Hartsell,  S.E.  (1952  a).  J.  Bacteriol.  64,  811. 
Sguros,  P.L.  and  Hartsell,  S.E.  (1952  b).  J.  Bacteriol.  64,  821. 
Shacter,  B.  (1953).  Arch.  Biochem.  Biophys.  46,  324. 
Shacter,  B.  (1957).  J.  Natl.  Cancer  Inst.  18,  455. 
Shah,  V.K.  and  Ramakrishnan,  C.V.  (1963).  Enzymologia  26,  53. 


REFERENCES  1053 

Shanes,  A.M.  and  Brown,  D.E.S.  (1942).  J.  Cellular  Comp.  Physiol.  19,  1. 

Shannon,  L.M.,  Young,  R.H.,  and  Dudley,  C.  (1959).  Nature  183,  683. 

Shapiro,  D.M.  and  Gellhorn,  A.  (1951).  Cancer  Res.  11,  35. 

Shapiro,  D.M.,  Kream,  J.,  and  Dietrich,  L.S.  (1952).  Proc.  Soc.  Ezptl.  Biol.  Med.  81,  616. 

Shapiro,  D.M.,  Shils,  M.E.,  and  Dietrich,  L.S.  (1953).  Cancer  Res.  13,  703. 

Shapiro,  D.M.,  Dietrich,  L.S.,  and  Shils,  M.E.  (1956).  Cancer  Res.  16,  575. 

Shapiro,  D.M.,  Dietrich,  L.S.,  and  Shils,  M.E.  (1957).  Cancer  Res.  17,  600. 

Sharon,  N.  and  Lipmann,  F.  (1957).  Arch.  Biochem.  Biophys.  69,  219. 

Shaw,  E.  and  Woolley,  D.W.  (1952).  J.  Biol.  Chem.  194,  641. 

Shaw,  P.D.  and  Hager,  L.P.  (1959).  J.  Biol.  Chem.  234,  2565. 

Shaw,  W.H.R.  (1954).  Scieyice  120,  361. 

Shaw,  W.H.R.  and  Raval,  D.N.  (1961).  J.  Am.  Chem.  Soc.  83,  2866. 

Sheets,  R.F.  and  Hamilton,  H.E.  (1958).  J.  Lab.  Clin.  Med.  52,  138. 

Sheets,  R.F.,  Hamilton,  H.E.,  and  DeGowin,  E.L.  (1956  a).  Proc.  Soc.  Exptl.  Biol. 

Med.  91,  423. 
Sheets,  R.F.,  Hamilton,  H.E.,  DeGowin,  E.L.,  and  King,  R.L.  (1956  b).  J.  Appl. 

Physiol.  9,  145. 
Sheinin,  R.  and  Crocker,  B.F.  (1961).  Can.  J.  Biochem.  Physiol.  39,  55. 
Shemin,  D.  and  Kumin,  S.  (1952).  J.  Biol.  Chem.  198,  827. 
Sherman,  I.R.  and  Corley,  R.C.  (1952).  Federation  Proc.  11,  285. 
Shibasaki,  I.  and  Terui,  G.  (1953).  J.  Ferment.  Technol.  {Japan)  31,  238. 
Shichi,  H.  and  Hackett,  D.P.  (1962).  Phytochemistry  1,  131. 
Shideman,  F.E.  and  Rene,  R.M.  (1951a).  J.  Pharmacol.  Exptl.  Therap.  101,  32. 
Shideman,  F.E.  and  Rene,  R.M.  (1951b).  Am.  J.  Physiol.  166,  104. 
Shideman,  F.E.,  Woods,  L.A.,  and  Seevers,  M.H.  (1950a).  J.  Pharmacol.  Exptl.  Therap. 

98,  29. 

Shideman,  F.E.,  Woods,  L.A.,  and  Seevers,  M.H.  (1950b).  J.  Pharmacol.  Exptl.  Therap. 

99,  98. 

Shifrin,  S.  and  Kaplan,  N.O.  (1960).  Advan.  Enzymol.  22,  337. 

Shigeura,  H.T.  and  Gordon,  C.N.  (1962).  J.  Biol.  Chem.  237,  1937. 

Shiio,  I.,  Otsuka,  S.-I.,  and  Takahashi,  M.  (1961).  J.  Biochem.  (Tokyo)  49,  397. 

Shilo,  M.  (1957).  J.  Gen.  Microbiol.  16,  472. 

Shimazono,  H.  (1955).  J.  Biochem.  (Tokyo)  42,  321. 

Shimi,  I.R.  and  Nour  El  Dein,  M.S.  (1962).  Arch.  Mikrobiol.  41,  261. 

Shimomura,  0.,  Johnson,  F.H.,  and  Saiga,  Y.  (1961).  J.  Cellular  Comp.  Physiol.  58, 
113. 

Shimomura,  0.,  Johnson,  F.H.,  and  Saiga,  Y.  (1963).  J.  Cellular  Comp.  Physiol.  61, 
275. 

Shinohara,  K.  (1935).  J.  Biol.  Chem.  Ill,  435. 

Shkol'nik,  M.Y.  (1954).  Dokl.  Akad.  Nauk  SSSR  98,  443. 

Shore,  V.  and  Shore,  B.  (1960).  Am.  J.  Physiol.  198,  187. 

Shore,  V.  and  Shore,  B.  (1962).  Am.  J.  Physiol.  203,  15. 

Shore,  V.,  Shore,  B.,  and  Hsieh,  K.  (1959).  Federation  Proc.  18,  144. 

Shrago,  E.  and  Falcone,  A.B.  (1963  a).  Biochim.  Biophys.  Acta  67,  147. 

Shrago,  E.  and  Falcone,  A.B.  (1963  b).  Biochim.  Biophys.  Acta  73,  7. 

Shug,  A.L.,  Hamilton,  P.B.,  and  Wilson,  P.W.  (1956).  In  "Inorganic  Nitrogen  Meta- 
bolism" (W.D.  McElroy  and  H.B.  Glass,  eds.),  p.  344.  Johns  Hopkins  Press,  Bal- 
timore, Maryland. 


1054  REFEKENCES 

Shug,  A.L.,  Fraser,  D.,  and  Mahler,  H.R.  (1959).  Biochim.  Biophys.  Acta  36,  142. 

Shug,  A.L.,  Mahler,  H.R.,  and  Fraser,  D.  (1960).  Biochim.  Biophys.  Acta  42,  255. 

Shukuya,  R.  and  Schwert,  G.W.  (1960).  J.  Biol.  Chem.  235,  1658. 

Shull,  K.H.  (1959).  Nature  183,  259. 

Shulman,  A.  (1956).  Nature  177,  703. 

Shwartzman,  G.  and  Hift,  H.  (1951).  J.  Nutr.  44,  575. 

Sibly,  P.M.  and  Wood,  J.G.  (1951).  Australian  J.  Sci.  Res.  Ser.  B4,  500. 

Sie,  H.-G.  and  Fishman,  W.H.  (1954).  J.  Biol.  Chem.  209,  73. 

Siebert,  G.,  Carsiotis,  M.,  and  Plant,  G.W.E.  (1957).  J.  Biol.  Che?n.  226,  977. 

Siebert,  G.,  Schmitt,  A.,  and  Trazler,  G.  (1963).  Z.  Physiol.  Chem.  332,  160. 

Siegel,  L.  and  Englard,  S.  (1962).  Biochiyn.  Biophys.  Acta  64,  101. 

Siegelman,  H.W.,  Chow,  C.T.,  and  Biale,  J.B.  (1958).  Plant  Physiol.  33,  403. 

Siekevitz,  P.,  Low,  H.,  Ernster,  L.,  and  Lindberg,  O.  (1958).  Biochim.  Biophys.  Acta 

29,  378. 
Sih,  C.J.,  Hamilton,  P.B.,  and  Knight,  S.G.  (1958).  J.  Bacteriol.  75,  623. 
Silber,  R.,  Huennekens,  F.M.,  and  Gabrio,  B.W.  (1962).  Arch.  Biochem.  Biophys.  99, 

328. 
Silberman,  H.R.  and  Wyngaarden,  J.B.  (1961).  Biochim.  Biophys.  Acta  47,  178. 
Sillen,  L.G.  (1949).  Acta  Chem.  Scand.  3,  539. 
Silverman,  J.L.  (1963).  Biochim.  Biophys.  Acta  78,  94. 
Simola,  P.E.  and  Kosunen,  T.  (1938).  Suomen  Kemistilehti  B11,  22. 
Simon,  E.W.  (1957).  J.  Exptl.  Botany  8,  20. 
Simon,  E.W.  (1959).  J.  Exptl.  Botany  10,  125. 

Simon,  L.N.  and  Myers,  T.C.  (1961).  Biochim.  Biophys.  Acta  51,  178. 
Simonds,  J.P.  and  Hepler,  O.E.  (1945).  Arch.  Pathol.  39,  103. 
Simonsen,  D.H.  and  van  Wagtendonk,  W.J.  (1956).  J.  Gen.  Microbiol.  15,  39. 
Simpson,  F.J.  (1960).  Can.  J.  Biochem.  Physiol.  38,  115. 
Simpson,  R.B.  (1961).  J.  Am.  Chem.  Sac.  83,  4711. 
Singer,  B.  and  Fraenkel-Conrat,  H.  (1962).  Biochemistry  1,  852. 
Singer,  T.P.  (1948).  J.  Biol.  Chem.  174,  11. 

Singer,  T.P.  and  Barron,  E.S.G.  (1944).  Proc.  Soc.  Exptl.  Biol.  Med.  56,  120. 
Singer,  T.P.  and  Barron,  E.S.G.  (1945).  J.  Biol.  Chem.  157,  241. 
Singer,  T.P.  and  Hofstee,  B.H.J.  (1948  a).  Arch.  Biochem.  18,  229. 
Singer,  T.P.  and  Hofstee,  B.H.J.  (1948  b).  Arch.  Biochem.  18,  245. 
Singer,  T.P.  and  Kearney,  E.B.  (1950).  Arch.  Biochem.  27,  348. 
Singer,  T.P.  and  Lara,  F.J.S.  (1958).  Proc.  Intern.  Symp.  Enzyme  Chem.,  Tokyo  Kyoto, 

1957  p.  330.  Maruzen,  Tokyo. 
Singer,  T.P.,  and  Lusty,  C.J.  (1960).  Biochem.  Biophys.  Res.  Commun.  2,  276. 
Singer,  T.P.,  Massey,  V.,  and  Kearney,  E.B.  (1956  a).  Biochim.  Biophys.  Acta  19,  200. 
Singer,  T.P.,  Kearney,  E.B.,  and  Bernath,  P.  (1956  b).  J.  Biol.  Chem.  223,  599. 
Singer,  T.P.,  Kearney,  E.B.,  and  Massey,  V.  (1957).  Advan.  Enzymol.  18,  65. 
Sipos,  J.C,  Sawatzky,  H.,  and  Wright,  G.F.  (1955).  J.  Am.  Chem.  Soc.  77,  2759. 
Siu,  P.M.L.  and  Wood,  H.G.  (1962).  J.  Biol.  Chem.  237,  3044. 
Sivaramakrishnan,  V.M.  and  Sarma,  P.S.  (1954).  Biochim.  Biophys.  Acta  14,  579. 
Sizer,  I.W.  (1942  a).  J.  Gen.  Physiol.  25,  399. 
Sizer,  I.W.  (1942  b).  J.  Biol.  Chem.  145,  405. 
Sizer,  I.W.  (1945).  J.  Biol.  Chem.  160,  547. 
Sizer,  I.W.  and  Tytell,  A. A.  (1941).  J.  Biol.  Chem.  138,  631. 


REFERENCES  1055 

Skarnes,  R.C.  and  Watson,  D.W.  (1955).  J.  Bacteriol.  70,  110. 

Skipper,  H.E.,  Mitchell,  J.H.,  Jr.,  and  Bennett,  L.L.,  Jr.  (1950).  Cancer  Res.  10,  510. 

Skoda,  J.,  Kara,  J.,  Sormova,  Z.,  and  Sorm,  F.  (1959).  Biochim.  Bilphys.  Acta  33,  579. 

Skou,  J.C.  (1963).  Biochem.  Biophys.  Res.  Commun.  10,  79. 

Slater,  E.C.  (1949).  Biochem.  J.  45,  130. 

Slater,  E.C.  and  Bonner,  W.D.,  Jr.  (1952).  Biochem.  J.  52,  185. 

Slater,  E.C.  and  Holton,  F.A.  (1954).  Biochem.  J.  56,  28. 

Slater,  T.F.  (1959).  Nature  183,  1679. 

Slechta,  L.  (1960).  Biochem.  Biophys.  Res.  Commun.  3,  596. 

Slein,  M.W.,  Cori,  G.T.,  and  Cori,  C.F.  (1950).  J.  Biol.  Chem.  186,  763. 

Slingerland,  D.W.  (1955).  J.  Clin.  Endocrinol.  Metab.  15,  131. 

Sloane,  N.H.  and  Boggiano,  E.M.  (1960).  Arch.  Biochem.  Biophys.  87,  217. 

Smalt,  M.A.,  Kreke,  C.W.,  and  Cook,  E.S.  (1957).  J.  Biol.  Chem.  IIA,  999. 

Smillie,  R.M.  (1956).  Australian  J.  Biol.  Sci.  9,  81. 

Smith,  C.G.  and  Bernstein,  I. A.  (1961).  Biochim.  Biophys.  Acta  52,  184. 

Smith,  C.G.,  Lummis,  W.L.,  and  Grady,  J.E.  (1959).  Cancer  Res.  19,  847. 

Smith,  D.E.,  Czarnetzky,  E.J.,  and  Mudd,  S.  (1936).  Am.  J.  Med.  Sci.  192,  790. 

Smith,  E.L.  (1958).  J.  Biol.  Chem.  233,  1392. 

Smith,  E.L.,  Lumry,  R.,  and  Polglase,  W.J.  (1951).  J.  Phys.  Colloid  Chem.  55,  125. 

Smith,  E.L.,  Davis,  N.C.,  Adams,  E.,  and  Spackman,  D.H.  (1954  a).  In  "The  Mech- 
anism of  Enzyme  Action"  (W.D.  McElroy  and  H.B.  Glass,  eds.),  p.  291.  Johns 
Hopkins  Press,  Baltimore,  Maryland. 

Smith,  E.L.,  Kimniel,  J.R.,  and  Brown,  D.M.  (1954b).  J.  Biol.  Chem.  207,  533. 

Smith,  E.L.,  Kimmel,  J.R.,  Brown,  D.M.,  and  Thompson,  E.O.P.  (1955).  J.  Biol. 
Chem.  215,  67. 

Smith,  F.  and  Spriestersbach,  D.  (1954).  Nature  174,  466. 

Smith,  H.M.  (1958).  J.  Cellular  Camp.  Physiol.  51,  161. 

Smith,  J.E.  (1963).  Can.  J.  Microbiol.  9,  345. 

Smith,  J.T.  (1963  a).  Biochem.  J.  87,  40p. 

Smith,  J.T.  (1963  b).  Nature  197,  900. 

Smith,  L.  and  BaltschefFsky,  M.  (1959).  J.  Biol.  Chem.  234,  1575. 

Smith,  L.H.,  Jr.  and  Sullivan,  M.  (1960).  Biochim.  Biophys.  Acta  39,  554. 

Smith,  M.E.  and  Greenberg,  D.M.  (1957).  J.  Biol.  Chem.  226,  317. 

Smith,  P.F.  (1959).  J.  Bacteriol.  77,  682. 

Smith,  P. J.C.  (1960  a).  Biochem.  J.  76,  508. 

Smith,  P.J.C.  (1960  b).  Biochem.  J.  76,  514. 

Smith,  S.E.  (1960  a).  Brit.  J.  Pharmacol.  15,  319. 

Smith,  S.E.  (1960  b).  Ciba  Found.  Symp.,  Adrenergic  Mechanisms  p.  25. 

Smith,  S.E.  (1963).  Brit.  J.  Pharmacol.  20,  178. 

Smyth,  D.H.  (1940).  Biochem.  J.  34,  1046. 

Sneed,  M.C.  and  Brasted,  R.C.  (1955).  "Comprehensive  Inorganic  Chemistry,"  Vol.  4, 
Van  Nostrand,  Princeton,  New  Jersey. 

Snodgrass,  P.J.  and  Hoch,  F.L.  (1959).  Federation  Proc.  18,  326. 

Snodgrass,  P.J.,  Vallee,  B.L.,  and  Hoch,  F.L.  (1960).  J.  Biol.  Chem.  235,  504. 

Snoswell,  A.M.  (1959).  Australian  J.  Exptl.  Biol.  Med.  Sci.  37,  49. 

Snoswell,  A.M.  (1962).  Biochim.  Biophys.  Acta  60,  143. 

Snyder,  R.,  Vogel,  W.H.,  and  Schulman,  M.P.  (1963).  Biochem.  Biophys.  Res.  Com- 
mun.  13,  335. 


1056  REFERENCES 

Snyder,  S.H.,  Silva,  O.L.,  and  Kies,  M.W.  (1961).  J.  Biol.  Chem.  236,  2996. 

Sohler,  M.R.,  Seibert,  M.A.,  Kreke,  C.W.,  and  Cook,  E.S.  (1952).  J.  Biol.  Chem.  198, 

281. 
Sokoloff,  B.,  Saelhof,  C,  Kato,  S.,  Chemelin,  I.,  and  Beach,  J.  (1956).  Growth  20,  265. 
Soldatenkov,  S.V.  and  Mazurova,  T.A.  (1957).  Biochemistry  (USSR)  (English  Transl.) 

11,  345. 
Solomon,  J.B.  (1958).  Biochem.  J.  70,  529. 
Sols,  A.  (1956).  Biochim.  Biophys.  Acta  19,  144. 
Sols,  A.  and  Crane,  R.K.  (1953).  Federation  Proc.  12,  271. 
Sols,  A.  and  Crane,  R.K.  (1954).  J.  Biol.  Chem.  206,  925. 
Sols,  A.  and  de  la  Fuente,  G.  (1957).  Biochim.  Biophys.  Acta  24,  206. 
Sols,  A.,  Cadenas,  E.,  and  Alvarado,  F.  (1960  a).  Science  131,  297. 
Sols,  A.,  Heredia,  C.F.,  and  Ruiz-Amil,  M.  (1960  b).  Biochem.  Biophys.  Res.  Commun. 

2,  126. 
Solvonuk,  P.F.  and  Collier,  H.B.  (1955).  Can.  J.  Biochem.  Physiol.  33,  38. 
Solvonuk,  P.F.,  McRae,  S.C,  and  Collier,  H.B.  (1956).  Can.  J.  Biochem.  Physiol.  34, 

481. 
Somers,  E.  (1959).  Nature  184,  475. 

Sondheimer,  E.  and  Griffin,  D.H.  (1960).  Science  131,  672. 
Soodak,  M.  and  Cerecedo,  L.R.  (1944).  J.  Am.  Chem.  Soc.  66,  1988. 
Sophianopoulos,  A.J.  and  Corley,  R.C.  (1959).  Federation  Proc.  18,  328. 
Sorbo,  B.  (1963).  Acta  Chem.  Scand.  17,  S  107. 
Sorbo,  B.H.  (1951).  Acta  Chem.  Scand.  5,  724. 

Sorsby,  A.,  Newhouse,  J.P.,  and  Lucas,  D.R.  (1957).  Brit.  J.  Ophthalmol.  41,  309. 
Sourkes,  T.L.  (1954).  Arch.  Biochem.  Biophys.  51,  444. 
Sourkes,  T.L.  and  Towsend,  E.  (1955).  Can.  J.  Biochem.  Physiol.  33,  735. 
Speck,  J.F.,  Moulder,  J.W.,  and  Evans,  E.A.,  Jr.  (1946).  J.  Biol.  Chem.  164,  119. 
Spector,  A.  (1963).  J.  Biol.  Chem..  238,  1353. 
Spector,  W.G.  (1953).  Proc.  Roy.  Soc.  B141,  268. 
Speer,  H.L.  and  Jones,  R.F.  (1964).  Physiol.  Plantarum  17,  287. 
Spencer,  A.F.  and  Lowenstein,  J.M.  (1962).  J.  Biol.  Chem.  237,  3640. 
Spencer,  B.  and  Williams,  R.T.  (1951).  Biochem.  J.  48,  538. 
Spencer,  D.  (1959).  Australian  J.  Biol.  Sci.  12,  181. 
Spencer,   H.C.,  Rowe,  V.K.,   and  McCollister,   D.D.  (1950  a).  J.  Pharmacol   Exptl. 

Therap.  98,  30. 
Spencer,   H.C.,  Rowe,  V.K.,  and  McCollister,   D.D.  (1950  b).  J.   Pharmacol.  Exptl. 

Therap.  99,  57. 
Spencer,  R.P.  and  Brody,  K.R.  (1964).  Am.  J.  Physiol.  206,  653. 
Spencer,  R.P.  and  Knox,  W.E.  (1962).  Arch.  Biochem.  Biophys.  96,  115. 
Spencer,  R.P.,  Valentine,  R.J.,  and  Field,  J.B.  (1956).  Acta  Pharmacol.  Toxicol.  12, 

196. 
Spensley,  P.C.  and  Rogers,  H.J.  (1954).  Nature  173,  1190. 
Speyer,  J.F.  and  Dickman,  S.R.  (1956).  J.  Biol.  Chem.  220,  193. 
Spiegelman,  S.,  Kamen,  M.D.,  and  Sussman,  M.  (1948).  Arch.  Biochem.  18,  409. 
Spiro,  R.G.  (1958).  J.  Biol.  Chem.  233,  546. 
Spizizen,  J.  (1943).  J.  Infect.  Diseases  73,  222. 
Spratt,  N.T.  (1950).  Biul.  Bull.  99,  120. 
Srinivasan,  P.R.  and  Sprinson,  D.B.  (1959).  J.  Biol.  Chem.  234,  716. 


REFERENCES  1057 

Srinivasan,  P.R.  Brunfaut,  M.,  and  Errera,  M.  (1964).  Exptl.  Cell.  Res.  34,  61. 

Stadtman,  E.R.  (1955).  In  "Methods  in  Enzymology"  (S.P.  Colowick  and  N.O.  Ka- 
plan, eds.),  Vol.  1,  p.  596.  Academic  Press,  New  York. 

Stadtman,  E.R.,  Cohen,  G.N.,  and  LeBras,  G.  (1961).  Ann.  N.  Y.  Acad.  Sci.  94,  952. 

Staemmler,  M.  (1956).  Arch.  Pathol.  Anat.  Physiol.  328,  1. 

Stafford,  H.A.  (1956).  Plant  Physiol.  31,  135. 

Stalder,  K.  (1958).  Z.  Physiol.  Chem.  311,  221. 

Stalder,  K.  (1959).  Z.  Physiol.  Chem.  314,  205. 

Stanley,  R.G.  and  Conn,  E.E.  (1957).  Plant  Physiol.  32,  412. 

Stare,  F.J.  and  Baumann,  C.A.  (1936).  Pwc.  Roy.  Soc.  B121,  338. 

Stare,  F.J.  and  Baumann,  C.A.  (1939).  Cold  Spring  Harbor  Symp.  Quant.  Biol.  7,  227. 

Stare,  F.J.  and  Baumann,  C.A.  (1940).  J.  Biol.  Chem.  133,  453. 

Stare,  F.J.,  Lipton,  M.A.,  and  Goldinger,  J.M.  (1941).  J.  Biol.  Chem.  141,  981. 

Stark,  J.B.,  Goodban,  A.E.,  and  Owens,  H.S.  (1951).  Anal.  Chem.  23,  413. 

Starr,  J.L.  and  Goldthwait,  D.A.  (1963).  J.  Biol.  Chem.  238,  682. 

Stauff,  J.  and  Uhlein,  E.  (1958).  Biochem.  Z.  329,  549. 

Stearn,  A.E.  (1935).  J.  Oen.  Physiol.  18,  301. 

Steberl,  E.A.,  Wasson,  G.W.,  and  Porter,  J.W.  (1960).  Biochem.  Biophys.  Res.  Com- 
mun.  2,  174. 

Stedman,  R.L.  and  Kravitz,  E.  (1955).  Arch.  Biochem.  Biophys.  59,  260. 

Stedman,  R.L.,  Engel,  S.L.,  and  Bilse,  I.M.  (1954).  J.  Am.  Pharm.  Assoc,  Sci.  Ed. 
43,  622. 

Steel,  K.J.  (1960).  J.  Pharm.  Pharmacol.  12,  59. 

Steele,  A.B.,  Boese,  A.B.,  and  Dull,  M.F.  (1949).  J.  Org.  Chem.  14,  460. 

Stein,  E.,  Magin,  J.,  and  Kleinfeld,  M.  (1960).  Am.  J.  Physiol.  199,  460. 

Steinberg,  D.  and  Fredrickson,  D.S.  (1955).  Proc.  Soc.  Exptl.  Biol.  Med.  90,  232. 

Steinberg,  D.  and  Vaughan,  M.  (1956).  Arch.  Biochem.  Biophys.  65,  93. 

Steiner,  D.F.,  Rauda,  V.,  and  Williams,  R.H.  (1961).  J.  Biol.  Chem.  236,  299. 

Stephenson,  M.L.,  Thimann,  K.V.,  and  Zamecnik,  P.C.  (1956).  Arch.  Biochem.  65, 
194. 

Stern,  J.R.  (1948).  Biochem.  J.  43,  616. 

Stern,  J.R.  (1956).  J.  Biol.  Chem.  ll^,  33. 

Stem,  J.R.  and  del  Campillo,  A.  (1956).  J.  Biol.  Chem.  218,  985. 

Stern,  J.R.  and  Friedman,  D.L.  (1960).  Biochem.  Biophys.  Res.  Commun.  2,  82. 

Stevens,  CO.  and  Henderson,  L.M.  (1959).  J.  Biol.  Chem.  234,  1188. 

Stewart,  H.B.  (1957).  Federation  Proc.  16,  256. 

Steyn-Parve,  E.P.  (1954).  Biochim.  Biophys.  Acta  14,  440. 

Stickland,  R.G.  (1959  a).  Biochem.  J.  73,  654. 

Stickland,  R.G.  (1959  b).  Biochem.  J.  73,  646. 

Stjernholm,  R.  and  Wood,  H.G.  (1961).  Proc.  Natl.  Acad.  Sci.  U.S.  47,  303. 

Stockell,  A.  (1959).  J.  Biol.  Chem.  234,  1293. 

Stockell,  A.  and  Smith,  E.L.  (1957).  J.  Biol.  Chem.  227,  1. 

Stocken,  L.A.  (1947).  Biochem.  J.  41,  358. 

Stoerk,  H.C.  (1947).  J.  Biol.  Chem.  171,  437. 

Stoerk,  H.C.  (1950).  Ann.  N.Y.  Acad.  Sci.  52,  1302. 

Stone,  B.A.  and  Hochster,  R.M.  (1956).  Can.  J.  Microbiol.  2,  623. 

Stone,  C.A.,  Ross,  C.A.,  Wenger,  H.C,  Ludden,  C.T.,  Blessing,  J.A.,  Totaro,  J.A., 
and  Porter,  CC  (1962).  J.  Pharmacol.  Exptl.  Therap.  136,  80. 


1058  REFERENCES 

Stone,  J.E.  and  Potter,  V.R.  (1957).  Cancer  Res.  17,  800. 

Stoneman,  F.,  Shideman,  F.E.,  and  Rathbun,  R.C.  (1951).  Federation  Proc.  10,  337. 

Stoppani,  A.O.M.  (1946).  Anales  Asoc.  Quim.  Arg.  34,  33. 

Stoppani,  A.O.M.  (1948).  Enzymologia  13,  165. 

Stoppani,  A.O.M.  and  Brignone,  J. A.  (1956).  Biochem.  J.  64,  196. 

Stoppani,  A.O.M.  and  Brignone,  J.A.  (1957).  Arch.  Biochem.  Biophys.  68,  432. 

Stoppani,  A.O.M.  and  Milstein,  C.  (1957  a).  Biochem.  J.  67,  406. 

Stoppani,  A.O.M.  and  Milstein,  C.  (1957  b).  Biochim.  Biophys.  Acta  24,  655. 

Stoppani,  A.O.M.,  Actis,  A.S.,  Deferrari,  J.O.,  and  Gonzalez,  E.L.  (1953).  Biochem.  J. 

54,  378. 
Stoppani,  A.O.M.,  Conches,  L.,  de  Favelukes,  S.L.S.,  and  Sacerdote,  F.L.  (1958  a). 

Biochem.  J.  70,  438. 
Stoppani,  A.O.M.,  de  Favelukes,   S.L.S.,  and  Conches,  L.  (1958  b).  Arch.  Biochem. 

Biophys.  73,  453. 
Storey,  I.D.E.  (1964).  Biochem.  J.  90,  15p. 

Stracher,  A.  and  Chan,  P.C.  (1961).  Arch.  Biochem.  Biophys.  95,  435. 
Straessle,  R.  (1951).  J.  Am.  Chem.  Soc.  73,  504. 
Straessle,  R.  (1954).  J.  Am.  Chem.  Soc.  76,  3138. 
Straub,  F.B.  (1937).  Z.  Physiol.  Chem.  249,  189. 
Strauss,  N.  and  Jann,  G.J.  (1956).  J.  Bacteriol.  71,  439. 
Strecker,  H.J.  (1960).  J.  Biol.  Chem..  235,  3218. 
Strecker,  H.J.  and  Korkes,  S.  (1952).  J.  Biol.  Chem.  196,  769. 

Strength,  D.R.,  Mondy,  N.I.,  and  Daniel,  L.J.  (1954).  Arch.  Biochem.  Biophys.  48,  141. 
Strickland,  K.F.  (1954).  Can.  J.  Biochem.  Physiol.  32,  50. 
Strickler,  J.C.  and  Kessler,  R.H.  (1963).  Am.  J.  Physiol.  205,  117. 
Stricks,  W.  and  Kolthoff,  I.M.  (1953).  J.  Am.  Chem.  Soc.  75,  5673. 
Stricks,  W.,  Kolthoff,  I.M.,  and  Heyndrickx,  A.  (1954).  J.  Am.  Chem.  Soc.  76,  1515. 
Strittmatter,  C.F.  (1959).  J.  Biol.  Chem.  234,  2794. 
Strittmatter,  P.  (1959).  J.  Biol.  Chem.  234,  2661. 
Strittmatter,  P.  (1961a).  J.  Biol.  Chem.  236,  2329. 
Strittmatter,  P.  (1961b).  J.  Biol.  Chem.  236,  2336. 
Strittmatter,  P.  and  Velick,  S.F.  (1956).  J.  Biol.  Chem.  221,  253. 
Strohman,  R.C.  and  Samorodin,  A.J.  (1962).  J.  Biol.  Chem.  237,  363. 
Strominger,  J.L.  and  Mapson,  L.W.  (1957).  Biochem.  J.  66,  567. 
Stumpf,  P.K.  (1948).  J.  Biol.  Chem.  176,  233. 
Stumpf,  P.K.  and  Barber,  G.A.  (1956).  Plant  Physiol.  31,  304. 
Stumpf,  P.K.,  Zarudnaya,  K.,  and  Green,  D.E.  (1947).  J.  Biol.  Chem.  167,  817. 
Stumpf.  P.K.,  Loomis,  W.D.,  and  Michelson,  C.  (1951).  Arch.  Biochem.  30,  126. 
Subramanian,  V.,  Stent,  H.B.,  and  Walker,  T.K.  (1929).  J.  Chem.  Soc.  p.  2485. 
Sumizu,  K.,  Yoshikawa,  M.,  and  Tanaka,  S.  (1961).  J.  Biochem.  (Tokyo)  50,  538. 
Summers,  W.A.  (1957).  Exptl.  Parasitol.  6,  194. 
Sundaram,  S.  and  Sarma,  P.S.  (1960  a).  Enzymologia  21,  219. 
Sundaram,  S.  and  Sarma,  P.S.  (1960b).  Biochem.  J.  77,  465. 

Sundaram,  T.K.,  Rajagopalan,  K.V.,  and  Sarma,  P.S.  (1958).  Biochem.  J.  70,  196. 
Siipfle,  K.  (1923).  Arch.  Hyg.  Bakteriol.  93,  252. 
Surtshin,  A.  (1957).  Am.  J.  Physiol.  190,  271. 
Sussman,  A.,  Holton,  R.,  and  von  Boventer-Heidenhain,  B.  (1958).  Arch.  Mikrobiol. 

29,  38. 


REFERENCES  1059 

Sutton,  C.R.  and  King,  H.K.  (1959).  Biochem.  J.  73,  43p. 

Sutton,  C.R.  and  King,  H.K.  (1962).  Arch.  Biochem.  Biophys.  96,  360. 

Sutton,  W.B.  (1954).  J.  Biol.  Chem.  210,  309. 

Suzuoki,  T.  (1912).  "Morphologie  der  Nierensekretion  unter  pathologischen  Bedin- 

gungen."  Fischer,  Jena. 
Suzuoki,  Z.  and  Suzuoki,  T.  (1951).  J.  Biochem.  (Tokyo)  38,  237. 
Suzuoki,  Z.  and  Suzuoki,  T.  (1954).  Nature  173,  83. 
Swan,  J.M.  (1959).  In  "Sulfur  in  Proteins"  (R.  Benesch  et  al.,  eds.),  p.  3.  Academic 

Press,  New  York. 
Swenson,  A.D.  and  Boyer,  P.D.  (1957).  J.  Am.  Chem.  Soc.  79,  2174. 
Swensson,  A.,  Lundgren,  K.-D.,  and  Linstrom,  O.  (1959).  A.M. A.  Arch.  Ind.  Health 

20,  432. 
Swim,  H.E.  and  Krampitz,  L.O.  (1954).  J.  Bacteriol.  67,  419. 

Swingle,  K.F.,  Axelrod,  A.E.,  and  Elvehjem,  C.A.  (1942).  J.  Biol.  Chem.  145,  581. 
Szabolcsi,  G.  and  Biszku,  E.  (1961).  Biochim.  Biophys.  Acta  48,  335. 
Szabolcsi,  G.,  Biszku,  E.,  and  Sajgo,  M.  (1960).  Acta  Physiol.  Acad.  Sci.  Hung.  17,  183. 
Szep,  O.  (1940).  Biochem.  Z.  307,  79. 

Tabachnick,  M.,  Srere,  P. A.,  Cooper,  J.,  and  Racker,  E.  (1958).  Arch.  Biochem.  Bio- 
phys. 74,  315. 

Tabor,  H.  and  Wyngarden,  L.  (1959).  J.  Biol.  Chem.  234,  1830. 

Tachibana,  S.,  Katagiri,  H.,  and  Yamada,  H.  (1958).  Proc.  Intern.  Symp.  Enzyme 
Chem.,  Tokyo  Kyoto  1957  p.  154.  Maruzen,  Tokyo. 

Tager,  J.M.  (1963).  Biochim,.  Biophys.  Acta  73,  341. 

Takagaki,  G.,  Hirano,  S.,  and  Tsukada,  Y.  (1957).  Arch.  Biochem.  Biophys.  68,  196. 

Takagaki,  G.,  Hirano,  S.,  and  Tsukada,  Y.  (1958).  J.  Biochem.  {Tokyo)  45,  41. 

Takagi,  Y.  and  Horecker,  B.L.  (1957).  J.  Biol.  Chem.  225,  77. 

Takahashi,  H.  and  Nomura,  M.  (1952).  J.  Agr.  Chem.  Soc.  Japan  26,  99. 

Takahashi,  N.  (1960  a).  J.  Biochem.  (Tokyo)  47,  230. 

Takahashi,  N.  (1960  b).  J.  Biochem.  (Tokyo)  48,  691. 

Takebe,  I.  (1961).  J.  Biochem.  (Tokyo)  50,  245. 

Takemori,  A.E.  and  Mannering,  G.J.  (1958).  J.  Pharmacol.  Exptl.  Therap.  123,  171. 

Takenaka,  Y.  and  Schwert,  G.W.  (1956).  J.  Biol.  Chem.  112,  157. 

Takeuchi,  T.  (1933).  J.  Biochem.  (Tokyo)  17,  47. 

Talalay,  P.  and  Dobson,  M.M.  (1953).  J.  Biol.  Chem.  205,  823. 

Talalay,  P.  and  Marcus,  P.I.  (1956).  J.  Biol.  Chem.  218,  675. 

Tanada,  T.  (1956).  Plant  Physiol.  31,  403. 

Tanaka,  K.  (1961).  J.  Biochem.  (Tokyo)  50,  102. 

Tanaka,  M.,  Kato,  Y.,  and  Kinoshita,  S.  (1961).  Biochem.  Biophys.  Res.  Commun.  4, 
114. 

Tanaka,  S.  (1938).  J.  Biochem.  (Tokyo)  28,  119. 

Tanaka,  S.,  Shimizu,  T.,  Ishishita,  Y.,  Koike,  M.,  and  Hiramitsu,  S.  (1952).  Nisshin 
Igaku  39,  220. 

Tanaka,  Y.  (1960).  J.  Biochem.  (Tokyo)  47,  1. 

Tapley,  D.F.  (1956).  J.  Biol.  Chem.  Ill,  325. 

Tarr,  H.L.A.  (1959).  Can.  J.  Biochem.  Physiol.  37,  961. 

Tashian,  R.E.  (1961).  Metab.,  Clin.  Exptl.  10,  393. 

Tata,  J.R.  (1959).  Biochim.  Biophys.  Acta  35,  567. 


1060  REFERENCES 

Tata,  J.R.  (1960).  Biochem.  J.  77,  214. 

Taylor,  E.L.,  Meriwether,  B.P.,  and  Park,  J.H.  (1963).  J.  Biol.  Chem.  238,  734. 

Taylor,  E.S.  and  Gale,  E.F.  (1945).  Biochem.  J.  39,  52. 

Taylor,  F.J.  (1960).  Proc.  Roy.  Soc.  B151,  483. 

Taylor,  H.D.  and  Austin,  J.H.  (1918).  J.  Exptl.  Med.  27,  635. 

Taylor,  J.J.  and  Johnson,  C.E.  (1962).  Enzymologia  25,  87. 

Telkka,  A.  and  Mustakallio,  K.K.  (1954).  Experientia  10,  216. 

Terner,  C.  (1951).  Biochem.  J.  50,  145. 

Terner,  C.  (1954).  Biochem.  J.  56,  471. 

Terner,  C,  Eggleston,  L.V.,  and  Krebs,  H.A.  (1950).  Biochem.  J.  47,  139. 

Theis,  F.V.  and  Freeland,  M.R.  (1940).  Arch.  Surg.  40,  190. 

Theorell,  H.  (1956).  In  "Currents  in  Biochemical  Research"  (D.E.  Green,  ed.),  p.  275. 

Wiley  (Interscience),  New  York. 
Theorell,  H.  and  Bonnichsen,  R.  (1951).  Acta  Chem.  Scand.  5,  329. 
Theorell,  H.,  Yagi,  K.,  Ludwig,  G.D.,  and  Egami,  F.  (1957).  Nature  180,  922. 
Thimann,  K.V.  and  Bonner,  W.D.,  Jr.  (1948).  Am.  J.  Botany  35,  271. 
Thimann,  K.V.  and  Bonner,  W.D.,  Jr.  (1949).  Am.  J.  Botany  36,  214. 
Thoelen,  H.  and  Pletscher,  A.  (1953).  Helv.  Physiol.  Pharmacol.  Acta  11,  64. 
Thomas,  C.A.  (1954).  J.  Am.  Chem.  Soc.  76,  6032. 
Thomas,  G.W.  and  Cook,  E.S.  (1947).  J.  Bacteriol.  54,  527. 
Thomas,  K.  and  Kalbe,  H.  (1953).  Z.  Physiol.  Chem.  293,  239. 
Thomas,  K.  and  Stalder,  K.  (1958).  Z.  Physiol.  Chem.  313,  22. 
Thomas,  K.  and  Stalder,  K.  (1959).  Z.  Physiol.  Chem.  315,  256. 
Thompson,  E.O.P.  (1954).  J.  Biol.  Chem.  207,  563. 
Thompson,  R.H.S.  and  Whittaker,  V.P.  (1947).  Biochem.  J.  41,  342. 
Thompson,  R.L.  (1947).  J.  Immunol.  55,  345. 
Thompson,  W.  and  Dawson,  R.M.C.  (1964).  Biochem.  J.  91,  237. 
Thomson,  J.F.  (1959).  Proc.  Soc.  Exptl.  Biol.  Med.  101,  1. 
Thomson,  J.F.  and  Sato,  T.R.  (1960).  Arch.  Biochem.  Biophys.  89,  139. 
Thorn,  M.B.  (1953).  Biochem.  J.  54,  540. 
Thorn,  M.B.  (1959).  Biochem.  J.  71,  23p. 

Thorne,  C.J.R.  and  Kaplan,  N.O.  (1963).  J.  Biol.  Chem.  238,  1861. 
Thome,  K.J.I,  and  Kodicek,  E.  (1962).  Biochim.  Biophys.  Acta  59,  295. 
Thunberg,  T.  (1909).  Skand.  Arch.  Physiol.  11,  430. 
Thunberg,  T.  (1933).  Biochem.  Z.  258,  48. 
Thyagarajan,  B.S.  (1958).  Chem.  Rev.  58,  439. 
Tietz,  A.  (1961).  J.  Lipid  Res.  2,  182. 

Tietze,  F.  and  Klotz,  I.M.  (1952).  Arch.  Biochem.  Biophys.  35,  355. 
Tinacci,  F.  (1953).  Boll.  Soc.  Ital.  Biol.  Sper.  29,  1457. 
T'ing-seng,  H.  (1961a).  Biochemistry  (USSR)  (English  Transl.)  25,  870. 
T'ing-seng,  H.  (1961b).  Biochemistry  {USSR)  {English  Transl.)  26,  493. 
Tokuyama,  K.  (1959).  J.  Biochem.  {Tokyo)  46,  1453. 

Tokuyama,  K.  and  Dawson,  C.R.  (1962).  Biochim.  Biophys.  Acta  56,  427. 
Tolba,  M.K.  and  Eskarous,  J.K.  (1962).  Arch.  Mikrobiol.  41,  393. 
Tolba,  M.K.  and  Salama,  A.M.  (1962).  Arch.  Mikrobiol.  43,  349. 
Tong,  W.,  Taurog,  A.,  and  ChaikofF,  I.L.  (1957).  J.  Biol.  Chem.  Ill,  713. 
Tonhazy,  N.E.  and  Pelczar,  M.J.,  Jr.  (1953).  J.  Bacteriol.  65,  368. 
Tonomura,  Y,  and  Furuya,  K.  (1960).  J.  Biochem.  {Tokyo)  48,  899. 


REFERENCES  1061 

Tonomura,  Y.  and  Kitagawa,  S.  (1957).  Biochim.  Biophys.  Acta  26,  15. 

Tonomura,  Y.  and  Kitagawa,  S.  (1960).  Biochim.  Biophys.  Acta  40,  135. 

Tonomura,  Y.  and  Yoshimura,  J.  (1962).  J.  Biochem.  (Tokyo)  51,  259. 

Tonomura,  Y.,  Sekiya,  K.,  and  Imamura,  K.  (1963  a).  Biochim.  Biophys.  Acta  69, 
296. 

Tonomura,  Y.,  Yoshimura,  J.,  and  Olmishi,  T.  (1963  b).  Biochim.  Biophys.  Acta  70, 
698. 

Topper,  Y.J.  and  Lipton,  M.M.  (1953).  J.  Biol.  Chem.  203,  135. 

Torda,  C.  and  Wolff,  H.G.  (1944  a).  Proc.  Soc.  Exptl.  Biol.  Med.  57,  234. 

Torda,  C.  and  Wolff,  H.G.  (1944  b).  Proc.  Soc.  Exptl.  Biol.  Med.  57,  236. 

Totter,  J.R.  and  Best,  A.N.  (1955).  Arch.  Biochem.  Biophys.  54,  318. 

Tower,  D.B.  (1958).  J.  Neurochem.  3,  184. 

Tower,  D.B.,  Peters,  E.L.,  and  Curtis,  W.C.  (1963).  J.  Biol.  Chem.  238,  983. 

Traniello,  S.  and  Vescia,  A.  (1964).  Arch.  Biochem.  Biophys.  105,  465. 

Trim,  A.R.  (1959).  Biochem.  J.  73,  298. 

Troger,  R.  (1959).  Arch.  Mikrobiol.  33,  186. 

Trudinger,  P.A.  and  Cohen,  G.N.  (1956).  Biochem.  J.  62,  488. 

Tsao,  T.-C.  and  Bailey,  K.  (1953).  Biochim.  Biophys.  Acta  11,  102. 

Tschudy,  D.P.,  Colhns,  A.,  Caughey,  W.S.,  and  Kleinspehn,  G.G.  (1960).  Science  131, 
1380. 

Tsen,  C.C.  and  Collier,  H.B.  (1959).  Nature  183,  1327. 

Tsen,  C.C.  and  Collier,  H.B.  (1960).  Can.  J.  Biochem.  Physiol.  38,  981. 

Tsuboi,  K.K.  and  Hudson,  P.B.  (1955  a).  Arch.  Biochem.  Biophys.  55,  191. 

Tsuboi,  K.K.  and  Hudson,  P.B.  (1955  b).  Arch.  Biochem.  Biophys.  55,  206. 

Tsuboi,  K.K.  and  Hudson,  P.S.  (1956).  Arch.  Biochem.  Biophys.  61,  197. 

Tsuboi,  K.K.  and  Hudson,  P.B.  (1957).  J.  Biol.  Chem.  224,  889. 

Tsuboi,  K.K.,  Penefsky,  Z.J.,  and  Hudson,  P.B.  (1957).  Arch.  Biochem.  Biophys. 
68,  54. 

Tsuboi,  K.K.,  Estrada,  J.,  and  Hudson,  P.B.  (1958).  J.  Biol.  Chem,  231,  19. 

Tsukada,  Y.  and  Takagaki,  G.  (1955).  Nature  175,  725. 

Tsukada,  Y.,  Nagata,  Y.,  Hirano,  S.,  and  Takagaki,  G.  (1958).  J.  Biochem.  (Tokyo) 
45,  979. 

Tsukamoto,  H.  (1951).  J.  Pharm.  Soc.  Japan.  71,  471. 

Tsunoda,  T.  and  Shiio,  I.  (1959).  J.  Biochem.  (Tokyo)  46,  1011. 

Tsurumaki,  T.,  Ogiu,  K.,  and  Marui,  E.  (1928).  Folia  Pharmacol.  Japon.  6,  329. 

Tsuyuki,  E.,  Tsuyuki,  H.,  and  Stahmann,  M.A.  (1956).  J.  Biol.  Chem.  Ill,  479. 

Tubbs,  P.K.  (1962).  Biochem.  J.  82,  36. 

Tubbs,  P.K.  and  Greville,  G.D.  (1961).  Biochem.  J.  81,  104. 

Turano,  C,  Giartosio,  A.,  and  Riva,  F.  (1963).  Enzymologia  25,  196. 

Turano,  C,  Giartosio,  A.,  and  Fasella,  P.  (1964).  Arch.  Biocheyn.  Biophys.  104,  524. 

Turba,  F.  and  Kuschinsky,  G.  (1952).  Biochim.  Biophys.  Acta  8,  76. 

Turner,  D.H.  and  Turner,  J.F.  (1961).  Biochem.  J.  79,  143. 

Turner,  J.S.  and  Hanly,  V.  (1947).  Nature  160,  296. 

Turner,  W.A.  and  Hartman,  A.M.  (1925).  J.  Am.  Chem.  Soc.  47,  2044. 

Turpaev,  T.M.  (1955).  Biochemistry  (USSE)  (English  Transl.)  20,  456. 

Turrian,  H.,  Grandjean,  E.,  Blattig,  K.,  and  Turrian,  V.  (1956).  Helv.  Physiol.  Phar- 
macol. Acta  14,  50. 

Tyler,  D.B.  (1942).  Proc.  Soc.  Exptl.  Biol.  Med.  49,  537. 


1062  REFERENCES 

Udaka,  S.  and  Moyed,  H.S.  (1963).  J.  Biol.  Chem.  238,  2797. 

Udenfriend,  S.  and  Cooper,  J.R.  (1952).  J.  Biol.  Chem.  194,  503. 

Udenfriend,  S.,  Creveling,  C.R.,  Ozaki,  M.,  Daly,  J.W.,  and  Witkop,  B.  (1959).  Arch. 
Biochem.  Biophys.  84,  249. 

Ukita,  T.,  Tamura,  T.,  Matsuda,  R.,  and  Kashiwabara,  E.  (1949).  Japan.  J.  Exptl. 
Med.  20,  109. 

Ulbricht,  T.L.V.  and  Gots,  J.S.  (1956).  Nature  178,  913. 

Ulrich,  F.  (1960).  Am.  J.  Physiol.  198,  847. 

Ulrich,  J. A.  and  Fitzpatrick,  T.B.  (1951).  Proc.  Soc.  Exptl.  Biol.  Med.  76,  346. 

Umbreit,  W.W.  (1955  a).  Natl.  Vitamin  Found.  Nutr.  Sym.p.  Ser.  11,  21. 

Umbreit,  W.W.  (1955  b).  In  "Amino  Acid  Metabolism"  (W.D.  McElroy  and  H.B. 
Glass,  eds.),  p.  48.  Johns  Hopkins  Press,  Baltimore,  Maryland. 

Umbreit,  W.W.  and  Waddell,  J.G.  (1949).  Proc.  Soc.  Exptl.  Biol.  Med.  70,  293. 

Umemura,  Y.,  Nishida,  J.,  Akazawa,  T.,  and  Uritani,  I.  (1961).  Arch.  Biochem.  Bio- 
phys. 92,  392. 

Ungar,  G.  and  Psychoyos,  S.  (1963).  Biochim.  Biophys.  Acta  66,  118. 

Unna,  K.  (1940).  Proc.  Soc.  Exptl.  Biol.  Med.  43,  122. 

Urata,  G.  and  Granick,  S.  (1963).  J.  Biol.  Chem.  238,  811. 

Urivetzky,  M.  and  Tsuboi,  K.K.  (1963).  Arch.  Biochem.  Biophys.  103,  1. 

Utzinger,  G.E.,  Strait,  L.A.,  and  Tuck,  L.D.  (1963).  Experientia  19,  324. 

Vagelos,  P.R.  (1960).  J.  Biol.  Chem.  235,  346. 

Vagelos,  P.R.  and  Alberts,  A.W.  (1960).  J.  Biol.  Chem.  235,  2786. 

Vagelos,  P.R.  and  Earl,  J.M.  (1959).  J.  Biol.  Chem.  234,  2272. 

Vallee,  B.L.,  Williams,  R.J.P.,  and  Coleman,  J.E.  (1961).  Nature  190,  633. 

VanArsdel,  P.P.  and  Bray,  R.E.  (1961).  J.  Pharmacol.  Exptl.  Therap.  133,  319. 

Van  Baerle,  R.R.,  Goldstein,  L.,  and  Dearborn,  E.H.  (1957).  Enzymologia  18,  190. 

Vandendriessche,  L.  (1956).  Arch.  Biochem.  Biophys.  65,  347. 

Vander,  A.J.  (1963).  Am.  J.  Physiol.  204,  781. 

Vander,  A. J.,  Malvin,  R.L.,  Wilde,  W.S.,  and  Sullivan,  L.P.  (1958).  Am.  J.  Physiol. 
195,  558. 

van  der  Schoot,  J.B.,  Creveling,  C.R.,  Nagatsu,  T.,  and  Udenfriend,  S.  (1963).  J. 
Pharmacol.  Exptl.  Therap.  141,  74. 

Van  Duuren,  A.J.  (1953).  Bee.  Trav.  Chim.  72,  889. 

van  Eys,  J.  (1956).  Federation  Proc.  15,  374. 

van  Eys,  J.  and  Kaplan,  N.O.  (1957  a).  Biochim.  Biophys.  Acta  23,  574. 

van  Eys,  J.  and  Kaplan,  N.O.  (1957  b).  J.  Biol.  Chem.  228,  305. 

van  Eys,  J.,  Ciotti,  M.M.,  and  Kaplan,  N.O.  (1958).  J.  Biol.  Chem.  231,  571. 

van  Eys,  J.,  Nuenke,  B.J.,  and  Patterson,  M.K.,  Jr.  (1959).  J.  Biol.  Chem.  234, 
2308. 

van  Eys,  J.,  Kretszchmar,  R.,  Tseng,  N.S.,  and  Cunningham,  L.W.,  Jr.  (1962).  Bio- 
chem. Biophys.  Res.  Commun.  8,  243. 

Van  Grembergen,  G.  (1945).  Enzymologia  11,  268. 

Van  Grembergen,  G.  (1949).  Enzymologia  13,  241. 

van  Heyningen,  R.  and  Pirie,  A.  (1953).  Biochem.  J.  53,  436. 

van  Heyningen,  T.  and  Waley,  S.G.  (1963).  Biochem.  J.  86,  92. 

Van  Oorschot,  J.L.P.  and  Hilton,  J.L.  (1963).  Arch.  Biochem.  Biophys.  100,  289. 

Vanov,  S.  (1962).  Arch.  Intern.  Pharmacodyn.  138,  51. 


REFERENCES  1063 

Van  Pilsun,  J.F.  (1953).  J.  Biol.  Chem.  204,  613. 

Van  Thoai,  N.  and  Olomucki,  A.  (1962).  Biochim.  Biophys.  Acta  59,  533. 

van  Vals,  G.H.  and  Emmelot,  P.  (1957).  Z.  Krebsforsch.  62,  63. 

van  Vals,  G.H.,  Bosch,  L.,  and  Emmelot,  P.  (1956).  Brit.  J.  Cancer  10,  792. 

Vardanis,  A.  and  Hochster,  R.M.  (1961).  Can.  J.  Biochem.  Physiol.  39,  1695. 

Vargas,  R.  and  Cafruny,  E.J.  (1959).  Federation  Proc.  18,  455. 

Varner,  J.E.  (1960).  Arch.  Biochem.  Biophys.  90,  7. 

Varner,  J.E.  and  Webster,  G.C.  (1955).  Plant  Physiol.  30,  393. 

Vasington,  F.D.  (1963).  J.  Biol.  Chem.  238,  1841. 

Vasington,  F.D.  and  Murphy,  J.V.  (1962).  J.  Biol.  Chem.  237,  2670. 

Vaslow,  F.  and  Doherty,  D.G.  (1953).  J.  Am.  Chem.  Soc.  75,  928. 

Veitch,  F.P.  and  Brierley,  G.P.  (1962).  Biochim.  Biophys.  Acta  58,  467. 

Velick,  S.F.  (1953).  J.  Biol.  Chem.  203,  563. 

VeUck,  S.F.  (1958).  J.  Biol.  Chem.  233,  1455. 

Velick,  S.F.  and  Ronzoni,  E.  (1948).  J.  Biol.  Chem..  173,  627. 

Velick,  S.F.  and  Vavra,  J.  (1962).  J.  Biol.  Chem.  137,  2109. 

Velluz,  L.  and  Herbain,  M.  (1951).  J.  Biol.  Chem.  190,  241. 

Vennesland,  B.  and  Evans,  E.A.,  Jr.  (1944).  J.  Biol.  Chem.  156,  783. 

Vennesland,  B.,  Evans,  E.A.,  Jr.,  and  Francis,  A.M.  (1946).  J.  Biol.  Chem.  163,  575. 

Venturi,  V.M.  and  Schoepke,  H.G.  (1960).  Proc.  Soc.  Exptl.  Biol.  Med.  103,  455. 

Vercauteren,  R.  (1957).  Enzymologia  18,  97. 

Vernberg,  W.B.  and  Hunter,  W.S.  (1960).  Exptl.  Parasitol.  9,  42. 

Vemon,  L.P.,  Mahler,  H.R.,  and  Sarkar,  N.K.  (1952).  J.  Biol.  Chem.  199,  599. 

Vescia,  A.  and  di  Prisco,  G.  (1962).  J.  Biol.  Chem.  237,  2318. 

Vickery,  H.B.  (1959).  J.  Biol.  Chem.  234,  1363. 

Vickery,  H.B.  and  Palmer,  J.K.  (1956  a).  J.  Biol.  Chem.  218,  225. 

Vickery,  H.B.  and  Palmer,  J.K.  (1956b).  J.  Biol.  Chem.  12^,  79. 

Vickery,  H.B.  and  Palmer,  J.K.  (1957).  J.  Biol.  Chem.  225,  629. 

Villee,  C.A.  and  Loring,  J.M.  (1961).  Biochem.  J.  81,  488. 

Villela,  G.G.  (1963).  Enzymologia  25,  261. 

Villela,  G.G.,  Afifonso,  O.R.,  and  Mitidieri,  E.  (1956).  Ezperientia  12,  477. 

Vincent,  P.O.  (1958).  Australian  J.  Exptl.  Biol.  Med.  Sci.  36,  589. 

Vincent,  P.C.  and  Blackburn,  C.R.B.  (1958).  Australian  J.  Exptl.  Biol.  Med.  Sci.  36, 

471. 
Vishniac,  W.  (1950).  Arch.  Biochem.  26,  167. 
Vishwakarma,  P.  (1957).  Federation  Proc.  16,  132. 
Vishwakarma,  P.  (1962).  Am.  J.  Physiol.  202,  572. 

Vishwakarma,  P.  and  Lotspeich,  W.D.  (1960).  Am.  J.  Physiol.  198,  819. 
Vitale,  J.J.,  Jankelson,  O.M.,  Connors,  P.,  Hasted,  D.M.,  and  Zamcheck,  N.  (1956). 

Am.  J.  Physiol.  187,  427. 
Vogel,  A.I.  (1929).  J.  Chem.  Soc.  p.  1476. 

Vogler,  K.G.,  LePage,  G.A.,  and  Umbreit,  W.W.  (1942).  J.  Gen.  Physiol.  26,  89. 
Volk,  W.A.  (1960).  J.  Biol.  Chem.  235,  1550. 

von  Bruchhausen,  F.  (1964).  Arch.  Exptl.  Pathol.  Pharmakol.  247,  87. 
von  Euler,  H.  (1942).  Ber.  75,  1876. 

von  Euler,  H.  and  Josephson,  K.  (1923).  Z.  Physiol.  Chem.  127,  99. 
von  Euler,  H.  and  Svanberg,  O.  (1920).  Fermentforschung  3,  330. 
Von  Holt,  C,  Blum,  K.U.,  and  Heinrich,  W.-D.  (1955).  Z.  Physiol.  Chem.  302,  68. 


1064  EEFERENCES 

Wachstein,  M.  and  Meisel,  E.  (1954).  Science  119,  100. 

Wachstein,  M.  and  Meisel,  E.  (1957).  J.  Histochem.  Cytochem.  5,  204. 

Wada,  E.  and  Kobashi,  Y.  (1953).  J.  Agr.  Chem.  8oc.  Japan  27,  561. 

Wada,  H.  and  Snell,  E.E.  (1961).  J.  Biol.  Chem.  236,  2089. 

Wada,  H.,  Yoshimatsu,  H.,  Koizumi,  T.,  Inoue,  F.,  Ito,  K.,  Morisue,  T.,  Nasu,  H., 

Ito,  H.,  Sakamoto,  Y.,  and  Ichihara,  K.  (1958).  Proc.  Intern.  Symp.  Enzyme  Chem., 

Tokyo  Kyoto  1957  p.  148.  Maruzen,  Tokyo. 
Wadkins,  C.L.  (1960).  J.  Biol.  Chem.  235,  3300. 

Wadkins,  C.L.  and  Lehninger,  A.L.  (1958).  J.  Biol.  Chem.  233,  1589. 
Wadkins,  C.L.  and  Mills,  R.C.  (1955).  Federation  Proc.  14,  299. 
Wadkins,  C.L.  and  Mills,  R.C.  (1956).  Federation  Proc.  15,  377. 
Wahl,  R.  (1939).  Compt.  Rend.  Soc.  Biol.  131,  231. 

Wakerlin,  G.E.  and  Loevenhart,  A.S.  (1926).  J.  Pharmacol.  Exptl.  Therap.  27,  385. 
Wakil,  S.J.  and  Mahler,  H.R.  (1954).  J.  Biol.  Chem.  207,  125. 
Walaas,  E.  and  Walaas,  O.  (1956).  Acta  Chem.  Scand.  10,  122. 
Wald,  G.  and  Brown,  P.K.  (1951).  Federation  Proc.  10,  266. 
Wald,  G.  and  Brown,  P.K.  (1952).  J.  Gen.  Physiol.  35,  797. 
Wald,  G.,  Brown,  P.K.,  and  Smith,  P.H.  (1955).  J.  Gen.  Physiol.  38,  623. 
Waley,  S.G.  and  van  Heyningen,  R.  (1962).  Biochem.  J.  83,  274. 
Walker,  D.G.  and  Rao,  S.  (1963).  Biochim.  Biophys.  Acta  77,  662. 
Walker,  E.  (1928).  Biochem.  J.  11,  292. 

Walker,  G.C.  and  Nicholas,  D.J.D.  (1961).  Biochem.  J.  78,  lOp. 
Walker,  P.G.,  Woollen,  J.W.,  and  Heyworth,  R.  (1961).  Biochem.  J.  79,  288. 
Walker,  T.K.,  Subramaniam,  V.,  and  Challenger,  F.  (1927).  J.  Chem.  Soc.  p.  3044. 
Wallace,  A.  and  Mueller,  R.T.  (1961).  Plant  Physiol.  36,  118. 
Wallace,  R.A.,  Kurtz,  A.N.,  and  Niemann,  C.  (1963).  Biochemistry  1,  824. 
Wallach,  D.P.  (1960).  Biochem.  Pharmacol.  5,  166. 
Wallach,  D.P.  (1961a).  Biochem.  Pharmacol.  5,  323. 
Wallach,  D.P.  (1961b).  Biochem.  Pharmacol.  8,  328. 
Wallenfels,  K.  and  Miiller-Hill,  B.  (1964).  Biochem.  Z.  339,  352. 
Wallenfels,  K.  and  Sund,  H.  (1957  a).  Biochem.  Z.  329,  17. 
Wallenfels,  K.  and  Sund,  H.  (1957  b).  Biochem.  Z.  329,  48. 
Wallenfels,  K.,  Sund,  H.,  Zarnitz,  M.L.,  Malhotra,  O.P.,  and  Fischer,  J.  (1959).  In 

"Sulfur  in  Proteins"  (R.  Benesch  et  al.,  eds.),  p.  215.  Academic  Press,  New  York. 
Wallgren,  H.  (1960).  Acta  Physiol.  Scand.  49,  216. 
Walsh,  E.  O'F.  and  Walsh,  G.  (1948).  Nature  161,  976. 
Walter,  C.  (1963).  Biochim.  Biophys.  Acta  77,  525. 
Walter,  P.  and  Kaplan,  N.O.  (1963).  J.  Biol.  Chem.  238,  2823. 
Waltman,  J.M.  and  Rittenberg,  S.C.  (1954).  J.  Bacteriol.  68,  585. 
Wang,  T.P.  and  Kaplan,  N.O.  (1954).  J.  Biol.  Chem.  206,  311. 
Warburg,  0.  and  Christian,  W.  (1941).  Naturwissenschaften  39,  589. 
Warburg,  O.  and  Christian,  W.  (1942).  Biochem.  Z.  310,  384. 
Warburg,  O.,  Gawehn,  K.,  and  Geissler,  A.-W.  (1957).  Z.  Naturforsch.  12b,  393. 
Warkany,  J.  and  Hubbard,  D.M.  (1948).  Lancet  I,  829. 
Warner,  C.  (1951).  Australian  J.  Sci.  Res.  4,  554. 
Warner,  R.C.  (1957).  J.  Biol.  Chem.  IV),  711. 
Warnock,  L.G.  and  van  Eys,  J.  (1963).  J.  Bacteriol.  85,  1179. 
Warringa,  M.G.P.J.  and  Giuditta,  A.  (1958).  J.  Biol.  Chem.  230,  HI. 


REFERENCES  1065 

Warringa,  M.G.P.J.,  Smith,  O.H.,  Giuditta,  A.,  and  Singer,  T.P.  (1958).  J .  Biol.  Chem. 

230,  97. 
Washio,  S.  and  Mano,  Y.  (1960).  J.  Biochem.  (Tokyo)  48,  874. 
Watanabe,  K.  and  Yamafuji,  K.  (1961).  Enzymologia  23,  353. 
Waterhouse,  D.F.  (1955).  Australian  J.  Biol.  Sci.  8,  514. 
Watkins,  W.M.  (1959).  Biochem.  J.  71,  261. 
Waton,  N.G.  (1956).  Brit.  J.  Pharmacol.  11,  119. 

Watt,  D.D.  and  Werkman,  C.H.  (1954).  Arch.  Biochem.  Biophys.  50,  64. 
Weaver,  M.E.  (1959).  Anat.  Record  135,  303. 
Webb,  E.G.  and  Morrow,  P.F.W.  (1959).  Biochem.  J.  73,  7. 
Webb,  J.L.  (1950  a).  Brit.  J.  Pharmacol.  5,  87. 
Webb,  J.L.  (1950  b).  Brit.  J.  Pharmacol.  5,  335. 
Webb,  J.L.  and  Hollander,  P.B.  (1959).  Circulation  Res.  7,  131. 
Webb,  J.L.,  Saunders,  P.R.,  and  Thienes,  C.H.  (1949).  Arch.  Biochem.  22,  458. 
Webb,  J.L.A.,  Bhatia,  I.S.,  Corwin,  A.H.,  and  Sharp,  A.G.  (1950).  J.  Am.  Chem.  Soc. 

72,  91. 
Weber,  H.H.  and  Portzehl,  H.  (1954).  Progr.  Biophys.  Biophys.  Chem.  4,  60. 
Weed,  R.I.,  Eber,  J.,  and  Rothstein,  A.  (1962).  J.  Oen.  Physiol.  45,  395. 
Weeks,  J.R.  and  Chenoweth,  M.B.  (1950).  J.  Pharmacol.  Exptl.  Therap.  98,  35. 
Weeks,  J.R.,  Chenoweth,  M.B.,  and  Shideman,  F.E.  (1950).  J.  Pharmacol.  Exptl. 

Therap.  98,  224. 
Weill,  C.E.  and  Caldwell,  M.L.  (1945  a).  J.  Am.  Chem.  Soc.  67,  212. 
Weill,  C.E.  and  Caldwell,  M.L.  (1945  b).  J.  Am.  Chem.  Soc.  67,  214. 
Weil-Malherbe,  H.  (1937).  Biochem.  J.  31,  299. 
Weil-Malherbe,  H.  and  Bone,  A.D.  (1951).  Biochem.  J.  49,  339. 
Weinbach,  E.G.  (1953).  Arch.  Biochem.  Biophys.  42,  231. 
Weinbach,  E.G.  (1961).  J.  Biol.  Chem.  236,  1526. 

Weiner,  I.M.  and  Miiller,  C.H.  (1955).  J.  Pharmacol.  Exptl.  Therap.  113,  241. 
Weiner,  I.M.,  Burnett,  A.E.,  and  Rennick,  B.R.  (1956).  J.  Pharmacol.  Exptl.  Therap. 

118,  470. 
Weiner,  I.M.,  Garlid,  K.,  Sapir,  D.,  and  Mudge,  G.H.  (1959).  J.  Pharmacol.  Exptl. 

Therap.  127,  325. 
Weiner,  I.M.,  Levy,  R.I.,  and  Mudge,  G.H.  (1962).  J.  Pharmacol.  Exptl.  Therap.  138, 

96. 
Weinhouse,  S.,  Millington,  R.H.,  and  Friedman,  B.  (1949).  J.  Biol.  Chem.  181,  489. 
Weinhouse,  S.,  Millington,  R.H.,  and  Wenner,  C.E.  (1951).  Cancer  Res.  11,  845. 
Weiss,  B.  (1951).  J.  Biol.  Chem.  193,  509. 

Weitzel,  G.  and  Buddecke,  E.  (1959).  Z.  Physiol.  Chem.  317,  150. 
Weitzel,  G.  and  Schaeg,  W.  (1959).  Z.  Physiol.  Chem.  316,  250. 
Welch,  A.D.  (1945).  Physiol.  Rev.  25,  687. 
Welch,  K.  (1962).  Am.  J.  Physiol.  202,  757. 
Wells,  H.G.  (1916).  J.  Biol.  Chem.  26,  319. 
Wells,  I.e.  (1954).  J.  Biol.  Chem.  207,  575. 
Welt,  L.G.,  Goodyer,  A.V.N.,  Darragh,  J.H.,  Abele,  W.A.,  and  Meroney,  W.H.  (1953). 

J.  Appl.  Physiol.  6,  134. 
Wenner,  C.E.  and  Cereijo-Santalo,  R.  (1962).  Arch.  Biochem.  Biophys.  98,  67. 
Wenner,  C.E.  and  Paigen,  K.  (1961).  Arch.  Biochem.  Biophys.  93,  646. 
Wenzel,  D.G.  and  Siegel,  LA.  (1962).  J.  Pharm.  Sci.  51,  1074. 


1066  REFERENCES 

Werkheiser,  W.C.  (1959).  Proc.  Am.  Assoc.  Cancer  Res.  3,  72. 

Werkheiser,  W.C.  (1960).  Proc.  Am.  Assoc.  Cancer  Res.  3,  161. 

Werkheiser,  W.C.  (1961).  J.  Biol.  Chem.  236,  888. 

Werkheiser,  W.C.  (1962).  J.  Pharmacol.  Exptl.  Therap.  137,  167. 

Werle,  E.  and  Koch,  W.  (1949).  Biochem.  Z.  319,  305. 

Wessels,  J.G.H.  (1959).  Acta  Botan.  Neerl.  8,  497. 

Wessels,  J.S.C.  (1958).  Biochim.  Biophys.  Acta  29,  113. 

Wessels,  J.S.C.  (1959).  Biochim.  Biophys.  Acta  35,  53. 

Wessels,  J.S.C.  and  Baltscheffsky,  H.  (1960).  Acta  Chem.  Scand.  14,  233. 

Wesson,  L.G.  (1962).  J.  Lab.  Clin.  Med.  59,  630. 

Westerfeld,  W.W.  and  Richert,  D.A.  (1952).  J.  Biol.  Chem.  199,  393. 

Westerfeld,  W.W.,  Richert,  D.A.,  and  Higgins,  E.S.  (1959).  J.  Biol.  Chem.  234,  1897. 

Westermann,  E.,  Balzer,  H.,  and  Knell,  J.  (1958).  Arch.  Exptl.  Pathol.  Pharmakol. 

234,  194. 
Westhead,  E.W.  and  Boyer,  P.D.  (1961).  Biochim.  Biophys.  Acta  54,  145. 
Westheimer,  F.H.  and  Shookhoff,  M.W.  (1939).  J.  Am.  Chem.  Soc.  61,  555. 
Westlake,  D.W.S.,  Shug,  A.L.,  and  Wilson,  P.W.  (1961).  Can.  J.  Microbiol.  7,  515. 
White,  E.G.  and  Vernon,  L.P.  (1958).  J.  Biol.  Chem.  233,  217. 
White,  H.L.,  Rolf,  D.,  Bisno,  A.L.,  Kasser,  I.S.,  and  Tosteson,  D.C.  (1961).  Am.  J. 

Physiol.  200,  885. 
Whitehouse,  M.W.,  Staple,  E.,  and  Gurin,  S.  (1959).  J.  Biol.  Che^n.  234,  276. 
Whiteley,  H.R.,  Osborn,  M.J.,  and  Huennekens,  F.M.  (1959).  J.  Biol.  Chem.  234,  1538. 
Whitmore,  F.C.  and  Woodward,  G.E.  (1941).  In  "Organic  Syntheses"  (H.  Gilraan 

and  A.H.  Blatt,  eds.).  Collective  Vol.  1,  p.  159.  Wiley,  New  York. 
Whittingham,  C.P.  (1956).  J.  Exptl.  Botany  7,  273. 
Wick,  A.N.,  Drury,  D.R.,  and  Morita,  T.N.  (1955).  Proc.  Soc.  Exptl.  Biol.  Med.  89, 

579. 
Wick,  A.N.,  Wolfe,  J.B.,  and  Nakada,  H.I.  (1956).  Proc.  Soc.  Exptl.  Biol.  Med.  93, 

150. 
Wick,  A.N.,  Drury,  D.R.,  Nakada,  H.I.,  and  Wolfe,  J.B.  (1957).  J.  Biol.  Chem.  224, 

963. 
Wick,  A.N.,  Serif,  G.S.,  Stewart,  C.J.,  Nakada,  H.I.,  Larson,  E.R.,  and  Drury,  D.R. 

(1959).  Diabetes  8,  112. 
Wickson-Ginzburg,  M.  and  Solomon,  A.K.  (1963).  J.  Gen.  Physiol.  46,  1303. 
Wieland,  O.  and  Weiss,  L.  (1963).  Biochem.  Biophys.  Res.  Commun.  13,  26. 
Wien,  R.  (1939).  Quart.  J.  Pharm.  Pharmacol.  12,  212. 
Wiesmeyer,  H.  and  Colin,  M.  (1957).  Federation  Proc.  16,  270. 
Wiesmeyer,  H.  and  Cohn,  M.  (1960).  Biochim.  Biophys.  Acta  39,  427. 
Wiethoff,  E.G.,  Leppla,  W.,  and  Holt,  C.V.  (1957).  Biochem.  Z.  328,  576. 
Wigler,  P.W.  and  Alberty,  R.A.  (1960).  J.  Am.  Chem.  Soc.  82,  5482. 
Wilbrandt,  W.  (1941).  Arch.  Ges.  Physiol.  244,  637. 
Wilbrandt,  W.  and  Rosenberg,  T.  (1961).  Pharmacol.  Rev.  13,  109. 
Wilde,  C.E.  and  Kekwick,  R.G.O.  (1964).  Biochem.  J.  91,  297. 
Willard,  H.H.  and  Young,  P.  (1930).  J.  Am.  Chem.  Soc.  52,  132. 
WilHams,  A.D.,  Slater,  G.G.,  and  Winzler,  R.J.  (1955).  Cancer  Res.  15,  532. 
Williams,  A.K.  and  Eagon,  R.G.  (1959).  J.  Bacteriol.  77,  167. 
Williams,  C.H.,  Jr.  and  Kamin,  H.  (1962).  J.  Biol.  Chem.  237,  587. 
Williams,  J.N.,  Jr.  (1952).  J.  Biol.  Chem.  195,  629. 


REFERENCES  1067 

Williams,  R.J.P.  (1956).  Chem.  Rev.  56,  299. 

Willing,  F.,  Neuhoff,  V.,  and  Herken,  H.  (1964).  Arch.  Exptl.  Pathol.  Pharmakol.  247, 

254. 
Wills,  E.D.  (1959).  Biochem.  Pharmacol.  2,  276. 
Wills,  E.D.  (1960).  Biochim.  Biophys.  Acta  40,  481. 
Wilson,  A.N.  and  Harris,  S.A.  (1949).  J.  Am.  Chem.  Soc.  71,  2231. 
Wilson,  G.B.,  Morrison,  J.H.,  and  Knobloch,  N.  (1959).  J.  Biophys.  Biochem.  Cytol. 

5,  411. 
Wilson,  J.E.,  Clark,  W.C,  and  Hulpieu,  H.R.  (1962).  J.  Pharmacol.  Exptl.  Therap. 

137,  179. 
Wilson,  P.W.,  Burris,  R.H.,  and  Lind,  C.J.  (1942).  Proc.  Natl.  Acad.  Sci.  U.  S.  28. 

243. 
Wilson,  T.G.G.  and  Roberts,  E.R.  (1954).  Biochim.  Biophys.  Acta  15,  568. 
Wilson,  T.H.  and  Landau,  B.R.  (1960).  Am.  J.  Physiol.  198,  99. 
Wilson,  T.H.,  Lin,  E.C.C.,  Landau,  B.R.,  and  Jorgenson,  C.R.  (1960).  Federation  Proc 

19,  870. 
Winer,  A.D.  (1960).  Biochem.  J.  76,  5p. 

Winer,  A.D.,  Schwert,  G.W.,  and  Millar,  D.B.S.  (1959).  J.  Biol.  Chem.  234,  1149. 
Winestock,  C.H.,  Aogaichi,  T.,  and  Plaut,  G.W.E.  (1963).  J.  Biol.  Chem.  238,  2866. 
Wingo,  W.J.  and  Awapara,  J.  (1950).  J.  Biol.  Chem.  187,  267. 
Winzor,  D.J.  and  Creeth,  J.M.  (1962).  Biochem.  J.  83,  559. 
Wiseman,  G.  (1954).  J.  Physiol.  {London)  124,  414. 
Wiseman,  M.H.  and  Sourkes,  T.L.  (1961).  Biochem.  J.  78,  123. 
Wiskich,  J.T.,  Morton,  R.K.,  and  Robertson,  R.N.  (1960).  Australian  J.  Biol.  Sci. 

13,   109. 
Wislicenus,  J.  (1874).  Ber.  7,  683. 
Witter,  A.  (1960).  Acta  Chem.  Scand.  14,  1717. 
Witter,  A.  and  Tuppy,  H.  (1960).  Biochim.  Biophys.  Acta  45,  429. 
Witter,  R.F.,  Newcomb,  E.H.,  and  Stotz,  E.  (1950).  J.  Biol.  Chem.  185,  537. 
Wockel,  W.,  Stegner,  H.-E.,  and  Janisch,  W.  (1961).  Arch.  Pathol.  Anaf.  Physiol.  334, 

503. 
Wohl,  A.  and  Glimm,  E.  (1910).  Biochem.  Z.  27,  349. 
Wojtczak,  L.,  Wlodawer,  P.,  and  Zborowski,  J.  (1963).  Biochim.  Biophys.  Acta  70, 

290. 
Wolbach,  R.A.  and  Riggs,  A.  (1955).  Federation  Proc.  14,  306. 
Wold,  F.  and  Ballou,  C.E.  (1957).  J.  Biol.  Chem.  117,  313. 
Wolf,  P. A.  (1950).  Food  Technol.  4,  294. 

Wolf,  P.A.  and  Westveer,  W.M.  (1950).  Arch.  Biochem.  28,  201. 
Wolfe,  J.B.  (1955).  Arch.  Biochem.  Biophys.  57,  414. 
Wolfe,  J.B.  and  Nakada,  H.I.  (1956).  Arch.  Biochem.  Biophys.  64,  489. 
Wolfe,  J.B.  and  Rittenberg,  S.C.  (1954).  J.  Biol.  Chem.  209,  885. 
Wolfe,  J.B.,  Ivler,  D.,  and  Rittenberg,  S.C.  (1954  a).  J.  Biol.  Chem.  209,  867. 
Wolfe,  J.B.,  Ivler,  D.,  and  Rittenberg,  S.C.  (1954  b).  J.  Biol.  Chem.  209,  875. 
Wolfe,  J.B.,  Ivler,  D.,  and  Rittenberg,  S.C.  (1955).  J.  Bacteriol.  69,  240. 
Wolfe,  R.G.  and  Neilands,  J.B.  (1956).  J.  Biol.  Chem.  ll'i,  61. 
Wolfe,  S.J.  (1957).  J.  Biol.  Chem.  229,  801. 

Wolff,  J.B.  and  Resnik,  R.A.  (1963).  Biochim.  Biophys.  Acta  73,  588. 
Wood,  H.G.  and  Stjernholm,  R.  (1961).  Proc.  Natl.  Acad.  Sci.  U.  S.  47,  289. 


1068  EEFERENCES 

Wood,  H.G.  and  Werkman,  C.H.  (1940).  Biochem.  J.  34,  7. 

Wood,  J.D.  (1959).  Can.  J.  Biochem.  Physiol.  37,  937. 

Wood,  R.C.  and  Hitchings,  G.H.  (1959  a).  J.  Biol.  Chem.  234,  2377. 

Wood,  R.C.  and  Hitchings,  G.H.  (1959  b).  J.  Biol.  Chem.  234,  2381. 

Woodruff,  L.L.  and  Bunzel,  H.H.  (1910).  Am.  J.  Physiol.  25,  190. 

Woods,  D.D.  (1940).  Brit.  J.  Exptl.  Pathol.  21,  74. 

Woods,  L.A.,  Shideman,  F.E.,  Seevers,  M.H.,  Weeks,  J.R.,  and  Kruse,  W.T.  (1950). 

J.  Pharmacol.  Exptl.  Therap.  99,  84. 
Woodward,  G.E.  and  Hudson,  M.T.  (1954).  Cancer  Pes.  14,  599. 
Woody,  B.R.  and  Lindstrom,  E.S.  (1955).  J.  Bacteriol.  69,  353. 
Woolley,  D.W.  (1944  a).  Proc.  Soc.  Exptl.  Biol.  Med.  55,  179. 
Woolley,  D.W.  (1944  b).  J.  Biol.  Chem.  154,  31. 
Woolley,  D.W.  (1945  a).  Proc.  Soc.  Exptl.  Biol.  Med.  60,  225. 
Woolley,  D.W.  (1945  b).  J.  Biol.  Chem.  157,  455. 
Woolley,  D.W.  (1950  a).  J.  Biol.  Chem.  191,  43. 
Woolley,  D.W.  (1950  b).  Ann.  N.  Y.  Acad.  Sci.  52,  1235. 
Woolley,  D.W.  (1951).  J.  Biol.  Chem.  191,  43. 

Woolley,  D.W.  (1952).  "A  Study  of  Antimetabolites."  Wiley,  New  York. 
Woolley,  D.W.  and  Merrifield,  R.B.  (1952).  Federation  Proc.  11,  458. 
Woolley,  D.W.  and  Pringle,  A.  (1951).  Federation  Proc.  10,  272. 
Woolley,  D.W.  and  White,  A.G.C.  (1943  a).  Proc.  Soc.  Exptl.  Biol.  Med.  52,  106. 
Woolley,  D.W.  and  White,  A.G.C.  (1943  b).  J.  Biol.  Chem.  149,  285. 
Woolley,  D.W^  and  White,  A.G.C.  (1943  c).  J.  Exptl.  Med.  78,  489. 
Woolley,  D.W.,  Strong,  F.M.,  Madden,  T.J.,  and  Elvehjem,  C.A.  (1938).  J.  Biol.  Chem. 

124,  715. 
Work,  T.S.  and  Work,  E.  (1948).  "The  Basis  of  Chemotherapy."  Wiley  (Interscience), 

New  York. 
Wormser,  E.H.  and  Pardee,  A.B.  (1958).  Arch.  Biochetn.  Biophys.  78,  416. 
Wortman,  B.  (1962).  Arch.  Biochem.  Biophys.  97,  70. 
Wosilait,  W.D.  (1960).  J.  Biol.  Chem.  235,  1196. 
Wosilait,  W.D.  and  Nason,  A.  (1954).  J.  Biol.  Chem.  206,  255. 
Wright,  C.I.  and  Sabine,  J.C.  (1944).  J.  Biol.  Chem.  155,  315. 
Wright,  C.I.  and  Sabine,  J.C.  (1948).  J.  Pharmacol.  Exptl.  Therap.  93,  230. 
Wright,  N.C.  (1926).  Biochem.  J.  20,  524. 

Wriston,  J.C,  Jr.,  Lack,  L.,  and  Shemin,  D.  (1955).  J.  Biol.  Chem.  215,  603. 
Wu,  H.L.C.  and  Mason,  M.  (1964).  J.  Biol.  Chem.  239,  1492. 
Wu,  L.-C.  and  Scheffer,  R.P.  (1960).  Plant  Physiol.  35,  708. 
Wyatt,  H.V.  (1964).  J.  Theoret.  Biol.  6,  441. 
Wylie,  D.W.,  Archer,  S.,  and  Arnold,  A.  (1960).  J.  Pharmacol.  Exptl.  Therap.  130, 

239. 
Wyngaarden,  J.B.  (1957).  J.  Biol.  Chem.  224,  453. 
Wyngaarden,  J.B.  and  Ashton,  D.M.  (1959).  Federation  Proc.  18,  174. 
Wyngaarden,  J.B.  and  Greenland,  R.A.  (1963).  J.  Biol.  Chem.  238,  1054. 
Wynn,  J.  and  Gibbs,  R.  (1963).  J.  Biol.  Chem.  238,  3490. 
Wyss,  O.  and  Strandskov,  F.B.  (1945).  Arch.  Biochem.  6,  261. 

Yagi,  K.  and  Nagatsu,  I.  (1960).  J.  Biochem.  (Tokyo)  47,  630. 
Yagi,  K.  and  Natsume,  K.  (1964).  J.  Biochem.  (Tokyo)  55,  529. 


REFERENCES  1069 

Yagi,  K.  and  Ozawa,  T.  (1959).  Nature  184,  1227. 

Yagi,  K.,  Okuda,  J.,  Ozawa,  T.,  and  Okada,  K.  (1957).  Biochem.  Z.  328,  492. 

Yagi,  K.,  Ozawa,  T.,  and  Okada,  K.  (1959).  Biochim.  Biophys.  Acta  35,  102. 

Yagi,  K.,  Ozawa,  T.,  and  Harada,  M.  (1960).  Nature  188,  745. 

Yamada,  E.W.  and  Jakoby,  W.B.  (1959).  J.  Biol.  Chem.  234,  941. 

Yamada,  K.  (1959).  J.  Biochem.  {Tokyo)  46,  361. 

Yamada,  T.  and  Yanagita,  T.  (1957).  J.  Biochem.  (Tokyo)  44,  661. 

Yamamura,  Y.,  Kusunose,  M.,  and  Kusunose,  E.  (1952).  J.  Biochem.  {Tokyo)  39,  227. 

Yamamura,  Y.,  Kusunose,  M.,  Nagai,  S.,  and  Kusunose,  E.  (1954).  J.  Biochem.  {Tokyo) 

41,  513. 
Yamane,  T.  and  Davidson,  N.  (1961).  J.  Am.  Chem.  Soc.  83,  2599. 
Yamashita,  J.  (1960).  J.  Biochem.  {Tokyo)  48,  525. 
Yamauchi,  M.  (1956).  Nippon  Saikingaku  Zasshi  11,  221. 
Yanari,  S.  and  Mitz,  M.A.  (1957).  J.  Am.  Chem.  Soc.  79,  1154. 
Yang,  W.C.T.  (1957).  Science  125,  1087. 

Yeas,  M.F.  (1950).  Ph.D.  thesis,  CaHf.  Inst.  Technol.,  Pasadena,  California. 
Yeas,  M.F.  (1954).  J.  Exptl.  Biol.  31,  208. 

Yee,  R.B.,  Pan,  S.F.,  and  Gezon,  H.M.  (1958).  J.  Bacteriol.  75,  51. 
Yonetani,  T.  and  Theorell,  H.  (1962).  Arch.  Biochem.  Biophys.  99,  433. 
Yonetani,  T.  and  Theorell,  H.  (1964).  Arch.  Biochem.  Biophys.  106,  243. 
Yoneya,  T.  and  Adams,  E.  (1961).  J.  Biol.  Chem.  236,  3272. 
Yoshiba,  A.,  Nakao,  K.,  Minakami,  S.,  and  Yoneyama,  Y.  (1958).  J.  Biochem.  {Tokyo) 

45,  913. 
Yoshida,  F.  and  Nagasawa,  M.  (1958).  Proc.  Intern.  Symp.  Enzyme  Chem.,  Tokyo 

Kyoto  1957  p.  504.  Maruzen,  Tokyo. 
Yoshida,  H.  and  Quastel,  J.H.  (1962).  Biochim.  Biophys.  Acta  57,  67. 
Young,  R.H.  and  Shannon,  L.M.  (1959).  Plant  Physiol.  34,  149. 
Yushok,  W.D.  (1964).  Cancer  Res.  24,  187. 
Yushok,  W.D.  and  Batt,  W.G.  (1957).  Ahstr.  Am.  Chem.  Soc,  132nd  Meeting,  1957 

pp.  23C-24C. 

Zahl,  P.A.  and  Albaum,  H.G.  (1955).  Proc.  Soc.  Exptl.  Biol.  Med.  88,  263. 

Zakrzewski,  S.F.  (1963).  J.  Biol.  Chem.  238,  1485. 

Zakrzewski,  S.F.  and  Nichol,  C.A.  (1958).  Biochim.  Biophys.  Acta  27,  425. 

ZambonelH,  C.  (1958  a).  Ann.  Microbiol.  Enzimol.  8,  27. 

ZamboneUi,  C.  (1958  b).  Ann.  Microbiol.  Enzimol.  8,  98. 

Zannoni,  V.G.  and  La  Du,  B.N.  (1959).  J.  Biol.  Chem.  234,  2925. 

Zatman,  L.J.,  Kaplan,  N.O.,  and  Colowick,  S.P.  (1953).  J.  Biol.  Chem.  200,  197. 

Zatman,  L.J.,  Kaplan,  N.O.,  Colowick,  S.P.,  and  Ciotti,  M.M.  (1954  a).  J.  Biol.  Chem. 

209,  453. 
Zatman,  L.J.,  Kaplan,  N.O.,  Colowick,  S.P.,  and  Ciotti,  M.M.  (1954  b).  J.  Biol.  Chem. 

209,  467. 
Zelitch,  I.  (1955).  J.  Biol.  Chem.  216,  553. 
Zelitch,  I.  (1957).  J.  Biol.  Chem.  224,  251. 
Zelitch,  I.  (1958).  Federation  Proc.  17,  341, 
Zelitch,  I.  (1959).  J.  Biol.  Chem.  234,  3077. 
Zeller,  E.A.  (1938).  Helv.  Chim.  Acta  21,  1645. 
Zeller,  E.A.  (1940).  Helv.  Chim.  Acta  23,  1418. 


1070  REFERENCES 

Zeller,  E.A.  (1941).  Helv.  Chim.  Acta  24,  539. 

Zeller,  E.A.  and  Maritz,  A.  (1944).  Helv.  Chim.  Acta  27,  1888. 

Zeller,  E.A.  and  Maritz,  A.  (1945).  Helv.  Chim.  Acta  28,  365. 

Zellweger,  H.  and  Wehrli,  S.  (1951).  Helv.  Paediat.  Acta  6,  397. 

Ziegler,  D.M.  (1961).  In  "Biological  Structure  and  Function"  (T.W.  Goodwin  and  0. 

Lindberg,  eds.),  Vol.  2,  p.  253.  Academic  Press,  New  York. 
ZiflF,  M.  (1944).  J.  Biol.  Chem.  153,  25. 
Zimmerman,  A.M.,  Landau,  J.V.,  and  Marsland,  D.  (1957).  J.  Cellular  Comp.  Physiol. 

49,  395. 
Zimmerman,  S.B.  and  Kornberg,  A.  (1961).  J.  Biol.  Chem.  236,  1480. 
Zipkin,  I.,  and  McClure,  F.J.  (1957).  J.  Am.  Dental  Assoc.  55,  15. 
Zipkin,  I.  and  McClure,  F.J.  (1958).  Proc.  Soc.  Exptl.  Biol.  Med.  97,  318. 
Zittle,  C.A.  (1945).  J.  Biol.  Chem.  160,  527. 
ZoUner,  N.  and  Fellig,  J.  (1952).  Naturwissenschaften  39,  523. 
Zollner,  N.  and  Fellig,  J.  (1953).  Am.  J.  Physiol.  173,  223. 
ZoUner,  N.  and  Rothemund,  E.  (1954).  Z.  Physiol.  Chem.  298,  97. 
Zygmunt,  W.A.  (1963).  J.  Bacteriol.  85,  1217. 


I 


AUTHOR  INDEX 


Numbers  in  italics  refer  to  pages  on  which  the  complete  references  are  listed. 


Abadom,  P.  N.,  267,  9S7 

Abe,  S.  982,  987 

Abele,  W.  A.,  925,  1065 

Abeles,  R.  H.,  590,  1030 

Abood,  L.  G.,  847,  873,  878,  987,  990 

Abraham,   E.   P.,  599,   987 

Abraham,  S.,  146,  147,  273,  987,  997 

Abul-Fadl,  M.  A.  M.,  440,  987 

Ackermann,  W.  W.,  30,  42,  127,  175,  193, 

309,  494,  987 
Acland,  J.  D.,  910,  987 
Actis,  A.  S.,  687,  712,  716,  775,  783,  810, 

853,   1058 
Ada,  G.  L.,  851,  987 
Adams,  E.,  269,  334,  355,  549,  775,  783, 

857,  987,  1013,  1055,  1069 
Adams,  R.,  618,  1047 
Adell,  B.,  3,  8,  987 
Adinolfi,  A.,  615,  616,  1049 
Adkins,  H.,  619,  987 
Adler,  E.,  61,  668,  987,  998 
Alfonso,  D.  R.,  289,  1063 
Agarwala,  S.  C,  692,  831,  987 
Agosin,  M.,  18,  25,  28,  173,  203,  383,  406, 

407,  411,  839,  843,  854,  882,  987,  1035 
Agranoflf,   B.  W.,  887,  987 
Ahmed,  K.,  266,  987 
Aikawa,  J.  K.,  959,  987 
Aikawa,  T.,  678,  727,  1022 
Airan,  J.  W.,   15,  987 
Aisenberg,  A.  C,   126,   128,  987 
Ajl,  S.  J.,  26,  79,  177,  187,  242,  987,  1043 
Akabori,  S.,  416,  417,  1025 
Akamatsu,  S.,  428,  987 
Akatsu,  S.,  983,  1042 
Akawie,  R.  I.,  310,  1016 
Akazawa,  T.,  439,  1062 


Akiba,  T.,  983,  984,  987 

Alanis,  J.,   724,  987 

Albaum,  H.  G.,  169,  196,  228,  585,  987, 

1069 
Albers,  R.  W.,  601,  856,  987 
Albert,  A.,  259,  261,  972,  973,  988 
Alberts,   A.   W.,   847,   1067 
Alberty,  R.  A.,  274,  277,  278,  1066 
Albrecht,   A.   M.,   590,   988 
Albrecht,  M.,  151,  1036 
Albright,  E.  C.,  602,  1030 
Alburn,  H.  E.,  459,  988 
Aleem,  M.  I.  H.,  547,  551,  988 
Alex,  T.,  910,  913,  1050 
Alivisatos,  S.  G.  A.,  485,  487,  492,  503, 

988 
Allen,  E.  H.,  307,  1052 
Allen,  F.  W.,  712,  1003 
Allen,  M.  B.,  892,  989 
Allen,  R.  P.,  698,  699,  700,  1012,  1045 
AUfrey,  V.  G.,  32,  189,  393,  622,  623, 1034 
Allison,  A.  C.,  756,  977,  978,  979,  980, 

981,  988 
Almkvist,  J.,  950,  988 
Alpert,  N.  R.,  446,  1039 
AltareUi,  V.  R.,  960,  997 
Altekar,  W.  W.,  274,  988,  1047 
Altszuler,  N.,  401,  988 
Alvarado,  F.,  414,  1056 
Alvarez-O'Bourke,  F.,  505,  1047 
Amaha,  M.,  546,  988 
Ames,  S.  R.,  593,  659,  662,  988 
Anagnostopoulos,   C.,   63,   440,   998 
Anastasi,   A.,   671,   1027 
Anderson,    B.   M,,   497,   988 
Anderson,   D.   G.,   832,   886,  988,   991 
Anderson,  D.  O.,  464,  1021 
Anderson,  E.  P.,  473,  480,  988,  996 


1071 


1072 


AUTHOR    INDEX 


Anderson,   H.  L.,  217,  1006 
Anderson,  J.  A.,  38,  1003 
Anderson,  K.  S.,  625,  992 
Anderson,  M.  V.,  Jr.,  625,  992 
Andrews,  C.   H.,  977,  978,  979,  988 
Andrews,  G.  S.,  55,  179,  228,  988 
Anfinsen,  C.  B.,  708,  826,  840,  863,  1043 
Angielski,  S.,  206,  988 
Anichkov,  S.  V.,  212,  988 
Annau,   E.,   138,   179,   183,  988 
Annison,   E.   F.,   77,   1052 
Anson,  M.  L.,  661,  670,  671,  672,  680, 
681,  690,  697,  754,  979,  980,  988,  1038 
Anthony,  D.  D.,  817,  820,  988 
Anthony,  W.  L.,  495,  992 
Antonin,  S.,  657,  1003 
Antopoll,   W.,   58,   1009 
Aogaichi,  T.,  539,  543,  1067 
Aposhian,  H.  V.,  539,  956,  988 
App,  A.  A.,  831,  988 
Appleton,  J.  M.,  613,  1044 
Applewhite,  T.  H.,  374,  988 
Appleyard,  G.,  836,  988 
Ap  Rees,  T.,  51,  53,  169,  172,  1016 
Araki,  K.,   126,  988 
Aravena,  L.,  383,  411,  843,  854,  987 
Arbuthnott,  J.  P.,  900,  901,  988 
Archer,  S.,  611,  612,  1068 
Archibald,  R.  M.,  550,  988 
Ardao,  M.  I.,  92,  722,  882,  991 
Arkin,  A.,  701,  727,  728,  989 
Armstrong,  J.  McD.,  787,  989 
Arndt,  F.,  620,  989 
Arnold,  A.,  449,  611,  612,  1008,  1068 
Arnon,  D.  I.,  892,  989 
Arrigoni-MarteUi,  E.,  236,  1028 
Asahi,  T.,  543,  554,  989 
Asano,  A.,  552,  864,  989 
Asano,  N.,  195,  1015 
Ashton,  D.  M.,  474,  480,  1068 
Ashton,  G.  C,  422,  1019 
Ashwell,  G.,  414,  465,  1004 
Asnis,  R.   E.,  26,  989 
Astrup,  T.,  464,  465,  989 
Atkinson,  D.  E.,  293,  512,  851,  892,  989, 

1030,  1034 
Atkinson,  M.  R.,  471,  481,  510,  989 
Aubel,  E.,  52,  60,  74,  81,  989 


Audereau,  A.,  62,  63,  81,  82,  1033 

Auditore,  G.  V.,  936,  989 

Augustinsson,  K.   B.,  817,  989 

Austen,  K.  F.,  374,  989 

Austin,  J.  H.,  956,  1060 

Avery,  G.  S.,  Jr.,  53,  60,  81,  596,  992 

Avi-Dor,   Y.,   20,   29,   61,   80,   844,   848, 

864,  872,  878,  989,  998,  1013 
Avigad,   G.,   421,   1018 
Avron,  M.,  74,  78,  79,  81,  82,  91,  557,  851, 

891,  892,  989,  1014,  1022 
Awapara,  J.,  328,  1067 
Axelrod,  A.  E.,  35,  36,  589,  989,  1059 
Axelrod,  B.,  238,  411,  440,  852,  854,  887, 

989,  1015,   1026,    7046 
Axelrod,  J.,  590,  5*^/2,  597,  611,  843,  989, 

996 
Ayengar,  P.,  336,  989 
Ayers,  J.,  639,  1027 
Azarkh,  R.  M.,  360,  995 
Azoulay,  E.,  831,  832,  989,  1018 
Azzone,  G.  F.,  17,  18,  33,  121,  122,  865, 

989,  1030 

B 

Bach,  M.  K.,   169,  881,  989 

Bach,  S.  J.,  335,  989 

Bach,  Z.  M.,  611,  938,  989 

Bachhawat,   B.   K.,  816,  1044 

Bachmann,  H.,  949,  989 

Bachrach,    H.  L.,    176,    181,    194,    401, 

996,  1045 
Bacila,  M.,  550,  1001 
Bacr,  J.  E.,  917,  993 
Baer,  H.,  783,  854,  989 
Baernstein,  H.  D.,  696,  697,  989 
Bassler,  K.  H.,  601,  1030 
Bagley,  E.  H.,  62,  85,  87,  1013 
Bagot,  A.  E.,  660,  1014 
Bahn,   R.   C,   926,  989 
Bailey,  J.  H.,  973,  983,  1027 
Bailey,  K.,  684,  691,  692,  704,  705,  721, 

722,  723,  865,  876,  938,  939,  989,  1061 
Bain,  J.  A.,  567,  568,  990 
Baker,  E.  E.,  196,  1040 
Baker,  G.  D.,  495,  992 
Baker,  N.,  695,  990 
Baker,  R.  S.,  852,  989 


AUTHOR    INDEX 


1073 


Baker,   Z.,   500,   990 

Baldwin,  D.  M.,  976,  979,  980,  1028 

Baldwin,  E.,  54,  174,  990 

Balinsky,  D.,  349,  593,  604,  607,  990 

Balis,  M.  E.,  585,  990 

Ball,  H.  A.,  388,  990 

Ballou,  C.  E.,  408,  409,  1008,  1067 

Balls,  A.  K.i  657,  667,  668,  990,  1023 

Baltscheffsky,  H.,  445,  548,  557,  864, 
990,  1066 

Baltscheffsky,  M.,  163,  168,  557,  892. 
990,  1055 

Balzer,  H.,  314,  315,  1066 

Bamberger,  J.  W.,   198,  1038 

Bandelin,  F.  J.,  633,   990 

Bandurski,  R.  S.,  71,  172,  189,  543, 
554,  852,  989,  990,  995 

Banga,  I.,  75,  175,  181,  187,  990 

Banks,  J.,  497,  1036 

Bagtist,  J.  N.,  437,  990 

Baralt-Perez,  J.,  978,  980,  1044 

Barany,  K.,  707,  939,  940,  990 

Bardny,  M.,  663,  664,  704,  707,  723,  803, 
806,   939,   940,   990 

Barban,  S.,  32,  388,  389,  391,  400,  990 

Barbato,'L.  M.,  847,  990 

Barber,  G.  A.,  137,  138,  389,  990,  1058 

Barbour,  H.  G.,  956,  990 

Barchas,  J.,  310,  1033 

Bargoni,  N.,  547,  549,  559,  990 

Barker,  H.  A.,  282,  287,  288,  995 

Barker,  S.  B.,  18,  783,  990 

Barlow,  A.  J.  E.,  168,  169,  190,  999 

Barman,  T.  E.,  630,  990 

Barnabas,  J.,  15,  987,  990 

Barnard,  E.  A.,  807,  906,  990 

Barnes,  H.,  730,  735,  962,  990 

Barnes,  J.  M.,  225,  990 

Barnes,  M.  N.,  1046 

Barnett,  R.  C,   117,   198,  990,  996 

Barone,  J.  A.,  531,  990 

Barrett,  F.  R.,  953,  990 

Barron,  E.  S.  G.,  29,  41,  60,  74,  92,  105, 
119,  121,  171,  174,  189,  227,  228,  643, 
666,  667,  668,  672,  673,  705,  707,  712, 
713,  714,  717,  718,  722,  781,  803,  805, 
826,  831,  840,  846,  852,  855,  856,  857, 


866,  870,  873,  876,  878,  882,  883,  991, 

1032,  1046,  1049,  1054 
Barry,  R.  J.  C,  387,  991 
Bartett,  W.  L.,  437,  998 
Bartlett,  G.  R.,  341,  991 
Bartlett,  M.  D.,  810,  835,  870,  1001 
Barton,  M.   N.,  577,   1040 
Bartley,  W.,  88,  991 
Basciak,  J.,  988 
Basford,  R.  E.,  543,  553,  711,  848,  851, 

1020 
Bass,  L.,  946,  1011 
Bassham,  J.  A.,   163,  991 
Bastarrachea,  F.,  832,  991 
Batt,   W.  G.,  385,  1069 
Battaglia,  F,  C,  264,  911,  991 
Bauchop,  T.,  432,  593,  600,  853,  991 
Bauer,   C.   W.,   622,   991 
Bauer,   S.,    137,   1037 
Bauer,  W.,  686,  1048 
Bauerle,  R.  H.,  972,  992 
Baugh,  C.   L.,   852,  991 
Baum,  H.  M.,  284,  713,  1035 
Baumann,  C.  A.,  55,  56,  79,  81,  164,  175, 

176,   530,   991,   1040,   1057 
Bauminger,  B.,  657,  658,  1032 
Baur,  H.,  269,  353,  1005 
Baxter,  C.  F.,  64,  857,  991 
Baxter,   R.   M.,   589,   991 
Bayne,  S.,  384,  991,  1038 
Beach,  J.,  399,  400,  1056 
Bean,   R.   C,   385,   991 
Beaton,  G.  H.,  566,  991 
Beaton,  J.   R.,  571,  573,  991 
Beattie,  D.  S.,  334,  810,  856,  1052 
Beck,  G.  E.,  938,  991 
Beck,  S.  D.,  22,  29,  1016 
Becker,  B.,  209,  267,  465,  991 
Becker,  C.  E.,  383,  384,  385,  991 
Becker,  M.,  269,  353,  1005 
Becker,  V.,  18,  219,  220,  991 
Beckmann,  F.,  212,  991 
Beechey,  R.   B.,   121,  991 
Beers,  R.  F.,  Jr.,  463,  991,  1018 
Beevers,  H.,  61,  77,  90,  164,  171,  172,  185, 

189,  191,  274,  394,  591,  848,  991,  992, 

998,   1013,   1023,   1034 
Behal,  F.  J.,  195,  782,  992 


1074 


AUTHOR    INDEX 


Beher,  W.  T.,  494,  495,  504,  992,  1010 

Behne,    I.,    396,    1019 

Beiler,  J.  M.,  586,  587,  1036 

Bein,   H.   J.,   938,   991 

Beinert,  H.,  138,  143,  233,  535,  668,  687, 

712,  853,  992,  999,  1028 
Beling,  C.  G.,  464,  992 
Belkin,  M.,  201,  670,  992 
Bell,  F.  E.,  831,  992 
Bellin,  225,  992 
Bellinger,  S.  B.,  913,  1050 
Bello,  L.  J.,  475,  992 
Benedict,  F.  G.,  951,  998 
Benesch,  R.  E.,  638,  640,  645,  646,  746, 

757,  763,  902,  903,  931,  932,  992 
Ben-Gershom,    R.,   383,   1044 
Benigno,  P.,  983,  984,  992 
Bennett,  D.  R.,  625,  992 
Bennett,  E.  D.,  972,  992 
Bennett,  H.  S.,  766,  767,  992 
Bennett,   L.   L.,  Jr.,  585,  1055 
Benson,  A.,  888,  1032 
Benson,  A.  A.,  991 
Bentley,  L.  E.,  225,  992 
Bentley,  R.,  226,  228,  678,  992 
Ben-Zvi,  E.,  739,  1008 
Beppu,  T.,  844,  992 
Ber,   A.,   532,   992 
Berends,  W.,  538,  1030 
Beresotskaya,  N.  A.,  395,  1006 
Berg,   v.,  875,  1028 
Bergel,  F.,  516,  992 
Berger,  A.,  457,  1024 
Berger,   J.,   53,   60,   81,   596,   992 
Berger,   M.,    119,   992 
Bergersen,  F.  J.,  292,  294,  992 
Bergquist,  P.  L.,  881,  908,  993 
Bergkvist,  R.,  687,  992 
Bergmann,  F.,  281,  285,  816,  992,  993 
Bergmeyer,   H.-U.,  841,  993 
Berk,  S.,  228,  993 
Berkowitz,  J.,  333,  1012 
Berl,   S.,  574,   1046 
Bertin,  M.,   958,  960,  993 
Berliner,  R.  W.,  917,  1043 
Berman,  D.  A.,  13,  75,  131,  191,  215,  216, 

228,  941,  943,  993,  1001,  1036,  1048 
Bernath,  P.,  18,  713,  783,  825,  856,  1054 


Bernfeld,   P.,   454,   459,   463,   993 
Bernhard,   S.   A.,   375,   993 
Bemheini,    A.    I.,    724,    993 
Bernheim,   F.,  225,  228,  239,  259,  272, 

693,  694,  888,  993,  1003,  1040 
Bernstein,  D.  E.,  52,  74,  81,  993 
Bernstein,  G.  S.,  391,  400,  993 
Bernstein,   I.   A.,   413,   1055 
Berry,  L.  J.,  203,  221,  222,  223,  228,  993 
Bersin,   T.,    686,    693,   993 
Bertelsen,  K.,  817,  1017 
Bertino,  J.  R.,  581,  993 
Besemer,   A.   R.,   971,   973,   1000 
Bessey,  O.  A.,  286,  287,  288,  289,  1033 
Bessman,  M.  J.,  475,  992 
Bessman,   S.   P.,   433,   1008 
Best,  A.  N.,  585,  1061 
Betzel,  R.,  974,  1018 
Beuzeville,  C.,  221,  223,  228,  993 
Beyer,  K.  H.,  264,  917,  993 
Beyer,   R.   E.,   540,   544,   547,   548,   993 
Bhatia,  I.  S.,  213,   747,  1065 
Bhausar,   M.   D.,   415,   1050 
Bhuvaneswaran,   C,   478,   993 
Bhuyan,  B.  K.,  713,  859,  993 
Biale,  J.  B.,  74,  78,  89,  81,  82,  91,  120, 

173,   182,  989,  1032,  1054 
Bickers,  J.  N.,  926,  927,  993 
Bickis,  I.  J.,   156,  1046 
Bide,  R.  W.,  553,  774,  851,  993 
Bieber,   R.   E.,   497,   1024 
Bieber,  S.,  283,  1006 
Bieleski,  R.  L.,  910,  993 
Bielinski,  T.  C,  587,  995 
Bilinski,   E.,   137,  994 
Bilodeau,   F.,  212,  994 
Bilse,  I.  M.,  632,  1057 
Binkley,   F.,   802,   1052 
Bird,  H.  R.,  530,  1040 
Birkenhager,  J.  C.,  674,  676,  677,  994 
Birmingham,  M.  K.,  150,  1051 
Birt,  L.  M.,  74,  75,  79,  82,  171,  1002 
Bishop,    J.    S.,    401,    988 
Bisno,  A.  L.,  937,  1066 
Biszku,  E.,  649,  788,  803,  812,  817,  827, 

1059 
Bivings,  L.,  954,  994 
Bjerrum,  J.,  732,  739,  994 


AUTHOR    IXDEX 


1075 


Black,  M.  M.,  202,  994 

Black,  R.  E.,  391,  400,  993 

Black,  S.,  832,  994 

Blackburn,  C.  R.  B.,  902,  908,  1063 

Blattig,   K.,   956,   1061 

Blair,  D.  G.  R.,  473,  479,  994 

Blakley,  E.  R.,  403,  404,  542,  839,  994, 

999 
Blakley,  R.  L.,   137,  581,  585,  994 
Blanchard,  M.,  60,  338,  348,  994 
Blaschko.   H.,   308,   363,   364,  592,   694, 

835,   994 
Blaylock,  B.  A.,  551,  843,  994 
Blessing,  J.  A..  315,  1057 
Blickenstaflf.  D.  D.,  985,  994 
Bloch,  D.   I.,   780,  798,  865,  1050 
Bloch-Frankenthal,  L.,  383,  396,  1047 
Blum,  J.  J.,  445,  446,  816,  819,  867,  994 
Blum,  K.  U..  520,  1063 
Blumberg,  A.  J.,  959,  987 
Blumenstein,  J.,  709,  994 
Blumenthal,  H.  J.,  356,  440,  443.  984, 

994,  1012,  1029 
Blumson,  N.  L.,  332,  994 
Bocher,  C.  A.,  462,  1043 
Bock,  K.  ,R.,  966,  994 
Bock,  M.,  977,  978,  1013 

Bock,  R.  M.,  549,  832,  844,  1035,  1050 

Bodansky.  O.,  413.  994 

Boeri,  E.,  65,  435,  437,  552,  708,  870,  994 

Boese,   A.   B.,  620,  1057 

Bogdonoff,   M.   D.,   399,   1030 

Boger,  W.  P.,  626,  1051 

Boggiano,  E.  M.,  356,  1055 

Eokman,  A.  H.,  136,  1021 

Bolker,  H.,  202,  994 

Bomford,  R.  R.,  951,  952,  953,  954,  1021 

Bonaduce,   L.,   469,   675,   706,   1051 

Bone,  A.  D.,  379,  1065 

Bone,   D.   H.,   878,   994 

Boney,  A.  D.,  967,  994 

Bonner,  B.  A.,  27,   168,  994 

Bonner,  D.  M.,  418,  595,  1023.  1031 

Bonner,  J.,  53,  58,  78,  80,  93,  120,  167, 
169,  170,  172,  182,  184,  189,  191,  297, 
516,  911,  967,  994,  995,  996, 1000, 1023 

Bonner,  W.  D.,  Jr.,  32,  33,  196,  966,  968, 

995,  1055,  1060 


Bonnichsen,  R.,  784,  1060 

Bonsignore,  A.,  412,  995 

Booth,  A.  X.,  601,  995 

Borghgraef,  R.  R.  M.,  923,  928,  929,  960, 

995 
Borries,  E.,  570,  1004 
Borsook,  H.,  887,  995 
Borst,  P.,  Ill,  152,  995 
Bortz,  W.  M.,  614,  995 
Bosch,  L.,   130,  1063 
Bose,  S.  M.,  682,  687,  693,  1003 
Boser,  H.,  837,  995 
Bosund,  I.,  349,  995 
Boucek,  R.  J.,  215,  995 
Bovarnik,  M.  R.,  341,  347.  547,  549,  556, 

557,   558,   560,   780,   995,   1017 
Bowman,  E.  M..  772,  827.  828,  838,  1011 
Bowman,  I.  B.  R.,  91,  127,  560,  995 
Boxer,  G.   E.,  587,  995 
Boyd,  E.  S.,  54,  58,  166,  175,  995 
Boyer,  P.  D.,  351,  403,  404,  635,  640, 

643,  645,  649,  714,  745,  760,  763,  764, 

765,  766,  780,  788,  803,  806,  810,  812, 

833,  853.  994,  995,  1025,  1027,  1052, 

1059,   1066 
Boyland,  E.,  138,  140,  177,  179,  201,  218, 

224,  225,  428,  990,  995 
Boyland,  M.  E.,  138,  140,  177,  179,  995 
Boylen,  J.  B.,  305,  995 
Bradley,  L.  B.,  807,  816,  856,  866,  873, 

1022,  1026 
Bradley,  R.  M.,  614,  1048 
Bradshaw,  W.  H.,  282,  287,  288,  995 
Brady,  C.  J.,  693,  995 
Brady,  R.  O.,   148,  613,  614,  950,  995, 

1048 
Brand,  E.,  665,  666,  667,  995 
Branster,  M.  V.,  485,  493,  995 
Brasted,  R.  C,  736,  1055 
Braun,  K.,  494,  995 
Braune,  W.,  660,  995 
Braunstein,  A.  E.,  360,  561,  995 
Bray,  R.   E.,  725,  949,  995,  1062 
Bregoff,  H.  M.,  293,  996 
Bremer,  J..  663,  1006 
Brendel,  R.,  586,  587,  1036 
Brenner,  B.  M.,  472,  1052 
Bresler,   E.   H.,   926,   927,   993 


1076 


AUTHOR    INDEX 


Breslow,  E.,  542,  547,  550,  781,  1035 
Bresnick,  E.,  429,  468,  469,  480,  522,  834, 

996,  998,  1013 
Bressler,  R.,  224,  234,  781,  887,  996 
Breuer,  H.,  708,  1038 
Breusch,  F.  L.,  60,  175,  177,  996 
Brewer,  C.  R.,  78,  996 
Brewer,  J.  H.,  978,  979,  980,  1027 
Brian,  P.  W.,  618,  632,  1034 
Briault,  P.  G.,  983,  996 
Bridger,  W.  A.,  466,  481,  831,  996,  1000 
Brierley,  G.  P.,  238,  1063 
Briggs,  F.  N.,  179,  996 
Brignone,  J.  A.,  38,  41,  713,  775,  779, 

783,  1058 
Brindley,  C;  O.,  538,  851,  1029,  1030 
Brink,  N.  G.,  410,  500,  996 
Brixovd,  E.,  954,  957,  1005 
Brock,   N.,  965,   996 
Brocket!,  F.  P.,  813,  1040 
Brocklehurst,  W.  E.,  374,  989 
Brockman,  R.  W.,  480,  996 
Brode,  E.,  585,  1022 
Brodersen,  R.,  617,  632,  996 
Brodie,  A.  F.,  120,  552,  846,  848,  989, 

996,  1024 
Brodie,  B.  B.,  851,  1012 
Brodman,  K.,  946,  947,  948,  1050 
Brody,  K.  R.,  267,  1056 
Brody,  J.  M.,  348,  996 
Bronk,  D.  W.,  211,  1030 
Brooke,  J.  L.,  393,  394,  999 
Brooks,  J.,  735,  875,  1023 
Brooks,  J.  L.,  392,  403,  996 
Brooks,  P.,  739,  740,  996 
Brooks,  S.  A.,  388,  389,  996 
Broquist,  H.  P.,  583,  996 
Brosteaux,'  J.,  437,  1013 
Brown,  D.  D.,  843,  996 
Brown,  D.   E.  S.,  211,  1053 
Browm,   D.   H.,  411,   561,   1012  1022, 
Brown,  D.  M.,  696,  770,  782,  844,  999, 

1055 
Brown,  E.  G.,  162,  996 
Brown,  G.  W.,  Jr.,  141,  998 
Brown,  I.  A.,  954,  996 
Brown,  J.,   401,   996 
Brown,   P.   K.,  953,  1064 


Brown,   R.   R.,   1003 

Brown,  W.  D.,   137,  996 

Bruce,  M.,  898,  1049 

Bruchmann,  E.  E.,  692,  694,  996 

Bruemmer,   J.   H.,    17,   26,   996 

Brim,  C.,  918,  923,  996 

Brunfaut,   M.,   969,   1057 

Brunnemann,   A.,  499,  996 

Brunngraber,  E.  G.,  473,  996 

Bruns,  F.  H.,  406,  855,  996,  1042 

Brust,   M.,    198,   996 

Bryce,  G.  F.,  787,  996 

Bublitz,  C.  842,  996 

Buch,  M.  L.,  15,  996 

Buchanan,  J.  M.,  333,  1009,  1031 

Buchman,  E.  R.,  516,  996 

Buckland,  F.  E.,  977,  978,  979,  988 

Buddecke,  E.,  852,  875,  881,  884,  1065 

Budillon,  G.,  616,  1049 

Budzilovich,  T.  V.,  587,  995 

Biicher,  T.,  61,  1027 

Bueding,  E.,  376,  381,  382,  383,  842,  997 

Bulen,  W.  A.,  614,  708,  840,  997 

Bullough,  W.  S.,  199,  200,  225,  990,  997 

Bu'Lock,  J.  D.,  231,  997 

Bunzel,  H.  H.,  981,  1068 

Burch,   G.,   928,  997 

Burch,  H.  B.,  540,  997 

Burchenal,  J.  H.,  582,  583,  996,  1009 

Burdette,  W.  J.,  77,  79,  82,  91,  179,  997 

Burgess,  E.  A.,  81,  176,  183,  997 

Burk.  D.,  292,  388,  392,  400,  401,  1029, 

1030,  1037 
Burkard,  W.  P.,  318,  1045 
Burnett,  A.   E.,  923,  1065 
Burnett,  G.  H.,  592,  997 
Burney,  T.  E.,  728,  979,  980,  981,  997 
Burns,   S.,   675,   1004 
Burr,  M.  J.,   887,  997 
Burris,  R.  H.,  53,  61,  171,  291,  292,  293, 

294,  997,  999,  1005,  1019,  1032,  1038, 

1039,  1067 
Burton,  K.,  268,  341,  348,  545,  547,  549, 

997 
Busch,  H.,  15,  91,  98,  100,  101,  102,  103, 

104,  152,  156,  201,  217,  218,  228,  232, 

437,  997,  1042,  1046 
Butler,   L.   G.,   672,   1031 


AUTHOR    INDEX 


1077 


Butt,   E.  M.,  960,  1014 
Butt,  W.  D.,  451,  997 
Buzard,  J.  A.,  709,  718,  997 
Bydgeman,   S.,  401,   1019 
Byers,   S.   O.,   288,   997 
Byrne,  W.  L.,  270,  390,  412,  853,  1017, 
1041 


Cadenas,  E.,  125,  414,  1039,  1056 

Cady,  P.,   146,  997 

Caffrey,  R.  W.,  543,  553,  851,  864,  997, 

1020 
Cafruny,  E.  J.,  689,  918,  920,  921,  922, 

933,   997,  1063 
Cailleau,  R.,  62,  63,  1033 
Cain,   R.   B.,  90,  997,  998 
Calcutt,  G.,  982,  997 
Caldwell,  E.  F.,  569,  997 
Caldwell,  M.  L.,  658,  660,  662,  673,  674, 

683,  684,  833,  1065 
Calesnick,  B.,  960,  997 
Caley,  E.  R.,  948,  1016 
Calkins,   E.,  304,  1031 
Callaghan,  O.  H.,  446,  772,  847,  997 
Callahan,  S.,  283,  1006 
Calo,  N.,  892,  1012 
Caltrider,  P.  G.,  133,  997 
Calvin,  M.,  145,  163,  409,  619,  637,  646, 

656,  658,  991,  997,  1002,  1036,  1043 
Cama,   H.  R.,  547,  549,  660,  685,  832, 

1029,   1047 
Cameron,  G.,  201,  998 
CampbeU,  A.  D.,  670,  690,  1006 
Campbell,   D.   E.   S.,   923,   924,   998 
Campbell,  L.  L.,  Jr.,  269,  833,  834,  1035 
Campbell,  P.  N.,  272,  405,  998 
Canellakis,  Z.  N.,  305,  306,  998 
Caniggia,  A.,  215,  1031 
Cannata,  J.  J.  B.,  711,  852,  998 
Cantero,   A.,  381,  1046 
Cantino,  E.  C,  169,  1033 
Capobianco,  G.,  616,  1049 
Carafoli,   E.,   121,   122,   865,  989 
Caravaca,  J.,   439,   1014 
Caraway,  W.   T.,   703,  998 
Carbon,  J.  A.,  481,  998 
Carney,  S.  A.,  391,  998 


Carpenter,  A.  T.,  600,  998 

Carpenter,  T.  M.,  951,  998 

Carpenter,  W.  D.,  61,  998 

Carr,  A.  J.,  428,  998 

Carsiotis,  M.,  844,  852,  1054 

Carson,  S.  F.,  15,  1044 

Carter,  C.   E.,  282,  359,  1002,   1029 

Carter,   J.   R.,   650,   1047 

Cartwrigt,  N.  J.,  90,  998 

Casal,  A.,  657,  1003 

Cascarano,  J.,  879,  895,  998 

Castelfranco,  P.,  136,  998 

Catterson,  D.  A.,  959,  987 

Caughey,  W.  S.,  61,  329,  591,  998,  1061 

Cavallito,  C.  J.,  621,  973,  983,  998,  1027 

Cecil,  R.,  756,  988,  998 

Cedrangolo,  F.,  668,  998 

Cerecedo,  L.  R.,  516,  518,  519,  522,  523, 

530,  531,  533,  998,  1005,  1006,  1056 
Cereijo-Santalo,   R.,   396,   397,   1065 
Cerfontain,  H.,  745,  998 
Cha,  C.-Y.,  683,  998 
Chaberek,  S.,  13,  739,  998 
Chadwick,  J.  B.,  169,  880,  1041 
Chaffee,  E.,  875,  1021 
Chaffee,  R.  R.,  437,  998 
Chaikoff,  I.  L.,  141,  145,  146,  147,  179, 

209,  226,  234,  273,  987,  996,  997,  998, 

1025,  1060 
Challenger,  F.,  2,  225,  226,  228,  998,  1064 
Chamberlain,  G.  T.,  966,  994 
Chambers,  R.,  201,  998 
Chan,  P.  C,  792,  866,  867,  1058 
Chance,   B.,   17,   786,  998 
Chang,  V    M.,  555,  1019 
Chanutin,   A.,   68,   1039 
Chaplain,  R.  A.,  874,  1014 
Chapman,  D.  W.,  944,  957,  998 
ChappeU,  J.  B.,  86,   121,  210,  816,  998 
Chari-Bitron,  A.,  31,  104,  239,  828,  998 
Charmandairjan,  M.   O.,   835,   999 
Chattaway,   F.   W.,   168,    169,   190,   999 
Chatter jee,  G.  C,  529,  530,  576,  999, 1002 
Chaudhuri,  D.  K.,  835,  999 
Chaure,  V.  J.,   187,  208,  218,  883,  914, 

915,  916,  1002 
Cheesman,  D.  F.,  939,  999 
Cheftel,  R.  L.,  15,  999 


1078 


AUTHOR    INDEX 


Cheldelin,  V.  H.,  38,  52,  132,  138,  143, 

705,  712,  837,  853,  999,  1002,  1023, 

1026,  1033 
Chemelin,  I.,  399,  400,  1056 
Chen,  K.  K.,  999 

Chen,  R.  F.,  509,  513,  696,  782,  844,  999 
Cheniae,  G.,  33,  65,  850,  855,  999 
Chenoweth,  M.  B.,  212,  625,  992,  1065 
Chernick,  S.,  179,  996 
Chiba,  H.,  675,  685,  692,  771,  772,  835, 

999 
Chiba,   Y.,   194,  999 
Chick,  H.,  971,  972,  999 
Chiga,  M.,  872,  874,  999 
Chin,  H.  P.,  403,  999 
Chinard,  F.  P.,  640,  643,  657,  662,  665, 

669,  697,  701,  702,  703,  727,  803,  999, 

1017 
Chirigos,  M.  A.,  266,  999 
Choppin,  P.  W.,  976,  977,  979,  980,  981, 

999 
Chow,  C.  T.,  173,  182,  1054 
Christensen,  E.,  392,  393,  394,  403,  435, 

996,  999 
Christensen,  H.  N.,  15,  155,  266,  575,  908, 

999,  1043,  1048 
Christian,  W.,  768,  1064 
Christie,  G.  S.,  81,  999 
Christman,  A.  A.,  236,  999 
Christophers,  S.  R.,  559,  1050 
Chung,  A.  E.,  781,  839,  999 
Cianci,  V.,  971,  972,  999 
Ciferri,  O.,  404,  411,  542,  839,  994,  999 
CifoneUi,  J.  A.,  385,  1004 
Ciotti,  M.  M.,  487,  496,  497,  498,  504, 

508,  784,  1024,  1062,  1069 
Citri,  N.,  249,  599,  615,  688,  999,  1011 
Clagett,  C.  O.,  61,  999 
Clark,  D.  S.,  74,  77,   190,  228,  999 
Clark,  I.,  562,  572,  1045 
Clark,  W.  C,  520,  1067 
Clark,  W.  G.,  310,  314,  1016,  1045 
Clark,  W.  M.,  656,  999 
Clarke,  D.  D.,  887,  1000 
Clarkson,  T.  W.,  916,  920,  1000 
Claus,  D.,  880,  971,  972,  984,  1000 
Claus,  G.  W.,  852,  991 
Clayton,  R.  A.,  839,  885,  1000 


Clayton,  R.  K.,  53,  77,  78,  81,  146,  1000 
Cleland,  K.  W.,  22,  29,  174,  175,  817,  855, 

874,  875,  882,  884,  1000 
Cleland,   R.,   967,   1000 
Cleland,  W.  W.,  640,  1000 
Clementi,  A.,  228,  1000 
Clements,  A.  N.,  54,  77,  1000 
Cliffe,  E.  E.,  544,  1000 
Clowes,  G.  H.  A.,  198,  999,  1028 
Coates,  J.  H.,  287,  989 
Cochin,  J.,  590,  597,  989 
Cochran,  D.  G.,  332,  446,  1038,  1050 
Cochrane,  V.  M.,  79,  97,  1000 
Coe,  E.  L.,  396,  397,  1021 
Coen,  L.  J.,  575,  1022 
Coggeshall,  R.  E.,  499,  1000 
Cohen,  E.  M.,  873,  927,  1000,  1003 
Cohen,  G.  N.,  321,  323,  356,  1057,  1061 
Cohen,  L.   H.,  466,  478,  481,  831,  996, 

1000 
Cohen,  P.   P.,   116,   157,  305,  306,   336, 

592,    781,   804,   840,   997,   998,   1000, 

1003,  1032,  1036 
Cohen,  R.  A.,  55,  177,  1000 
Cohen,   S.,  226,   1000 
Cohen,  S.   S.,  476,  479,   1008,  1036 
Cohn,  M.,  400,  415,  416,  780,  1000,  1066 
Coleman,   G.   H.,   974,   1000 
Coleman,  J.   E.,   770,   818,  1000,  1062 
Coleman,  R.,  662,  1000 
Collander,    R.,    189,    1000 
Collias,  E.  C,  29,  1000 
Collier,  H.  B..  553,  774,  825,  833,  836, 

845,  851,  854,  859,  901,  906,  993,  1056, 

1061 
Collins,  A.,  591,  1061 
Colowick,  S.  P.,  273,  405,  433,  434,  435, 

485,  487,  490,  493,  498,  504,  507,  508, 

509,    511,    512,    783,    896,    912,    1000, 

1001,   1006,   1024,   1027,   1041,   1043, 

1046,  1069 
Colpa-Boonstra,  J.,  32,  35,  1020 
Colter,   J.   S.,   791,   794,   815,   817,   820, 

1000,   1006 
Combs,  A.  M.,  540,  997 
Commoner,  B.,   18,  22,   196,  1000,  1019 
Conches,   L.,   74,   77,   81,   92,    116,    117, 

169,  1058 


AUTHOR  INDEX 


1079 


Conchie,  J.,  428,  429,  812,  814,  838,  1000 
Conn,  E.  E.,  27,  74,  80,  82,  119,  120,  355, 

852,  1000,   1028,  1057 
Connamacher,  R.  H.,  316,  317,  318,  139, 

320,  1018 
Connelly,  C.  C,  813,  1040 
Connor,  R.,  619,  987 
Connors,  P.,  208,  1063 
Conover,  T.  E.,  121,  1000 
Conrad,  619,  1000 

Contopoulou,  C.  R.,  136,  675,  998,  1004 
Contrera,  J.  F.,  320,  599,  1013 
Converse,  J.  L.,  971,  973,  1000 
Conway,  T.  W.,  354,  1001 
Cocul,  B.  J.,  190,  196,  1001 
Cook,  A.  M.,  971,  972,  1001 
Cook,  E.  S.,  744,  810,  835,  856,  870,  871, 

972,  973,  1001,  1028,  1052,  1055,  1056, 

1060 
Cook,  R.  P.,  3,  78,  228,  1001 
Cook,  S.  F.,  880,  1001 
Coombs,  T.  L.,  781,  827,  1001 
Coon,  M.  J.,  782,  834,  1029,  1048 
Cooper,  C,  444,  445,  705,  869,  872,  1001, 

1031 
Cooper,  E.  A.,  195,  972,  1001 
Cooper,  J.,  855,  1059 
Cooper,  J.  A.  D.,  550,  1001 
Cooper,  J.  R.,  354,  1062 
Cooperstein,  S.  J.,  175,  663,  1001,  1010 
Coore,  H.  G.,  376,  1001 
Copenhewer,  J.  H.,  Jr.,  119,  1001 
Coper,  H.,  499,  996 
Cori,  C.  F.,  376,  387,  388,  390,  405,  561, 

648,  789,  803,  808,  811,  813,  1001, 

1022,  1026,  1034,  1055 
Cori,  G.  J.,  376,  405,  1001,  1055 
Cori,   0.,  60,    119,    121,    707,    717,    873, 

1049 
Corley,  R.  C,  2,    219,    273,    610,    1001, 

1053,  1056 
Cormier,  M.  J.,  845,  891,  1001 
Corner,  E.  D.  S.,  967,  994 
Corte,  G.,  586,  1023 
Corwin,  A.  H.,  213,  747,  1065 
Corwin,  L.  M.,  63,  878,  1001 
Cotter,  G.  J.,  696,  1007 
Cottrell,  T.  L.,  744,  1001 


Cousins,  F.  B.,  543,  1001 

CoveUi,  I.,  546,  972,  1007 

Covin,  J.  M.,  128,  216,  1001 

CowgiU,  R.  W.,  405,  852,  853,  1001 

Coxon,   R.   v.,   76,  1001 

Coyne,  B.  A.,  575,  999 

Cramer,  F.  B.,  386,  389,  392,  1001 

Cramer,  H.  I.,  619,  987 

Crandall,  D.  I.,  541,  545,  709,  771,  772, 

787,  1001,  1008 
Crane,  F.  L.,  16,  17,  26,  996,  1014 
Crane,  R.  K.,  262,  264,  379,   383,  388, 

390,  782,  824,  843,  1001,  1008,  1028, 

1056 
Cravens,  W.  W.,  530,  576,  1001,  1040 
Crawford,  E.  J.,  286,  287,  288,  289,  407, 

474,  1024,  1033 
Crawford,  M.  A.,   109,  1001 
CrawhaD,  J.   C,   165,   1001 
Creaser,  E.  H.,  122,  177,  200,  395,  1001 
Creasey,  N.  H.,  272,  405,  998 
Creeth,  J.  M.,  682,  1067 
Cremer,  J.  E.,    75,   76,    127,   153,  1001, 

1002 
Cremona,  T.,  65,  435,  436,  437,  510,  798, 

872,  994,  1002,  1038 
Creveling,   C.   R.,   316,   320,    611,   1002, 

1062 
Creveling,  R.  K.,  466,  1037 
Crewther,  W.  G.,  593,  660,  693,  708,  1002 
Crocker,  B.  F.,  418,  1053 
Cronin,  M.  A.,  358,  1002 
Cross,  A.  C,  916,  1000 
Cross,  R.  J.,  204,  928,  1002 
Crout,  J.   R.,  592,  611,  1002 
Cruickshank,  C.  N.  D.,  225,   990 
Cruickshank,  D.   H.,   64,   1002 
Csonka,  E.,  957,  959,  961,  1038 
Cumings,  J.  N.,  952,  1002 
Cummins,   J.  T.,  154,   712,   1002,  1050 
Cunningham,  L.  W.,  681,  745,  750,  1002 
Cunningham,   L.   W.,  Jr.,  513,   1062 
Cunningham,  M.,    137,    138,    141,    142, 

1011 
Cuong,  T.  C,  225,  226,  1009 
Curl,  A.  L.,  1023 
Curtis,  W.  C,  834,  1061 
Cutolo,   E.,  552,   708,  994 


1080 


AUTHOR    INDEX 


Czarnetzky,  E.  J.,  956,  972,  1055 
Czekalowski,  J.  W.,   194,  1002 


D 


D'Abramo,   F.,    164,   1049 

Da  Costa,  F.  M.,  315,  1012 

Dahl,  J.   L.,   478,   1002 

Dahlqvist,  A.,  416,  1002 

Dajani,  R.  M.,  91,  1002 

Dale,  R.  A.,  920,  921,  1002 

Dalgarno,  L.,  74,  75,  79,  82,  171,  1002 

Daly,  J.  W.,  316,  611,  1002    1062 

Dalziel,  K.,  513,  1002 

Damaschke,   K.,   891,   1002 

Damjanovich,  S.,  949,  1016 

Dancis,  J.,  585,  990 

Danforth,  W.,  28,  51,  53,  56,  77,  228, 1002 

Daniel,   L.  J.,   530,  585,   1002,  1058 

Danielski,  J.,  91,  1002 

Danielson,  L.,  708,  849,  1006 

Dann,  O.  T.,  359,  1002 

Darlington,   W.   A.,   207,   1002 

Darragh,  J.  H.,  925,  1065 

Das,  H.  K.,  153,  1002 

Das,  N.  B.,  32,  38,  48,  62,  177,  178,  437, 

1002 
Das,  S.  K.,  529,  1002 
Das,  S.  N.,  9,  1002 
Datta,  A.,  548,  556,  705,  865,  869,  873, 

1014,  1046 
Datta,   A.   G.,  519,   1002 
Daus,  L.,   145,  1002 
d'Aurac,  J.,  225,  226,  1009 
Da  Vanzo,  J.   P.,  358,  1002,  1052 
Davenport,  H.  W.,   187,  208,  218,  883, 

914,  915,  916,  1002 
Davenport,  V.   D.,  883,  914,  915,  1002 
Davidson,  E.,  356,  1012 
Davidson,  H.  M.,  63,  442,  1041 
Davidson,  J.  D.,  281,  282,  466,  477,  1007 
Davidson,  N.,  739,  740,  741,  996,  1008, 

1069 
Davies,  D.  D.,  349,  593,  604,  607,  857, 

990,  1006 
Davies,  G.  E.,  224,  236,  1003 
Davies,  J.  H.,  428,  995 
Davies,  R.  E.,  88,  991 


Davies,  R.  I.,  264,  911,  1031 

Davis,  F.  F.,  712,  1003 

Davis,  J.,  348,  1024 

Davis,  N.  C,  783,  1055 

Davison,   D.   C.,   170,   1003 

Dawes,  E.  A.,   124,  432,  593,  600,  853, 

991,  1003 
Dawid,  I.  B.,  333,  1009 
Dawson,  C.  R.,  697,  1060 
Dawson,   R.  M.  C,  75,   151,  858,  1003, 

1060 
Day,  B.,  978,  979,  980,  1027 
Day,  H.  G.,  385,  520,  525,  530  991,  1010 
Day,  M.  D.,  318,  1003 
Day,  R.  A.,  333,  1009 
Dearborn,  E.  H.,  332, 550, 551, 1013, 1062 
Debay,  C.  R.,  834,  1006 
de   Bernardinis,   C,   217,   218,   1015 
De  Bodo,  R.  C,  401,  988 
De  Boer,  C.  J.,  461,  462, 1018 
de  Caro,  L.,  520,  522,  526,  527, 1003, 1048 
Dechary,  J.  M.,  593,  1028 
Decker,  R.  H.,  610,  782,  843,  1003 
De  Eds,   F.,  601,  995 
de  Favelukes,  S.  L.  S.,  74,  77,  81,  92, 

116,   117,   169,  1058 
Deferrari,  J.  O.,  687,  712,  716,  775,  783, 

810,  853,  1058 
De  Flora,  A.,  407,  1013 
De  Gowin,  E.  L.,  902,  905,  1053 
De  Graff,  A.  C.,  944,  1003 
De  Groot,  C.  A.,  927,  1003 
De  Groot,  N.,  840,  1003 
de  Issaly,  I.  S.,  74,  79,  81,  1003 
Deitrich,  R.  A.,  662,  1003 
Deitz,  V.  R.,  643,  662,  669,  803,  1017 
Dejung,  K.,  925,  1003 
de  Kock,  L.  L.,  133,  1040 
de  Kock,  P.  C,  133,  1040 
de  la  Fuente,  G.,  389,  1056 
Delaunay,   A.,   203,   1030 
del  Campillo,  A.,  708,  837,  1057 
Dellert,  E.  E.,  458,  1003 
Delmonte,   L.,  581,  1003 
De  Luca,  H.  A.,  150,  151,  1027 
Demis,   D.  J.,  876,  883,  893,   894,  895, 

897,  898,  1003 
De  Moss,  J.  A.,  89,  1003 


AUTHOR    INDEX 


1081 


De  Moss,  R.  T>.,  831,  1003 

Dempsey,  W.  B.,  569,  1009 

Dengler,  H.,  314,  1003 

Denison,  F.  W.,  Jr..  15,  1003,  1044 

Dennis,   D.,  434,   1003 

Denstedt,  O.  F.,  55,  65,   179,  436,  485, 

487,   489,   492,   503,   988,  1035,  1043, 

1049 
de  Petrocellis,   B.,  469,   675,   706,   1051 
Derbyshire,  J.  E.,  203,  223,  993 
De  Robertis,  E.,  203,  727,  1003 
Dervartanian,  D.  V.,  18,  38,  1003 
Desnuelle,   P.,   649,   657,   664,   666,   668, 

1003,  1028 
De  Stevens,  G.,  461,  462,  1018 
Dettmer,  F.  H.,  53,  77,  78,  81,  146,  1000 
De  Turk,  W.  E.,  228,  888,  993,  1003 
Devlin,  J.  M.,  872,  1031 
Dewan,  J.  G.,  61,  594,  1014 
Dewey,  D.   L.,  593,   708,  1003 
Deykin,   D.,  835,  1003 
Dhar,  S.  C,  682,  687,  693,  1003 
Diamant,  E.  J.,  495,  1015 
Di  Carlo,  F.  J,,  659,  660,  662,  674,  683, 

684,   792,   833,   1003 
Dick,   G.   F.,   666,   667,  991 
Dickens,  F.,  36,  41,  503,  659,  662,  695, 

711,  855,  1003 
Dickerman,  H.  W.,  490,  850,  1003 
Dickman,   S.,   852,  991 
Dickman,  S.  R.,  273,  462,  592,  1003,  1056 
Diczfaliisy,  E.,  461,  464,  992,  1004 
Dieterle,  W.,   15,   1004 
Dietrich,  E.  V.,  40,  1004 
Dietrich,  L.  S.,  36,  37,  41,  240,  288,  505, 

538,  569,  570,  1004,  1010,  1053 
Dikstein,  S.,  387,  436,  437,  991,  1004 
Dilley,   R.  A.,  557,  1004 
Dils,   R.,    147,   1004 
Dilworth,   M.   J.,   293,   1044 
d'Incan,    E.,   236,   1049 
Dintzis,  H.  M.,  757,  1021 
D'lorio,  A.,  612,  725,  947,  1004 
Dippy,  J.  F.  J.,  299,  1004 
di  Frisco,  G.,  594,  1063 
Di  Sabato,  G.,  808,  1004 
Dische,  Z..  414,  465,  1004 
di  Stefano,  H.  S.,  689,  921,  922,  997 


Dittrich,   S.,    15,   1004 

Dixon,  D.  H.,  594,  694,  1004,  1051 

Dixon,  K.  C,  384,  1004 

Dixon,  M.,  280,  693,  694,  993,  1004 

Dizon,  F.  Y.,  574,  577,  1051 

do  Amaral,  D.  F.,  511,  710,  849,  1035 

Dobson,   M.   M.,   447,   1059 

Doctor,   V.   M.,   577,   582,   1004 

Dodgson,  K.  S.,  444,  684,  1004 

Doherty,   D.   G.,  373,  1063 

Doherty,  M.  E.,  776,  777,  788,  836,  1013 

Dohi,  S.  R.,  461,  1046 

Doi,  R.  H..  552,  553,  710,  848,  1004 

Doi,  Y.,  969,  1025 

Doisy,  E.  A.,  839,  1004 

Doisy,  R.  J.,  283.  859,  1004 

Dohn,  M.  I.,  430,  547,  553,  676,  693,  847, 

849,  1004 
Dominguez,  A.  M.,  1004 
Domonkos,  J.,  179,  1004 
Donovan,  J.  W.,  638,  1004 
Doran,  D.  J..   173,  1004 
Dorfman,  A.,  385,  1004 
Dorman,  P.  J.,  928,  1012 
Doudoroff,  M.,  675,  1004 
Dove,  W.  F.,  741,  1004 
Dowdle,  E.  B.,  208,  909,  914,  1004,  1051 
Downey,  M.,  198,  990 
Downing,  S.  J.,  264,  1049 
Downs,  C.  E.,  335,  1021 
Doyle,  M.  L.,  839,  1004 
Drabikowski,  W.,  939,  940,  1004 
Dragsdorf,  R.  D.,  975,  1016 
Dreser,  H.,  941,  1005 
Dresse,  A.,  611,  989 
Dreyfus,  P.  M.,  519,  1005 
Driver,  G.  W.,  224,  236,  1003 
Druckrey,  H.,  965,  996 
Drury,  D.  R.,  387.  390,  399,  404,  1066 
Dube,  S.  K.,  683,  783,  845,  1005 
Du  Bois,  K.  P.,  22,  30,  41,  659,  676,  687, 

784,  856,  1045 
du  Bay,  H.  G.,  128,  177,  1005 
Duchateua-Bosson,  G.,  297,  1008 
Dudley,   C.,   227,   1053 
Duerkson,  J.  D.,  782,  1005 
Dufait,   R.,   693,   1036 
Duggan,  D.  E.,  281,  282,  285,  1005 


1082 


AUTHOR    INDEX 


Dull,  M.  F.,  620,  1057 

Dulskas,  A.,  926,  1050 

Dumont,  J.   E.,  76,  131,  1005 

Dunn,  A.,  401,  988 

Duperon,  R.,  15,  1005 

Durham,  N.  N.,  267,  613,  1005,  1020 

Duthie,  R.,  363,  994 

Duysens,  L.  N.  M..  891,  1043 

Dyer,  H.  M.,  260,  1005 

Dzurik,   R.,   926,  954,   957,   1005 


Eadie,  G.  S.,  459,  1025 

Eagon,  R.  G.,  389,  1066 

Earl,  J.  M.,   751,   774.   838,   1012.   1062 

Eaton,   M.   D.,    193,   1005 

Ebata,  M.,   842,   1051 

Eber,  J.,  902,  903,  904,  905,  906,  907, 1065 

Eberson,  L.,  9,  1005 

Ebert,  H.,  228.  993 

Eberts,  F.  S.,  171,  1005 

Edelhoch,  H.,  758,  761,  1005 

Edlbacher,   S.,   269,   335,   353,   1005 

Edman,  K.  A.  P.,  938,  1005 

Edsall,  J.   T.,   638,   664.   665.   667,   758, 

759,  761,  1005,  1014 
Edson.  N.  L.,  55,  56.  138,  140,  144,  177, 

237,  238,  1005 
Edwards,  B.  B.,  948.  1016 
Edwards,  J.  G.,  924.  1005 
Eeg-Larsen,  N.,  384,  1005 
Egami,    F.,    537.    541,    544,    817,    1005, 

1041,  1060 
Eggerer,  H.,  836,  887,  987,  1005 
Eggleston.  L.  V.,  22,  27,  74,  75.  77,  94, 

96,  115,  116,  124.  153,  158,  830,  1028, 

1060 
Ehlers,  K.  H.,  221,  223,  993 
Ehrmantraut,   H.,   163,  1005 
Eich,  S.,  518,  519,  522,  523,  531,  998, 1005 
Eichel,  B.,  60,  61,  63,  169,  196,  228,  550, 

987,  1005 
Eichel,  H.  I..  20,  543.  547,  552,  555,  710, 

848,   1005 
Eisen,  H.  N.,  263,  387,  392,  394,  1018 
Eisenstadt,  J.  M.,  269,  1006 
Eisenstark,  A.,  975,  1016 


Eldjarn,  L.,  639,  663,  1006 

El  Hawary,  M.  F.  S.,  110,  111,  1006 

Elion,  G.  B.,  283,  1006 

Elkins-Kaufman,  E.,  365.  1006 

Ellem.  K.  A.  O.,  791.  794,  815,  817,  820, 

1000,  1006 
Ellfolk.  N..  355,  685,  707,  1006 
Ellias,   L.,   423,   772,   1016 
Elhott.  H.  W.,  179,  181.  1006 
ElUott.  K.  A.  C,  75,  91,  178,  185,  212, 

994,  1006 
Elliott,  W.  H.,  471,  1006 
Ellis,  R.  J.,  857,  1006 
Ellis,  S.,  217,  1006 
Ellman,  G.  L.,  640,  1006 
Elodi,  P..  650,  788,  804,  809,  826,  1006, 

1023 
El'tsina,  N.  V.,  395.  1006 
Eltz,  R.  W.,  886.  1006 
Elvehjem,  C.  A.,  29,  35,  36,  37,  41.  175, 

240,  260,  288,  489,  503,  504,  593,  659, 

662,  988,  1004,  1007,  1045,  1059,  1068 
Ely,  J.  O.,  400,  1006 
Emerson,  G.  A.,  518,  531,  538,  1006 
Emerson,  P.  M.,  436,  1006 
Emery,   J.   F.,   834,   1006 
Emmelot,   P.,   130,   149,   150,   156,  1063 
Endahl,    B.    R.,   848,   1006 
Engel,  L.  L..  555,  713,  781,  1030,  1050 
Engel,   S.   L.,   632,   1057 
Engelhardt,  W.,   865,   1006 
England,  S.,  273,  675,  703,  706.  716,  718, 

804,  833,  1006,  1054 
Engle,  C.  G.,  978,  1015 
Ennor,  A.  H.,  467,  685,  704,  707,  710, 

833,  836,  845,  1006,  1010,  1014,  1039 
Eny,   D.   M.,   228,   1006 
Eppley,  R.  W.,  881,  908,  909,  912,  1006 
Erbland,   J.,    151,    1036 
Erdos,  F.  G.,  834,  1006 
Erf.  L.  A.,  217,  218,  1015 
Erlenmeyer,   H.,   518,   530,   1006 
Ernster,   L.,   17,    18,  33,  444,  445,   708, 

849.  865,  872,  989,  1006.  1030.  1033, 

1054 
Errera,   M.,   969.   1057 
Eskarous,  J.  K.,  973,  978,  1060 
Estabrook,  R.  W.,  395,  1035 


AUTHOR    INDEX 


1083 


Estes,  E.  H.,  Jr.,  399,  1030 

Estler,  C.  J.,  750,  881,  884,  885,  1000 

Estrada,  J.,  711,  839,  1061 

Eubank,  L.  L.,  670,  690,  1006 

Eusebi,  A.  J.,  518,  530,  533,  998,  1000 

Evang,   A.,   662,   663.   1046 

Evans,  E.  A.,  Jr.,  53,  63,  73,  74,  78.  91, 

93,   124,  226,  1007,  1056.  1063 
Evans,  H.  J..  33,  65,  850,  855,  999 
Evans,  J.  B..  78,  83,  1044 
Evans,  J.  I.,   11,  13,  1007 
Evert,  H.  E.,  773,  1007 
Ewers,  A.,  865.  912.  1030 
Eyer,  H.,  694.  1009 
Eyring,  H.,  5.  1007 
Eysenbach,   H.,  49,  1008 
Ezaki,  S.,  271,  417,  1007 


Fahey,  J.   L..  538.  1030 

Fahrlander,   H.,    157,   1007 

Fain,  J.  N.,  391,  1007 

Fairclough,  R.  A.,   3,  1007 

Fairhurst,   A.   S.,   1035 

Falcone,  A.  B.,  846,  851,  1053 

Falcone,  G.,  546,  972,  1007 

Falk,  J.  E.,  972.  973,  988 

Fanestil,  D.  D.,  869,  1007 

Fanshier,  D.  W.,  334,  1029 

'Fantes,  K.  H.,  422,  1019 

Farah,  A.,  204,  205.  214.  626,  689,  918. 

921,  922,  943.  944,  945,  957,  997,  1007, 

1011,  1032 
Farah,  A.  E.,  626,  917,  934,  935,  1007, 

1038 
Farkas.  G.  L..   169.   170,   195,  1007 
Farkas,  W.,  451,  1034 
Farnham,  A.  E.,  435,  1043 
Farrar,  W.  V.,  576,  1015 
Fasella,  P.,  788,  803,  827,  1007,  1061 
Fasold,   H.,   641,   1007 
Fastier.  F.  X..  363,  364.  365,  994,  1007 
Favarger,  P.,   157,  1007 
Favelukes,  G.,  708,  717.  778,  781,  838, 

1007 
Fawaz,  E.  X.,  41,  55,  56,  64,  95,  112,  128, 

129,  221,  236,  925,  927,  1007 


Fawaz,  G.,  41,  55,  56,  64,  95,  112,  128, 
129,  221,  236.  925,  927,  1007 

Fazekas.   J.   F.,   500,   990 

Featherstone,  R.  M.,  40,  596,  1004,  1028 

Feeney,  R.  E.,  745,  754,  760,  1034 

Feher,   0..   949,   1016 

Feigelson,  P.,  281,  282,  466,  477,  503, 
603,  1007,  1014 

Feigl,  F..   15,  1007 

Feinberg,  R.  H..  660,  693,  713,  1007 

Feinstein,  R.  X.,  696,  1007 

Feist,  E.,  146,  224,  234,  1007 

Feist,  F.,  617,  618,  1007 

Fekete,  J.,  460,  1015 

Feldberg,  W.,  949,  1007 

Feldman,  W..  848,  855,  1007 

Felenbok,  B..  842.  1007 

Feller,  D.  D.,  146,  224,  234,  1007 

Fellig,   J.,   461,  462,   712,   1008,  1070 

Fellman.  J.  H..  314.  1008 

Felsher.  R.  Z.,  687.  1013 

Felton,  S.  P.,  539,  543,  544,  1020 

Felts,  J.   M.,   613,  1008 

Feng,  J.   Y.,   853,   1016 

Fenn,  P.,  965,  1048 

Ferguson,  J.  H.,  456,  1012 

Fenley,  H.  X..  772,  791,  793,  795,  796, 
797,  1008 

Ferno,   0.,  461,  464,   1004,  1008 

Ferrari,  G.,  520,  526,  578,  1003,  1048 

Ferrari.  R.  A.,  379,  390,  449,  1008 

Ferreira,    R.,    739,   1008 

Festenstein,  G.  X.,  417,  429.  1008 

Fewson,  C.  A.,  547,  554,  1008 

Fewster,  J.   A.,  842,  1008 

Fex,  H.,  461,  464,  1004,  1008 

Field,   J.   B.,   268,   1056 

Fildes,  P..  971,  975,  1008 

Filho,  J.  B.  M.,  510,  552,  850,  1047 

Finamore,  F.  J.,  887,  997 

Findlay,  J.,  429,  1008 

Fine,  A..  19,  29,  1044 

Finean.  J.  B.,  950,  1038 

Fink.  J.,  210,  1021 

Finkelman,  F.,   704,   723,  939,   990 

Finkle,   B.  J.,  804,  1008 

Fischer,   A..  465,  1008 

Fischer,  E.  H.,  453,  887,  995,  1028 


1084 


AUTHOR    INDEX 


Fischer,  F.  G.,  49,  1008 

Fischer,  J.,  810,  1064 

Fischer,  P.,  697,  698,  699,  700,  1012 

Fishbein,  W.  N.,  433,  1008 

Fisher,   A.   L.,   40,   1004 

Fisher,  F.  M.,  Jr.,  839,   1008 

Fisher,  H.  F.,  293,  1008 

Fishgold,  J.  T.,  30,  50,  200,  1008 

Fishman,  R.  A..  401,  1008 

Fishman,    W.    H.,    63,    428,    442,    1008, 

1041,  1054 
Fiskin,  R.  D.,  913,  1050 
Fitzgerald,  R.  J.,  632.  1008 
Fitzpatrick,  J.  B.,  304,  529,  1031,  1062 
Fiume,  L.,  238,  1008 
Flaks,  J.   G.,  476,   1008 
Flamm,   W.   G.,  771,  772,  1008 
Flatow,   L.,   670,   1008 
Flavin,  M.,  64,  145,  224,  226,  234,  235, 

357,  1008 
Fleischer,  S.,  16,  1014 
Flesch,  P.,  767,  1008,  1037 
Flickinger,  R.  A.,  Jr.,  203,  1008 
Flood.  A.  E.,  195,  1026 
Florkin,  M.,  297,  1008 
Fluharty,  A.  L.,  408,  1008 
Fluri,   R.,  529,   1008 
Foa,  P.  P.,  394,  1038 
Fodor,  P.  J.,  595,  1008 
Foldeak,  S.,  503,  1037 
Folk,   J.    E.,   367,   1008 
Folkers,  K.,  538,  1019 
Foltz,  V.  D.,  974,  1033 
Fonnesu,  A.,  210,  1050 
Fonnum,  F.,  857,  1008 
Ford,  L.,  873,  883,  1021 
Forney,  R.  B.,  960,  1008 
Forsen,  S.,  619,  1009 
Forssman,  S.,  18,  48.  98,  104,  105,  213, 

215,  217,  218,  959,  1009 
Forster,  R.  P.,  205,  921,  1009 
Forsyth,  F.  R.,  196,  529,  1050 
Forti,  G.,  623,  1009,  1036 
Foss,  O.,  639,  1009 
Foster,  C.,  530,  1023 
Foster,  R.  J.,  349,  372,  373,  374,  1009 
Foulerton,  A.  G.  R.,  954,  1009 
Foulkes,  E.  C,  912,  1009 


Fountain,  J.  R.,  582.  1009 
Fourneau,  E.,  952,  1009 
Fournier,  P.,  225,  226,  1009 
Fonts,  J.  R.,  429,  1013 
Fraenkel-Conrat,  H.,  681,  741,  980,  1009, 

1054 
Francis,  A.  M.,  226,  1063 
Francis,  M.  J.  O.,  845,  1009 
Frank,   E.,   461,   1015 
Frank,   S.,   695,   1020 
Franke,  W.,  18,  32,  37,  40,  44,  664,  665, 

1009,  1028 
Frankel,  S.,  569,  1048 
Franshier,  D.  W.,  596,  1028 
Eraser,  D.,  877,   981,  1054 
Eraser,  D.  M.,  551,  1009 
Frear,   D.   S.,  597,   846,  1023 
Fredrickson,   D.   S.,  614,  1057 
Freeland,  M.   R.,  696,  1060 
Freedland,  R.  A.,  325,  1009 
Freedlander,  B.  L.,  202,  218,  1009 
Freeman,   M.,   581,  1043 
French,  D.,  421,  1038 
French,  F.  A.,  202,  218,  1009 
French,  R.  C,  591,  991 
French,  T.  V.,  333,  1009 
Freudenberg,   K.,   694,   1009 
Friberg,  L.,  959,  960,  1009 
Fridhandler,  L.,  179,  183,  391,  393,  574, 

1009 
Fridovich,   I.,   288,   451,   549,   555,   803, 

807,  1009,  1034,  1046 
Fried,  G.  H.,  58,  1009 
Friedemann,   T.    E.,   349,   1009 
Frieden,  C.,  508,  514,  1009,  1010 
Frieden,  E.,  325,  744,  778,  1010 
Friedenwald,  J.  S.,  444,  465,  991,  1034 
Friedland,  I.  M.,  505.  1004,  1010 
Friedman,  B.,  137,  141,  1065 
Friedman,   D.   L.,   228,   233,   235,   1037, 

1057 
Friedmann,  E.,  34,  35,  1039 
Friedmann,  H.  C,  470,  1010 
Frimmer,  M.,  548,  552,  553,  676,  793,  850, 

1010 
Frisell,  W.   R.,  60,   341,  346,   601,   704, 

706,  717,  772,  780,  1010 
Fritz,  C.  T.,   175,  1010 


AUTHOR    IXDEX 


1085 


Frohman.  C.  E.,  89,  520,  525,  530.  1010 

Fromageot,  C,  60,  357.  1010 

Fromm.  H.  J..  357.  376,  1010 

Frommel,  E..  835,  1010 

Fruton,  J.  S..  375,  1010 

Fu,  T.-H.,  792,  813,  1043 

Fuentes,   V.,   877,   1010 

Fujie.  Y.,  865,  1010 

Fujimoto,  D.,  554.  684,  1010,  1022 

Fukui,  T.,  887,  1016 

Fukunaga,  K.,  843,  1029 

Fuld,  M.,   71,    140,  995,   1010 

Fulton,  J.  D.,  168, 179,  559,  693,  882, 1010 

Furgiuele,   F.   P.,   978,   981,   1052 

Furst,   A.,  202,  218.  1009 

Furuya,   K.,  820,   867.   WHO 

Futterman,  S.,  581,  582,  1010 


Gabrio,  B.  W..  543,  553.  581.  582.  711, 
848.   851,   864.   9V3,   997,   1020.    1054 

Gaebler,  O.  H.,  494,  504,  992,  1010 

Gaffney,  P.  E.,  978,  1015 

Gaffney,  T.  J.,   710,  845,  1010 

Gage,   J.   C,   961,   1010 

Gajdos,   A.,   62,   1010 

Gale.  E.  F.,  660,  845.  1060 

Gale,  G.  R.,  430,  1010 

Galoyan,  A.  A.,  951,  956,  985,  1010 

Galoyan,  S.  A.,  951,  958,  959,  1010 

Galston,   A.   W.,   750,   1010 

Gamble.  J.  L.,  Jr.,  383,  722,  872,  909,  914, 
1010,  1031,  1052 

Gamble,  W.,  91,  1002 

Gamborg,  O.  L.,  599,  1011 

Gammon,  G.  D.,  573,  1011 

Gane,  R.,  5.,  1011 

Ganguly,  J.,  547,  554,  1035 

Garber,  N.,  249,  599,  615,  688,  999,  1011 

Garcia -Hernandez,   M.,   355.   1011 

Gardier,  R.  W.,  921,  1011 

Gardner,   E.   A.,  214,   1011 

Garen,  A.,  439,  1011 

Garfinkel.   D.,  849,  1011 

Garlid,  K.,  924,  1065 

Gause,  G.  F.,  981,  1011 

Gawehn,   K.,   695,  1064 


Gayer.  J.,   922.   1011 
Gehrig,  R.  F.,  353.  1042 
Geiger,  P.  J..  789,  819,  1048 
Geissler,   A.-W.,   695,   1064 
Gelfant,  S.,  200,  968,  1011 
Gellert,  M..  803,  807,  814,  1012 
Gelles,   E..  4,   1011 
Gellhorn,  A.,  546,  576,  577,  1011. 

1053 
Gemma,  F.  E.,  594,  1034 
Gemmill,  C.  L.,  772,  827,  828,  838, 

1011 
Geppert,  J.,  971,  1011 
Gerard,  R.  W.,  55,  177,  212,  1000, 
Gerarde,  H.  W.,   156.  1011 
Gerardin,   C,   983.   984,   1011 
Gergely.    J.,    475,    866.    868,    939. 

1004,  1011 
Gerhardt,  P.,  768,  1011 
Gerhart,  J.  C,  468,  480,  481,  816, 
Gerlach,  D.,  802,  1044 
Gershanovich.    V.    N.,    387,    1011 
Gershenfeld,  L.,  690,  1011 
Gassier,  U.,  945,  946,  1011 
Gest,  A.,  49,  294,  1011,  1044 
Geuther,   617,   619,  1011 
Gey,  K.  F.,  318,  1045 
Gey,  M,  K.,  878,  1018 
Geyer,   R.   P.,   114,   137,   138,   141, 

148,  150,  595,  1011,  1033,  1049 
Gezon,  H.  M.,  90,  1069 
Ghiretti,  F.,  54,  92,  105,  114,  174, 

228,  336,  709,  718,  840,  991,  1011, 
Ghiretti-Magaldi,  A.,  54,  114,  174, 
Ghosh,  A.  K.,  660,  662,  675,  692, 

773,  833,  1011 
Ghosh,  S.,  356,  1012 
Giartosio,  A.,  808,  827,  1061 
Gibbins,  L.  N.,  833,  1012 
Gibbs,  M.,  838,  892,  1012 
Gibbs,   R.,   602,   1068 
Gibson,  F..  321,  1012 
Gibson,  K.  D.,  751,  888,  1012 
Gibson,  Q.  H.,  757,  1012 
Gibson,  S.,  960,  993 
Giebel,  O.,  187,   188,  1012 
Giebisch,  G.,  923,  928,  936.  1012 
Gilbert,  D.   A.,   797,  803,  1012 


1047, 


876, 


1032 


940. 


1011 


142, 


227, 
1048 
1011 
772, 


1086 


AUTHOR    INDEX 


Gilfillan,  R.  F.,  224,  1012 

Gillespie,  J.  M.,   712,  1012 

Gillespie,  L.,  315,  1042 

Gillespie,   R.   E.,   710,  1030 

Gillette,  J.   R.,   851,   1012 

Gilman,  A.,  698,  699,  700,  1012,  1045 

Gilmour,   D.,   803,   807,   814,    816,    819, 

820,  821,  866,  869,  1012 
Gil  y  Gil,  C,  985,  1012 
Ginoza,  Y.  W.,  737,  973,  1050 
Ginsberg,  T.,  458,  459,  lOlH 
Ginsburg,   S.,   532,   1041 
Giovanelli,  J.,   145,  228,  231,  1012 
Girerd,  R.  J.,  333,  1012 
Giri,  K.  V.,  660,  839,  854,  1012,  1047 
Gitlin,  J.,  548,  551,   1024 
Giuditta,  A.,  33,  38,  49,  773,  850,  855, 

1012,  1064,  1065 
Gladner,  J.  A.,  367,  369,  370,  lOOS,  1041 
Glahn,  P.  E.,  268,  340,  104S 
Glaid,  A.  J.,  432,  433,  435,  436,  1015, 1042 
Glaser,  L.,  411,  1012 
Glasziou,  K.  T.,  834,  1012 
Glazko,   A.   J.,  456,   1012 
Glenn,  J.   L.,   17,  26,  996 
Glick,  M.  C,  26,  989 
Glimm,  E.,  259,  421,  1067 
Glock,   G.   E.,  503,   708,   711,   838,   839, 

1003,  1012 
Glynn,  I.  M.,  705,  864,  1012 
Gnuchev,  N.  V.,  359,  1026 
Goddard,  A.  E.,   195,  1001 
Godeaux,  J.,  938,  1012 
Godzeski,  C,  26,  36,  1012 
Goebel,  W.  F.,  657,  1012 
Goedde,  H.  W.,  432,  1020 
Goedkoop,   J.  A.,   4,   1012 
Gonnert,   R.,   977,   978,   1013 
Goerz,  R.  D.,  292,  1012 
Goffart,  M.,  697,  698,  699,  700,  1012 
Gold,  A.  H.,  641,  1012 
Goldberg,  G.,  206,  1012 
Goldberg,   L.,   63,   1012 
Goldberg,   L.   I.,   315,   1012 
Goldemberg,   S.   H.,  476,  1031 
Goldin,  A.,  400,  401,  496,  1024,  1030 
Goldinger,  J.  M.,  75,  79,  115,  174,  991, 

1057 


Goldman,  D.  S.,  60,  268,  596,  831,  832, 

843,  854,  991,  1010,  1013 
Goldschmidt,  E.  P.,  77,  237,  992,  1013 
Goldstein,  A.,  776,  777,  788,  836,  1013 
Goldstein,  L.,  205,  332,  550,  551,  1009, 

1013,  1062 
Goldstein,  M.,  320,  599,  1013 
Golstein,  M.  H.,  920,  921,  1013 
Goldstone,  A.,  334,  355,  857,  987,  1013 
Goldthwait,  D.  A.,  817,  820,  988,  1057 
Gollub,   E.   G.,  480,  1013 
Golub.  0.  J.,  728,  979,  980,  981,  997 
Gonda,  O.,  20,  29,  61,  80,  844,  864,  872, 

989,  1013 
Gonnard,  P.,  308,  1013,  1045 
Gonzales,  E.  L.,  687,  716,  775,  783,  810, 

853,   1058 
Gonzalez-Monteagudo,  O.,  574,  1046 
Goodban,  A.  E.,  15,  1043,  1057 
Gooder,   H.,   268,   324,   1013 
Goodland,  R.  L.,  524,  1052 
Goodman,   D.   S.,   835,   886,   887.   1003, 

1013 
Goodman,   I.,   429,   696,   948,   1013 
Goodwin,  M.   E.,  571,  573,  991 
Goodyer,  A.  V.  N.,  925,  1065 
Gopinathan,  K.  P.,  490,  847,  1013 
Gordon,  C.   N.,  467,  1053 
Gordon,  H.,  407,  684,  692,  833,  1018 
Gordon,  J.  J.,  699,   865,   1013 
Gore,   I.   Y.,   886,   1045 
Gore,  M.   B.   R.,  445,   1013 
Gorin,  G.,  671,  672,  673,  1025,  1031 
Gosselin,  L.,  611,  886,  989,  1045 
Goth,   A.,   401,   1013 
Gots,  J.  S.,  480,  530,  1013,  1062 
Gotto,   A.   M.,   602,   1013 
Gottlieb,  D.,   125,  133,  997,  1013,  1047 
Gould,   R.   G.,   886,   1045 
Gourevitch,   A.,   623,   1013 
Gouvea,  M.  A.,  868,  939.  940,  1011,  1036 
Govorov,   N.   P.,   948,   1013 
Grady,  J.  E.,  201,  577,  968,  1055 
Graff,   M.,   587,   1036 
Grafflin  A.  L.,  62,  85,  87,  1013 
Grana,  E.,  520,  522,  526,  1003 
Grand,   R.,   60,  357,  1010 
Grandjean,   E.,   956,   1061 


AUTHOR    INDEX 


1087 


Granick,  S.,  161,  162,  674,  859,  888, 
1013,  1037,  1062 

Grant,  B.  R.,  394,  1013 

Grant,  P.  T.,  560,  693,  842,  995,  1013 

Grant,  W.  M.,  687,  1013 

Grassetti,   D.   R.,   334.   596.   102S,   1029 

Grassmann,  W.,  368,  1013 

Graves,  D.  J.,  453,  709,  1028,  1030 

Gray,  C.  T.,  52,  187,  228,  229,  1013 

Gray,  N.  M.,   62,   85,   87,   1013 

Gray,  S.  J.,  687,  1013 

Grazi,  E.,  407,  412,  995.  1013 

Green,  A.   A.,   845,   1013 

Green,  D.  E.,  16,  60,  61,  62,  64,  146, 
237,  259,  338,  348,  407,  437,  509,  549, 
555,  599,  684,  692.  832,  833,  843,  994, 
1013.  1014.  1018,  1034,  1035,  1052, 
1058 

Green,  D.  M.,  333,  1012 

Green,   D.   W.,   755,   1014 

Green,  I.,  445,  1014 

Green,  J.  W.,  908,   1014 

Green,  L.  S.,  15,  1049 

Green,  N.  M.,  735,  797,  798,  1014 

Green,   S.,   428,   1008 

Greenawalt,  J.  W.,  548,  864,   1014 

Greenbaum,  L.  M.,  367,  1014 

Greenberg,  D.  M..  155.  156,  200,  336, 
351,  357,  507,  582,  660,  684,  685,  687, 
693,  706,  709,  712,  713,  717,  718,  787, 
841,  843,  847,  856,  1007,  1014,  1010, 
1024,  1026,  1037,  1041,  1042,  1044, 
1046,   1052,   1055 

Greengard,  O.,  603,  1014 

Greengard,  P.,  211,  266,  949,  999,  1014 

Greenland,  R.  A.,  467,  480,  1068 

Greenstein,  J.  P.,  664,  665,  666,  667,  1014 

Gregg,  D.  C,  297,  1014 

Gregg,  J.  R.,  199,  882,  964,  965,  1014, 
1043 

Gregolin,   C,   693,   845,   1014 

Gregory,   M.    E.,   475,   1033 

Greif,  R.  L.,  873,  923,  927,  429,  1014 

Greig,  M.  E.,  75,  91,  178,  185,  358,  1002, 
1006 

Grein,   L.,  810,  857,  1014 

Gremels,  H.,  879,  927,  1014 

Greville,  G.  D.,  128,  175,  177,  180,  183, 


187,    210,    552,    789,    793,    816,    819, 

866,   869,   998,  1014,  1038,   1061 
Grey,   E.  C,  2,  228,  1014 
Griffin,  D.  H.,  595,  1056 
Griffith,   G.   C,  960,   1014 
Griffith,    W.    H.,    348,    1014 
Griffiths,  D.  E.,  467,  707,  833,  848,  874, 

1014,  1017,  1039 
Griffiths,  M.,  816,  820,  821,  866,  869,  1012 
Grillo,  M.  A.,  712,  1031 
Grimm,  M.  R.,  304,  1028 
Grisolia,  S.,  413,  439,  1014,  1023 
Groschel-Stewart,  U.,  641,  1007 
Gromet-Elhanen,    Z.,    557,    1014 
Gros,   F.,  479,   1014 
Gross,  R.   E.,  259,  335,  1014 
Grossman,  L.,  269.  487,  1006,  1014,  1015 
Grossowicz,  X.,  269,  428,  504,  1015 
Grouge,  V.,  978.   1015 
Grove,W.  E.,  701,  702,  723,  724,  725, 1032 
Gruber,  C.  M.,  217,  218,  1015 
Gruber,   W.,    122,   814,   861,   1015 
Grunberg-Manago,  M.,  705,  830,  1049 
Grunert,  R.  R.,  751,  956,  1015 
Guarino,  A.  J.,  413,  1015 
Gubler,  C.  J.,  520,  521,  532,  533,  534, 1015 
Giinther,  G.,  61,  987 
Guerra,  F.,  210,  1021 
Guest,  J.  R.,  590,  1015 
Guggenheim,  K.,  495,  1015 
Guha.   S.   R.,  550,   841,   1015 
Gulick,  Z.  R.,  458,  459.  461,  462,  1018 
GuUand,  J.  M.,  576,  1015 
Gumbmann,  M.,  75,  80,  91,  121,  1015 
Gumnit,  R.,  573,  1011 
Gunja,  Z.  H.,  429,  843,  1015 
Gunsalus,  I.  C,  430,  1004 
Gurd,  F.  R.  N.,  649,  739,  759,  789,  853, 

862,   1015,   1034 
Gurd,  R.  S.,  924,  1026 
Gurin,  S.,  22,  27,  74,  77,  124.  148,  150, 

995,  1028,  1066 
Gurley,  H.,  926,  1018 
Gurtner,   H.   P.,   526,   1015 
Gutfreund.   H..  85,   105,   152,   375,  453, 

993,  1023 
Guthrie,  R.,  531,  990 
Gwin,  B.  A.,  15,  1044 


1088 


AUTHOR    INDEX 


H 

Haan,  J.,  1020 

Haarniann,  W.,  737,  753,  1015 

Haas,  E.,  547,  549, '^550,  1015 

Haas,  H.   T.   A.,  965,  1015 

Haas,  W.  J.,  657,  1015 

Haavaldsen,  R.,  857,  1008 

Haavik,  A.  G.,  848,  849,  1017 

Habeeb,  A.  F.  S.  A.,  762,  1015 

Hachisuka,   Y.,    195,   1015 

Hackett,  D.  P.,  120,  553,  848,  873,  878, 

881,  1015,  1041,  1053 
Hackney,  F.  M.  V.,  296,  1015 
Haddow,  A.,  225,  990 
Hafez,  E.S.E.,  179,  183,  1009 
Haft,  D.  E.,  883,  884,  1015 
Hagen,  U.,  956,   957,   1015 
Hager,  L.  P.,  453,  1053 
Hagihira,  H.,  265,  394,  1015 
Hagstrom,   B.   E.,  726,  1015 
Hahn,  H.,  950,  1015 
Hahn,  L.,  460,  461,  1015 
Hahn,   M.,   974,   1015 
Haining,  J.   L.,  887,  1015 
Hakala,  M.   T.,  432,  1015 
Halasz,   P.,   949,   1016 
Haider,  D.  K.,  530,  576,  999 
Halenz,  D.  R.,  853,  1016 
Haley,  T.  J.,  942.  948,  1016 
Hall,  A.  N.,  321,  1016 
Hall,  G.  A.,  Jr.,  3,  1016 
Halliday,  K.  A.,  694,  1051 
Halliday,  S.  L.,  504,  505,  1016 
Hallman,  N.,   105,   108,   110,  1016 
Halpern,   C,   195,   1038 
Halpern,  Y.  S.,  269,  498,  504,  1015 
Halvorson,  H.,  326,  351,  423,  552,  553, 

710,  772,   782,  848,  1004,  1005,  1016 
Halvorson,   H.    0.,   236,   237,   270,   354, 

674,  705,  831,  1040,  1042 
Hamilton,  H.  E.,  902,  905,  1053 
Hamilton,  P.  B.,  293,  294,  708,  837,  870, 

1053,  1054 
Hamilton,  R.  D.,  782,  992 
Hamilton-Miller,  J.  M.  T.,  598,  615,  1016 
Hammes,  G.  G.,  788,  803,  1007 
Hamolsky,  M.,  497,  1024 
Hampton,  A.,  466,  467,  481,  1016 


Hampton,  M.  M.,  696,  1007 

Hand,  W.  C.,  948,  1016 

Handler,    P.,    138,    144,    217,    218,    219, 

332,  451,  485,  488,  512,  549,  555,  772, 

783,  803,  807,  1009,  1016,  1023,  1027 

1034,  1046 
Handley,  C.  A.,  856,  925,  933,  1016,  1039 
Handschumacher,  R.  E.,  472,  478,  480, 

1016,  1044,  1049 
Hanly,  V.  F.,  22.  97,  171,  181,  182,  185, 

189,   190,  1016,  1061 
Hannoun,  C,  193,  1016 
Hanshoff,  G.,  707,  779,  780,  1047 
Hanson,  A.,  329, 1016 
Hanzlik,  P.  J.,  952,  1016 
Happold,  F.  C,  268,  324,  858,  1013,  1016 
Harada,  M.,  347.  1069 
Harary,  I.,  705,  1016 
Hard,  J.  A.,  400,  1006 
Hardy,  W.  G.,  201,  670,  992 
Harger,  R.  N.,  960,  1008 
Hargreaves,  A.  B.,  675,  835,  1016 
Harigaya,  S.,  428,  987 
Harkness,  D.  R.,  817,  1018 
Harlan,  W.  R.,  399,  1030 
Harley,  J.  L.,  51,  53,  169,  172,  1016 
Harman,  J.  W.,  119,  992 
Harper,  E.,  553,  1016 
Harpur,  R.  P.,  381,  593, 1016 
Harris,  J.,  668,  669,  713,  718,  1016 
Harris,  J.  E.,  911,  1016 
Harris,  J.  O.,  975,  1016 
Harris,  S.  A.,  518,  1067 
Harrison,  K.,  384,  1004 
Harting,  J.,  721,  876,  1016 
Hartman,  A.  M.,  2,  225,  1061 
Hartman,  S.  C,  333,  853,  1016 
Hartman,  W.  J.,  310,  1016 
Hartmann,  K.-U.,  476,  1016 
Hartree,  E.  F.,  259,  283,  696,  698,  1025 
Hartsell,  S.  E.,  168,  225,  1052 
Hartshorne,  D.,  713,  856,  1016 
Harvey,  G.  T.,  22,  29,  1016 
Harvey,  S.  C.,  173,  1017 
Hasenfuss,  M.,  125,  1029 
Hashimoto,  T.,  817,  1017 
Hashizume,  T.,  969,  1025 
Haskell,  T.  H.,  621,  998 


AUTHOR    INDEX 


1089 


Haskins,  F.  A.,  53,  1017 

Haslam,  R.  J.,  153,  1017 

Hass,  L.  F.,  390,  412,  853, 1017, 1025 

Hassall,  K.  A.,  414,  1017 

Hasse,  K.,  485,  487,  490,  1017 

Hasselbach,  W.,  938,  1017 

Hasselbring,  H.,  2,  225,  1041 

Hassid,  W.  Z.,  385.  991 

Hastings,  A.  B.,  869,  1007 

Hatch,  M.  D.,  53,  77,  79,  80,  82,  173, 
830,  853.  1017 

Hatefi,  Y.,  848,  849,  1017 

Hatt,  D.  L.,  980, 1047 

Haugaard,  N.,  178,  185,  659,  1017 

Hauge,  J.  G.,  591,  839,  1017 

Haughton,  B.  G.,  352,  578, 1017 

Haurowitz,  F.,  645,  1017 

Hauser,  A.  D.,  920,  921,  1013 

Hawtrey,  A.  O.,  79,  1017 

Hay,  A.  J.,  427,  429,  812,  814,  838,  1000, 
1032 

Hayaishi.  O..  226,  228,  230,  324,  853, 
1017 

Hayano,  M.,  116,  157,  1000 

Hayano,  S.,401,m5(? 

Hayashi,  T.,   802,  844,  1019 

Hayes,  A.  D.,  958,  959,  960,  1049 

Heald,  P.  J.,  134,  176,  184,  1017 

Hearon,  M.,  92,  991 

Hecht,  L.,  798,  865,  1042 

Heegaard,  E.,  516,  996 

Hegre,  C.  S.,  853,  1016 

Heidelberger,  C,  476,  1016 

Heilbrunn,  L.  V.,  963,  965,  1017 

Heim,  F.,  750,  881,  884,  885,  927,  1006, 
1031 

Heimann-Hollaender,  E.,  340,  1017 

Heimbiirger,  G.,  817,  989 

Heiney.  R.  E.,  639,  1027 

Heinrich,  W.-D.,  520,  1063 

Heinz,  F.,  817,  1017 

Heinz.  R.,  701,  724,  1017 

Heinzelman,  R.  V.,  358, 1052 

Hellerman,  L.,  38,  40,  42,  60,  61,  76, 
240,  329,  340,  341,  346,  347,  547,  549, 
556,  557,  558,  560,  640,  643,  657,  658, 
662,  664,  665,  666,  668,  669,  673,  676, 
683,  686,  697,  701,  702,  703,  704,  706, 


713,  717,  718,  742,  772,  780,  789,  803, 
816,  819,  876,  878,  950,  995,  998, 
999,  1003,  1010,  1011,  1016,  1017, 
1018,  1041,   1045,   1047,   1048,   1051 

Hellig,  H.,711,852,  iW<5 

Helmert,  E.,  685,  1036 

Helmreich,  E.,  263,  387,  392,  394,  1018 

Hemker,  H.  C,  547,  549,  550,  556,  558, 
1018 

Hemming,  H.  G.,  618,  632, 1034 

Henderson,  J.  F.,  481,  1018 

Henderson,  J.  H.  M.,  171,  1018 

Henderson,  L.  M.,  610,  709,  772,  782, 
843,  1003,  1038,  1057 

Henderson,  M.  J..  125,  1039 

Hendley.  D.  D.,  463,  1018 

Hendlin,  D.,  582,  589,  1018 

Hendricks,  S.  B.,  122,  1022 

Henle,  W.,  530,  1023 

Henning,  U.,  887,  987 

Hepler,  O.  E.,  924,  926,  1018,  1054 

Hepp,  P.,  931,  941,  1018 

Heppel,  L.  A.,  471,  473,  817,  988,  1018, 
1021 

Herbain,  M.,  518,  1063 

Herbert,   D.,   407,   684,   692,   833,   1018 

Herbert,  E.,  864,  1018 

Herbert,  M.,  77,  78,  1019 

Heredia,  C.  F.,  387,  388,  400,  547,  554, 
1018,    1056 

Herken,   H..  499,   965,   996,  1018,  1067 

Herman,   E.,   927,  947,  1043 

Herner,  B.,  348,  1018 

Herring,  P.  J.,  384,  1018 

Herriott,  R.  M.,  683,  688,  1018 

Herrmann,  H.,  465,  i005 

Herschberg,  A.  D.,  835,  1010 

Herz,  R.,  Jr.,  208,  911,  913,  1018 

Herzog,  R.  O.,  974, 1018 

Herzog,    U.,    750,    881,    884,    885,    1006 

Hess,  R.,  850,  1018 

Hess,  S.  M.,  316,  317,  318,  319,  320,  1018 

Hesselbach,  M.  L.,  128,  177,  1005 

Hested,  D.  M.,  208,  1063 

Hestrin,  S.,  421,  1018 

Heydeman,   M.  T.,  831,   832,   989,  1018 

Heymann,  H.,  297,  458,  459,  461,  462, 
1018 


1090 


AUTHOR    INDEX 


Heymans,  J.  F.,  1,  187,  218,  1018 

Heyndrickx,  A.,  747,  748,  1058 

Heyworth,  R.,  419,  420,  1064 

Hiai,  S.,  292,  293,  1018 

Hiatt,  A.  J.,  33,  166,  855,  1018 

Hiatt,  R.  B.,  696,  948,  1013 

Hicks,   S.  P.,   499,   504,   670,   724,   965, 

1018 
Hierholzer,  K.,  924,  1026 
Hietanen,  S.,  736,  739,  1018 
Hift,  H.,  569,  1054 

Higgins,  E.  S.,  554,  614,  814,  1018,  1066 
Hildebrandt,  A.  C,  197,  1019 
Hilden.  T.,  918,  923,  996 
Hill,  B.  R.,  772,  774,  845, 1019 
Hill,  R.  L.,  1019 
Hill,  R.  M.,  41,  177,  1020 
Hillman,  R.  S.  L.,  574,  575,  1022 
Hillmann,  G.,  910,  1051 
Hilmoe,  R.  J.,  817,  1018 
Hilton,  J.  L.,  588,  597,  1019,  1062 
Hilz,  H.,  693,  818,  1019 
Himms,  J.  M.,  363,  994 
Himwich,  H.  E.,  500,  990 
Hinman,  R.  L.,  326,  1016 
Hino,  S.,  292,  293,  1018 
Hirade,  J.,  802,  844,  873,  1019 
Hiramitsu,  S.,  179,  1059 
Hirano,  S.,  126,  127,  135,  153.  176,  1059, 

1061 
Hirashima,  K.,  968,  969,  1019 
Hird,  F.  J.  R.,  657,  1019 
Hirohata,  R.,  792,  813,  1043 
Hirschhorn,  L.,  724,  993 
Hitchings,   G.   H.,   283,  429,  584,   1006, 

1013,  1019,  1068 
Hitchcock,  P.,  212,  1019 
Hjort,  A.  M.,  956,  990 
Ho,  J.  Y.  C,  639, 1027 
Hoadley,  L.,  963,  964,  1019 
Hoare,  D.  S.,  593,  708,  1003 
Hoare,  J.  L.,  762,  1019 
Hoberman,  H.  D.,  61,  1019 
Hoch,   F.   L.,   743,   746,   780,   785,   789, 

806,  825,  831,  1019,  1055 
Hoch,  G.  E.,  293,  1019 
Hochachka,  P.  W.,  169,  1019 
Hochster,  R.  M.,   18,  30,  31,   132,  261, 


474,    511,    553,    555,    847,    878,    994, 

1019,  1057,  1063 

Hockenhull,  D.  J.  D.,  77,  78,  422,  1019 

Hofling,  E.,  4,  1019 

Hogberg,  B.,  461,  464,  1004,  1008 

Hohl,  R.,  122,  1015 

Hokfelt,  B.,  401,  1019 

Hofmann,  E.,  396,  1019 

Hofmann,  E.  C.  G.,  487,  489,  493,  1019 

Hofmann,   K.,   589,   989 

Hofstee,   B.   H.   J.,  286,  288,  289,   457, 

676,  718,  721,  1019,  1054 
Hogg,  J.  F.,  263,  389,  390,  391,  1041 
Hoggarth,  E.,  224,  236,  1003 
Hohnholz-Merz,  E.,  802,  1044 
Holden,  M.,  693,  712,  1019 
Holland,  J.,  71,  74,  79,  80,  81.  82,  173, 

203,  273,  274,  1019 
Holland,  J.  F.,  531,  581,  990,  1019 
Holland,  W.  C,  936,  989 
Hollander,    P.    B.,    214,    625,    896,    943, 

944,  945,  1065 
Holldorf,  A.,  430,  841,  1020 
Holliday.  W.  M.,  494,  992 
HoUis,  V.  W.,  Jr.,  449,  1034 
HoUocher,  I.  C,  Jr.,  18,  22,  1000,  1019 
Hollunger,   G.,   17,  998 
Holly,  F.  VV.,  538,  1019 
Holmberg,  C.  G.,  691,  693,  1019 
Holmes,  W.  L.,  473,  1020 
Holms,  W.  H.,  124,  1003 
Holt,   A.,   803,   1020 
Holf,  C.  v.,  518,  1066 
Holtkamp,  D.  E.,  41,  177,  1020 
Holtman,  D.  F.,  224,  1012 
Holton,  F.  A.,  32,  35,  61,  80,  83,  121,  122, 

1020,  1055 
Holton,  R.,  195,  1058 
Holtz,  P.,  359,  577,  1020 
Holz,  G.,  841,  993 
Holzer,  E.,  1020 

Holzer,  H.,  430,  432,  695,  817,  841,  852, 

1020 
Hommes,  F.  A.,   18,  1020 
Honda,  S.  I.,  23,  27,  33,  46,  47,  51,  170, 

1020 
Hopkins,  F.  G.,  18,  25,  34,  35,  41,  661, 

662,  664,  1020 


AUTHOR    INDEX 


1091 


Hopper,   S.,    334,    358,    810,   856,    1020, 

1052 
Horecker,    B.    L.,    412,    413,    471,    855, 

995,  1021,  1059 
Hori,  K.,  549,  554,  555,  837,  1020 
Horibata,  K.,  400,  1000 
Horio,  T.,  228,  229,  557,  880,  1020 
Horning,  M.  G.,  234,  1020 
Horowitz,   M.   G.,   639,   760,   767,   1020, 

1027 
Horwitt,  M.  K.,  456,  1020 
Horwitz,  L.,  881,  892,  1020 
Hoshino,  M.,  543,  1020 
Hoskins,  D.  D.,  783,  847,  1020 
Hosoya,  N.,  55,  78,  147,  149,  179,  228, 

233,  1020 
Hosoya,  T.,  686,  1020 
Hospelhorn,  V.  D.,  681,  703,  762.  1023 
Hostynova,  D.,  954,  957,  1005 
Hotchkiss,  R.  D.,  910,  1020 
Houck,  C.  R.,  888,  889,  890,  897,  1020 
Houck,  J.  C,  460,  461,  462,  1020 
Houlihan,  R.  K.,  37,  241,  1052 
Houwing,  C,  980,  1024 
Howard,  R.  L.,  849,  1026 
Howe,  W.  B.,  359,  1020 
Howell,  R.  S.,  574,  577,  1051 
Howes,  W.  v.,  61,  1034 
Hsieh,  K.,  927,  1053 
Huang,  H.  T.,  271,  1020 
Hubbard,  D.  M.,  953,  1064 
Hubbard,   J.    S.,    267,    613,    1005,    1020 
Hubbard,  N.,  888,  1029 
Hubbard,  R.  W.,  459,  993 
Hudson,  M.T.,  385,  391,  392,  1020,  1068 
Hudson,  P.  B.,  440,  441,  442,  711,  713, 

788,  839,  842,  1032,  1061 
Hudson,  P.  S.,  452,  1061 
Hiibscher,  G.,  284,  662,  713,  1000,  1035 
Hulsmann,  W.  C,  86,  547,  549,  550,  556, 

558,  1018,  1021 
Huennekens,  F.  M.,  539,  543,  544,  553, 

581,  582,  585,  711,  848,  851,  864,  993, 

997,  1020,  1043,  1051,  1054,  1066 
Huf,  E.  G.,  670,  690,  1006 
Huffaker,  R.  C,  225,  226,  1021 
Huggins,  C,  681,  703,  762,  1023 
Hughes,  A.  F.  W.,  199,  742,  1021 


Hughes,  C,  852,  1021 

Hughes,  D.  E.,  550,  845,  1009,  1021 

Hughes,  W.  J.,  Jr.,  757,  1021 

Hughes,  W.  L.,  Jr.,  681,  743,  744,  748, 

750,  753,  754,  755,  757,  758,  759,  761, 

862,  930,  1005,  1021 
Huisman,  I.  H.  J.,  755,  760,  1021 
Hulme,  A.  C,  120,  1023 
Hulpieu,  H.  R.,  520,  1067 
Hultquist,  G.,  401,  1019 
Hummel,  J.  P.,  464,  475,  1021,  1041 
Humphrey,  B.  A.,  28,  1021 
Humphrey,  G.  F.,  22,  28,  71,  74,  79,  80, 

81,  82,   173,   174,  203,  273,  274,  713, 

882,  884,  1019,  1021 
Humphreys,  S.  R.,  400,  401,  496,  1024, 

1030 
Humphreys,  T.  E.,  136,  1024 
Hundley,  J.  M.,  495,  1033 
Hunter,  A.,  335,  1021 
Hunter,  D.,  951,  952,  953,  954,  1021 
Hunter,  F.  E.,  Jr.,  210,  873,  883,  1021 
Hunter,  F.  R.,  22,  23,  54,  58,  1021 
Hunter,  G.  J.  E.,   164,  238,  1005,  1021 
Hunter,  N.  W.,  27,  28,  1021 
Hunter,  R.,  27,  1021 
Hunter,  W.  C,  985,  1021 
Hunter,  W.  R.,  882,  961,  1021 
Hunter,   W.   S.,    173,    183,   1063 
Hurd,  C.  D.,  619.  997 
Hurlbert,  R.  B.,  15,  28,  997 
Hurwitz,    A.,    210,    1021 
Hurwitz,   J.,    413,    471,    475,    477,    564, 

832,  855,  1021 
Hurwitz,  L.,  875,  1021 
Husa,  W.  J.,  196,  1033 
Huszak,  S.,  124,  138,  140,  177,  1021 
Hutchinson,  D.  J.,  582,  583,  996,  1009 
Huxley,  J.  S.,  964,  1021 
Hylin,  J.  W.,  33,  244, 1021 
Hynd,  A.,  384,  1018,  1021 


lacono,  J.,  574,  1040 
Ibsen,  K.  H.,  396,  397,  1021 
Ichihara,  A.,  857,  1021 


1092 


AUTHOR    INDEX 


Ichihara,  E.  A.,  857,  1021 

Ichihara,    K.,    64,    564,    578,    676,    858, 

1022,  1039,  1043,  1064 
Igarasi,  H.,  225,  1022 
Iliffe,  J.,  149,  1022 
Illingsworth,  B.,  561,  1022 
Imaizumi,  R.,  611,  612,  1042 
Imamoto,   F.,   27,   1022 
Imamura,  K.,  868,  1061 
Imshenetsky,  A.  A.,  983,  984,  985,  1022 
Inada,   A.,   194,  999 
Inagaki,  T.,  741,  1022 
Ingold,  C.  K.,  5,  1011 
Ingraham,  R.  C,  910,  1022 
Ingram,  V.  M.,  649,  755,  1014,  1022 
Inkson,  R.  H.  E.,  133,  1040 
Inoue.  F.,  578,  676,  858,  1064 
Irvin,  E.  M.,  556,  1022 
Irvin,  J.  L.,  556,  1022 
Isaka,   S.,   678,   727,   1022 
Ishervvood,    F.    A.,    64,    773,    774,   1002, 

1035 
Tshii,   K.,  983,  984,  987 
Ishikawa,  S.,  873,  1022 
Ishimoto,  M.,  551,  554,  555,  684,  1010. 

1022 
Ishishita,  Y.,  179,  1059 
Isles,  T.  E.,  663,  1022 
Itatani,  M.  K.,  175,  194,  1039 
Ito,  E.,  439,  1022 
Ito,  H.,  578,  676,  858,  1064 
Ito,  K.,  578,  676,  858,  1064 
Ives,  D.   H.,  481,  1022 
Ives,  D.  J.,  G.,  9,  1002 
Ivler,  D.,  228,  230,  231,  1067 
Iwainsky,  H.,   168,  228,  237,  504,  1040 
Iwasa,  K.,  27,  1022 
Iwatsubo,  M.,  547,  559,  1022 
Iwatsuka,  H.,  892,  1022 
Izaki,  K.,  349,  1038 
Izar,  G.,  920,  1022 

J 

Jackson,  D.  E.,  918,  944,  946,  947,  952, 

1022 
Jackson,  J.  F.,  510,  989 
Jackson,  L.  J.,  599,  1051 
Jackson,  P.  C.,  122,  1022 


Jackson,  W.  T.,  709,  1030 

Jacob,  H.  S.,  877,  904,  906,  908,  912, 

1022 
Jacob,  M.,  856,  873,  1022 
Jacobs,  E.  E.,  856,  873,  1022 
Jacobs,  F.  A.,  574,  575,  1022 
Jacobs,  G.  S.,  873,  927,  929,  1014 
Jacobs,  H.  I.,  212,  1022 
Jacobsen,  C.  F.,  646,  1032 
Jacobsohn,  K.,  279,  1022 
Jacobson,  K.  B.,  154,  503,  511,  586, 

1022 
Jacobson,  L.,  170,  209,  273,  274,  1043 
Jacoby,  G.  A.,  305,  1022 
Jacoby,  M.,  794,  1022 
Jacquez,  J.  A.,  266,  1022 
Jaattela,  A.  J.,  Q\\,  1022 
Jaenicke,  L.,  585,  1022 
Janisch,  W.,  923,  1067 
Jarnefett,  J.,  548.  1023 
Jagendorf,  A.  T.,  851,  874,  891,  892, 

989,  1022,  1024 
Jahn,  F.,  701,  723,  724,  725,  727,  1022 
Jakoby,  W.  B.,  60,  547,  550,  595,  1022, 

1023,  1041,  1069 
James,  W.  O.,  171,  1023 
Janacek.  K.,  912,  950,  1023,  1032 
Jandl,  J.  H.,  877,  904,  900,  908,  912, 

1022 
Jang,  R.,  411,  657,  989,  1023 
Jankelson,  O.  M.,  208,  1063 
Jann,  G.  J.,  33,  38, 1058 
Jansen,  B.  C.  P.,  519,  774,  775,  854, 

1043 
Jansen,  E.  F.,  657,  1023 
Jansz,  H.  S.,  561,  1022 
Jasmin,  R.,  586,  1023 
Jecsai,  G.,  804,-809,  1023 
Jedeikin,  L.  A.,  149,  151,  1023 
Jeffree,  G.  M.,  443,  464,  1023 
Jencks,  W.  P.,  887,  1023 
Jenerick,  H.  P.,  211,  1023 
Jenkins,  W.  T.,  64,  334,  1023 
Jenrette,  W.  V.,  664,  667,  1014 
Jensen,  E.  V.,  681,  703,  762,  1023 
Jensen,  K.  B.,  15,  225,  1023 
Jermstad,  A.,  15,  225,  1023 
Jerstad,  A.  C,  961,  1038 


AUTHOR    INDEX 


1093 


Jervis,  E.  L.,  265,  1023 

Jiracek,  V.,  389,  1027 

Joachimoglu,  G.,  875,  1023 

Jocelyn,  P.  C,  663,  1022 

Jodrey,  L.  H.,  54,  174,  1023 

Johnson,  C.  E.,  859,  1060 

Johnson,  F.  H.,  676,  1053 

Johnson,  G.  T.,  28,  1023 

Johnson,  J.  E.,  941,  943,  1049 

Johnson,  M.,  199.  997 

Johnson,  M.  A.,  597,  846, 1023 

Johnson,  M.  P.,  911,  1023 

Johnson,  O.  H.,  538,  1006 

Johnson,  R.,  944,  945,  1007 

Johnson,  W.,  586,  1023 

Johnson,  W.  A.,  55,  58,  98,  99,  104,  110, 

1028 
Johnson,  W.  J.,  504,  505,  601,  1023 
Johnston.  R.  L.,  942.  1023 
Johnstone.  J.  H.,  384.  1023 
Johnstone,  R.  M.,  338,  1023 
Jokhk,  W.  K.,  685,  709,  711,  1023 
Jolley,  R.  L.,   132,  1023 
Jonas,  R.  E.  E.,  137,  994 
Jones,  E.  A.,  85,  105,  152,  453,  1023 
Jones,  J.  D.,  120,  1023 
Jones,  J.  H.,  530. 1023 
Jones,  J.  R.,  882,  963,  1023 
Jones,  L.  O.,  576,  1011 
Jones,  M.,   156,  1011 
Jones,  O.  T.  G.,  267,  547,  553.  1023,  1041 
Jones,  R.  F.,  169,  1056 
Jordan,  H.  V.,  632,  1008 
Jorgenson,  C.  R.,  263,  403,  1067 
Josephson,  K.,  685,  1063 
Joshi,  J.  G.,  512,  783.  1023 
Joshi,  J.  v.,  15,  987 
Josten,  J.  J.,  547,  553,  847,  1047 
Jowett,  M.,  78,  87,   138,   144,   176,   177, 

238,  349,  613,  735,  875,  1023 
Joyce,  B.  K.,  413,  439,  1014,  1023 
Joyce,  C.  R.  B.,  901,  905,  1024 
Judah,  J.  D.,  81,  999 
Judis,  J.,  294,  1011 
Jukes,  T.  H.,  581,  1003 
Jung,  F.,  900,  901,  902,  906,  1024 
Junowicz-Kocholaty,  R.,  349,  1039 
Jurtshuk,  P.,  Jr.,  843,  849,  1017,  1052 


K 

Kappner,  W.,  982,  1024 

Kagawa,  Y.,  163,  888,  1024 

Kahana,  S.  E.,  407,  474,  1024 

Kahn,  J.  S.,  874,  1024 

Kaisch,  K.,  674,  816,  1028 

Kaiser,  C,  261,  1024 

Kaiser,  E.,  459, 1024, 1043 

Kalbe.  H..  225.  1060 

Kalckar,  H.  M.,  285,  286,  287,  288,  859, 
1024, 1037 

Kaldor,  G.,  548,  551,  556,  1024 

Kallen,  R.  G.,  147,  1024 

Kalmus.  A.,  285,  992 

Kalner,  H.  S..  383,  396,  1047 

Kalnitsky,  G.,  18,  26,  90,  650,  826,  855, 
991,  1024,  1047 

Kaltenbach,  J.  P.,  18,  26,  1024 

Kalyankar,  G.  D.,  15, 1024 

Kamen,  M.  D.,  293,  557,  875,  880,  996, 
1020, 1056 

Kameyama,  T.,  551,  555,  1022 

Kamin,   H.,   341,   817,   850,   1027,   1066 

Kamrin,  A.  A.,  573,  1011,   1024 

Kamrin,  R.  P.,  573,  1011,  1024 

Kanamori,  M.,  708,  1024 

Kandler,  O.,  163,  1024 

Kang,  H.  H.,  610,  772,  782,  843,  1003, 
1038 

Kann.  E.  E.,  189,  1024 

Kantarjian,  A.  D.,  952,  954,  1024 

Kaper,  J.  M.,  980,  1024 

Koplan,  C.,  979,  1024 

Kaplan,  E.  H.,  348,  1024 

Kaplan,  L.  A.,  505,  1004,  1010 

Kaplan,  N.  O.,  211,  434,  485,  487,  490, 
493,  496,  497,  498,  500,  503,  504, 
507,  508,  509,  511,  512,  784,  786,  807, 
808,  846,  849,  850,  851,  980,  981, 
988,  1003,  1004,  1014,  1015,  1022, 
1024,  1025,  1041,  1053,  1060,  1062, 
1064,  1069 

Kara,  J.,  474,  1055 

Karasek,  M.  A.,  706,  713,  718,  1024 

Karlsson,  J.  L.,  227,  228,  1024 

Karmen,  A.,  234.  1020 

Karnofsky,  D.  A.,  576,  1024 


1094 


AUTHOR    INDEX 


Karnovsky,  M.  L.,  435,  104S 

Karpeiskii,  M.  Y.,  359,  102G 

Karunaivatnam,    M.    C,    61,    424,    1024 

Kasamaki,  A.,  78,  79,  86,  349,  1024 

Kashiwabara,  E.,  618,  632,  1062 

Kashket,  E.  R.,  120,  1024 

Kashket,  S.,  487,  492,  988 

Kassanis,  B.,  976,  979,  1024 

Kassel,  B.,  665,  666,  667,  995 

Kasser,  I.  S.,  937,  1066 

Kassowitz,  H.,  952,  1044 

Katagiri,  H.,  543,  554,  1059 

Katagiri,  M.,  853,  1017 

Katchalski,  E.,  457,  758,  761,  1005,  1024 

Kato,   H.,   195,   1015 

Kato,  I.,  416,  417,  1025 

Kato,  S.,  399,  400,  1056 

Kato,  Y.,  542,  1059 

Katoh,  S.,  787,  1025 

Katoh,  T.,  554,  1025 

Katsuya,   H.,   969,   1025 

Katyal,  J.  M.,  671,  672,  673,  1025 

Katz,  A.  M.,  939,  940,  1025 

Katz,  H.,  664,  1028 

Katz,  J.,  145,  226,  234,  1025 

Katz,  R.  L.,  532,  1041 

Katz,  S.,  741,  980,  1025 

Katzman,  P.  A.,  839,  1004 

Kauder,  E.  M.,  841,  993 

Kaufman,  B.,  849,  851,  1025 

Kaufman,  B.  T.,  211,  816,  1025 

Kaufman,  S.,  369,  1025 

Kavanau,  J.  L.,  937,  1025 

Kawabata,  S.,  224,  1025 

Kawada,  N.,  55,  78,  147,  149,  179,  228, 

233,  1020 
Kawai,  F.,  675,  685,  692,  771,  772,  835, 

999 
Kawano,  Y.,  224,  1025 
Kawasaki,  E.   H.,  52,  1026 
Kay,  C.  M.,  759,  1025 
Kayser,  F.,  983,  984,  1011 
Kaziro,  Y.,  475,  853,  1025 
Kearney,  E.  B.,  16,  17,  18,  32,  33,  38, 

42,    45,   46,   48,    539,   541,    542,    545, 

547,    549,    713,    783,    825    856,    1025, 

1054 
Kearns,  C.  W.,  675,  1032 


Keay,  L.,  551,  709,  843,  1048 

Keech,  D.  B.,  676,  853, 1025 

Keelee,  M.  M.,  939,  999 

Keighley,  G.,  887,  995 

Keil,  J.  G.,  226,  228,  992 

Keilin,   D.,   22,   23,   259,  283,   696,   698, 

1025 
Keister,  D.  L.,  510,  891,  1025 
Kekwick,  R.  G.  O.,  836,  1066 
Keleti,  T.,  409,  1025 
Kellaway,  C.  H.,  949,  1007 
Keller,  D.  M.,  81,  210,  1025 
Keller,  V.,  529,  1051 
Kellerman,  G.  M.,  466,  1025 
Kelley,  J.  F.,  463,  993 
Kelly,  F.  J.,  928,  997 
Kelly,  S.,  169,  209,  1025 
Kelner,  A.,  349,  1039 
Kench,  J.  E.,  920,  1000 
Kennedy,  E.  P.,  137,  1031 
Kennedy,  J.,  348,  363,  1024,  1051 
Kenney,  F.  T.,  713,  857,  1025 
Kenten,  R.  H.,  524,  1025 
Keodara,  J.  C,  1025 
Kerby,  G.  P.,  459,  1025 
Kerly,  M.,  81,  176,  183,  997 
Kermack,    W.    0.,    551,    560,    593,   995, 

1009,  1025 
Kernan,  R.  P.,  212,  1026 
Kerr,  D.  S.,  817,  820,  988 
Keskin,  H.,  60,  996 
Kessler,  D.,  980,  1026 
Kessler,  R.  H.,  206,  918,  923,  924,  928, 

929,   930,   933,   935,   995,   1026,   1058 
Keston.  A.  S.,  413,  414,  1026 
Ketchel,  M.,  199,  1051 
Kharasch,  N.,  637,  639,  1026,  1043 
Khomutov,  R.  M.,  359,  1026 
Khorana,  H.  G.,  473,  1047 
Kielley,  R.  K.,  121,  445,  446,  1026 
Kielley,  W.  W.,  121,  445,  446,  807,816, 

866,  1026 
Kies,  M.  W.,  693,  1056 
Kiesow,  L.,  387,  395,  678,  1026 
Kiessling,  K.  H.,  515,  518,  1026 
Kihara,   H.,   833,   1026 
Kilbourne,  E.  D.,  400,  1026 
Kilsheimer,  G.  S.,  238,  440,  1026 


AUTHOR    INDEX 


1095 


Kimmel,  J.  R.,  375,  769,  770,  804,  1026, 

1055 
Kimura,  H.,  137,  1026 
Kimura,  T.,   845,  1026 
Kimura,  Y.,  124,  135,  176,  184,  1026 
King,  C.  G.,  259,  1040 
King,  E.  E.,  943,  944,  956,  957,  10:il 
King,    E.    J.,    440,    987 
King,  H.  K.,  352,  578,  780,  783,  810,  814, 

1017,  1059 
King,  R.  L.,  1053 
King,  T.  E.,  22,  33,  38,  552,  705,  712, 

832,  849,  853,  1002,  1025,  1026,  1033, 

1050 
Kini,  M.  M.,  76,  135,  153,  1026 
Kinoshita,  S.,  542,  1059 
Kinoshita,  T.,  969,  1025 
Kjaer,  A.,  617,  632,  996 
Kjeldgaard,  N.   D.,  285,  286,  287,  288, 

1024 
Kinsey,   V.    E.,   687,    1013 
Kinsky,  S.  C,  850,  1026 
Kipnis,  D.  M.,  387,  388,  390,  1026 
Kiraly,  Z.,  170,  1007 
Kirkham,  D.  S.,  195,  1026 
Kirkman,  H.  N.,  839,  1026 
Kirkpatrick,  H.  C,  979,  1032 
Kirshner,  N.,  399,  1030 
Kishimoto,  U.,  866,  1026 
Kistiakowsky,     G.     B.,     64.     603,     610, 

1026 
Kistner,  S.,  547,  1026 
Kit,  S.,  155,  200,  1026 
Kitagawa,  S.,  816,  866,  867,  868,  1061 
Kiyomoto,  A.,  428,  987 
Klavano,  P.  A.,  957,  959,  961,  1038 
Kleczkowski,    A.,    976,    979,    1024 
Klein,  A.  O.,  675,  1026 
Klein,  H.  P.,  269,  1006 
Klein,   J.   R.,    152,   341,   342,   347,   350, 

485,  488,  1016,  1027 
Klein,  M.,  978,  979,  980,  1027,  1044 
Klein  R.  F.,  399,  1030 
Klein,   R.  M.,   74,   171,   189,  595,   1032, 

1046 
Klein,  W.,  472,  1027 
Kleiner,  I.  S.,  202,  994 
Kleinfeld,  M.,  942,  945,  946,  1027,  1057 


Kleinspehn,  G.  G.,  591,  1061 
Kleinzeller,  A.,  267,  908,  1027 
Kleitman,  N.,  942,  944,  946,  1050 
Klemperer,   H.   G.,   688,   881,   910,   912, 

1027 
Klempien,  E.  J.,  693,  818,  1019 
Klenk,  L.,  368,  1013 
Klenow,    H.,   285,   286,   287,   288,    1024 
Kleppe,  K.,  389,  1044 
Klimek,  J.  W.,  973,  983,  1027 
Kline,  D.  L.,  150,  151,  1027 
Klingenberg,  M.,  Gl.,  178,  1027 
Klingman,  J.  D.,  332,  1027 
Klotz,   I.  M.,  29,  32,  35,  37,   242,  243, 

639,  757,  760,  767,  1020,  1027,  1049, 

1060 
Klotz,  T.  A.,  757,  1027 
Klybas,  V.,  408,  409,  1046 
Knell,  J.,  314,  315,  1066 
Knight,  S.  G.,  708,  837,  870,  1054 
Knobloch,  N.,  197,  1067 
Knopfmacher,  H.  P.,  683,  685,  691,  692, 

1027 
Knox,   R.,  615,  697,  699,  1016,  1045 
Knox,  W.  E.,  686,  1056 
Kobashi,  Y.,  225,  362,  1027,  1064 
Koch,  R.,  970,  1027 
Koch,  W.,  352,  1066 
Kochakian,  C.  D.,  848,  1006 
Kocholaty,  W.,  349,  687,  1027,  1039 
Kocouvek,  J.,  389,  1027 
Kodicek,  E.,  886,  1060 
Koditschek,  L.  K.,  582,  1018 
Koedam,  J.  C,  521,  522,  525,  1027 
Koelle,  E.  S.,  698,  699,  700,  1012,  1045 
Konig,  S.,  396,  1019 
Koeppe,  O.  J.,  812,  1027 
Koffler,  H.,  77,  237,  992,  1013 
Kohler,  A.,  291,  1041 
Kohler,  A.  R.,  583,  996 
Kohn,  P.,  376,  388,  1030 
Koike,  K.,  817,  1027 
Koike,  M.,  179,  578,  1035,  1059 
Koishi,  T.,  204,  883,  921,  1027 
Koisumi,  T.,  578,  1064 
Koivusalo,  M.,  349,  1027 
Koizumi,  T.,  676,  858,  1064 
Kolesar,  P.,  954,  957,  1005 


1096 


AUTHOR    INDEX 


Kolthoff.  I.  M.,  671.  747,  748,  763,  1027, 

1058 
Kominz,    D.    R.,    789,    866,    868,    939, 

1027 
Kondo,  K.,  675,  685,  692,  771,  772,  835, 

999 
Kondo,  S.,  439,  1022 
Kondo,  Y.,  551,  555,  1022 
Kono,  M.,  382,  1027 
Kono,  T.,  896,  912,  1027 
Konrad,  fi.,  170,  1007 
Konsanszky,  A.,  503,  1037 
Kopac,  M.  J.,  201,  998 
Korey,  S.,  696,  723,  938,  1027 
Korey,  S.  R.,  705,  830,  1049 
Korkes,  S.,  410,  1058 
Korn,  E.  D.,  463,  1027 
Kornberg,  A.,  226,  228,  446,  470,  1017, 

1042,  1070 
Kornberg,  H.  L.,  71,  594,  602,  845.  1004, 

1009,  1013,  1027 
Kornguth,  S.  E.,  456,  1027 
Koshland,  D.  E.,  Jr.,  868,  1032 
Kostif,  J.,  389,  1027 
Kostytschew,  S.,  875,  1028 
Kosunen,  T.,   109,  1054 
Kotaka,  S.,  541,  1028 
Kotyk,   A.,  267,   1027 
Koukol,  J.,  355,  852, 1028 
Kovac,  L.,  49,  267,  1027,  1028 
Kovachevich,  R.,  885,  1028 
Koval,  G.  J.,  601,  856,  987 
Kowa,  Y.,  148,  228,  1029 
Krahe,  E.,  974,  1028 
Krahl,  M.  E.,  198,  1028 
Krajci-Lazary,   B.,   926,   954,   957,   1005 
Kramer,  M.,  236,  1028 
Kramer,  P.  J.,  209,  1028 
Kramer,  S.,  29,  1034 
Krampitz,  L.  O.,  26,  89,  168,  1051,  1059 
Krane,  S.  M.,  262,  1028 
Krasna,  A.  I.,  279,  293,  294,  676,  1008, 

1028 
Krause,  R.  M.,  980,  1026 
Kravitz,  E.,  74,  77,  1057 
Kream,  J.,  288,  1053 
Krebs,  E.  G.,  289,  453,  1028 
Krebs,  H.  A.,  22,  27,  55,  58,  70,  71,  72, 


74,   75,   77,   89,   94,   96,   98,   99,    104, 

110,  115,  116,  124,  153,  158,  332,  333, 

769,  830,  1017,  1027,  1028,  1060 
Krejci,  L.  E.,  687,  1027 
Kreke,  C.  W.,  744,  810,  835,  856,  870, 

871,  880, 1001,  1028,  1052,  1055,  1056 
Kretszchmar,  R.,  513,  1062 
Krishna  Murti,  C.  R.,  692,  831,  987 
Krishnaswamy,    P.    R.,     15,    660,    854, 

1012,  1024 
Kriszat,  G.,  726,  964,  1028,  1049 
Kroger,  M.  H.,  856,  870,  871,  1028 
Kronenberg,  G.  H.  M.,  891,  1043 
Krop,  S.,  956,  957,  1038 
Krueger,  A.  P.,  976,  979,  980,  1028 
Krueger,  R.,  876,  938,  1028 
Krueger,  R.  C,  300,  1028 
Kruse,  R.,  922,  1007 
Kruse,  W.  T.,  620,  622,  623,  624,  625, 

627,  628,  630,  631,  1052,  1068 
Krusius.  F.  E.,  109,  110,  1028 
Kubowitz,  F.,  768,  1028 
Kuby,  S.  A.,  60,  836,  1028 
Kuczynski,  M.,  878,  989 
Kuhn,  J.,  791,  794,  817,  820,  1000 
Kuhn,  R.,  259,  421,  664,  665,  666,  668, 

687,  712,  853,  1028 
Kulka,  R.  G.,  444,  445,  1001 
Kull,  F.  C,  304,  1028 
Kumar,  S.  A.,  660,  685,  1047 
Kumin,  S.,  160,  1053 
Kummerovv,  F.  A.,  137,  1026 
Kun,  E.,  334,  355,  593,  596,  767,  1008, 

1011,  1028,  1029 
Kuner,  E.,  945,  946,  1011 
Kuno,  M.,  892,  1022 
Kuno,  T.,  195,  1015 
Kunz,  H.  A.,  519,  531,  1029 
Kunz,  W.,  40,  1029 
Kuo,  M.  H.,  440,  443,  1029 
Kupiecki,  F.  P.,  706,  834,  1029 
Kuratomi,  K.,  519,  676,  774,  783,  843, 

854,  1029 
Kurtz,  A.  N.,  373,  1029,  1064 
Kurup,  C.  K.  R.,  615,  1029 
Kuschinsky,    G.,    937,    938,    939,    1029, 

1061 


AUTHOR    INDEX 


1097 


Kusunose,  E.,  26,  36,  62,  63,   148,  228, 

855,  1029,  1069 
Kusunose,  M.,  26,  36,  62,  63,  148,  228, 

855,  1029.  1069 
Kutscha,  W.,  938,  1029 
Kutscher,  W.,  125,  1029 
Kuttner,  E.,  298,  299,  301,  1029 
Kvam,  D.  C,  478,  1029 
Kvamme,  E.,  104,  111,  126,  1029 
Kwientny,  H.,  281,  992 
Kwienthy-Govrin,  H..  285,  992,  993 


Labbe,  R.  F.,  888,  1029 

Labeyrie,  F.,  435,  437,  547,  559,  1022, 

1029 
Lachnit,  V.,  1045 
Lack,  L.,  160,  1068 
Ladd,  J.  N.,  15,  52,  1029 
La  Du,  B.  N.,   272,  305,  306,  595,  851, 

1012,  1022,  1029,  1069 
Ladygina,  M.  E.,  170,  1049 
Lagnado,  J.  R.,  816,  1029 
Laki,  K.,  63,  1029 
Lakshmanan,  M.  R.,  547,  549,  837, 

1029 
Laland,  S.  G.,  384,  1005 
Lamberg,  S.  L.,  540,  544,  993 
Lambie,  A.  T.,  921,  1029 
Lambooy,  J.  P.,  537,  538,  539,  988, 

1029,  1051 
Lamprecht,  W.,  817, 1017 
Landau,  B.,  392,  400,  401,  1030 
Landau,  B.  R.,  263,  387,  388,  401,  403, 

404,  1029,  1067 
Landau,  J.  V.,  964,  965,  1029,  1070 
Lands,  W.  E.  M.,  372,  1029 
Landon,  E.  J.,  282,  864,  1029 
Lane,  M.,  538,  1029,  1030 
Lane,  M.  D.,  853,  1016 
Lang,  C.  A.,  851,  1030 
Lang,  H.  M.,  891,  1050 
Lang,  K.,  601,  1030 
Langdon,  R.  G.,  412,  781,  815,  817, 

839,  841,  999,  1030,  1038,  1045 
Lange,  C.  F.,  Jr.,  376,  388,  1030 
Lange,  R.,  662,  663,  697,  699,  1045 


Langer,  L.  J.,  781,  1030 

Lansford,  E.  M.,  Jr.,  354,  1001 

Lara,  F.  J.  S.,  16,  49,  225,  226,  435,  547, 

551,  557,  558,  710,  717,  773,  778,  781, 

844,  1030,  1037,  1038,  1054 
Lardy,  H.  A.,  60,  63,  77,  87,  119,  124, 

128,  176,  203,  377,  381,  382,  383,  475, 

710,  779,  783,  798,  836,  846,  869, 

1001,   1028,   1030,   1035,   1042,   1044, 

1045,  1050 
Laris,  P.  C.,  865,  912,  1030 
Larner,  J.,  391,  706,  709,  710,  718,  1030, 

1049 
Laroche,  M.  J.,  592,  611,  989 
Larrabee,M.G.,  211, 1050 
Larson,  E.  R.,  404, 1066 
Larson,  F.  C.,  602,  1030 
La  Sala,  E.  F.,  622,  991 
Lascelles,  J.,  163,  888,  1030 
Lassen,  U.  V.,  267,  1030 
Lasser,  N.  L.,  38,  40,  42,  240,  1018 
Laszlo,    J.,    388,    392,    399,    400,    401, 

1029,  1030 
Lathe,  G.  H.,  428,  1030 
Laties,   G.   G.,   78,   79,  87,   91,   94,   132, 

170,  171,  173,  181,  182,  183,  185,  189, 

616,  1030,  1044 
Latuasan,  H.  E.,  538,  1030 
Latzkovits,  L.,  179,  1004 
Launoy,  L.,  984,  985,  1030 
Laurence,  E.  B.,  200,  997 
Laver,  W.  G.,  751,  888,  1012 
Lavik,  P.  S.,  856,  925,  1016 
Lawrence,  J.  C,  388,  389,  391,  996,  998 
Lazarow,  A.,  175,  1000,  1010 
Lazarus,  A.  S.,  968,  972,  1050 
Lazdunski,  M.,  711,  772,  773,  1030 
Lazzarini,  R.  A.,  512,  553,  851,  1030 
Lea,  D.  J.,  321,  1016 
Leaback,  D.  H.,  419, 1046 
Leach,  F.  R.,  610,  782,  843,  1003 
Leach,  S.  J.,  763,  1030 
Le  Bras,  G.,  356,  1057 
Lebrun,  J.,  203,  1030 
Leder,  I.  G.,  781,  839,  1030 
Lederer,  E.,  683,  1040 
Ledingham,  G.  A.,  169,  195,  1007 
Ledoux,  L.,  712,  715,  716,  744,  815,  1030 


1098 


AUTHOR    INDEX 


Lee,  C.  P.,  18,  33,  1030 

Lee,  H.  A.,  Jr.,  590, 1030 

Lee,  J.  S.,  228,  1031 

Lee,  Y.-P.,  674,  697,  774,  804,  852,  1031 

Lees,   H.,  450  451,  547,   551,  988,  997, 

1031 
Le  Fevre,  C.  G.,  618,  1031 
Le  Fevre,  P.  G.,  264,  690,  905,  906,  911, 

912,  1031 
Le  Fevre,  R.  J.  W.,  618,  1031 
Legge,  J.  W.,  707,  1031 
Lehman,  I.  R.,  462,  849,  1031 
Lehman,    R.    A.,    943,    944,    956,    957, 

1003,  1031 
Lehnann,  J.,  1031 
Lehmnger,  A.  L.,  91,  119,  137,  138,  178, 

705,  872,  873,  874,  1001,  1022,  1031, 

1064 
Leiby,  C.  M.,  316,  317,  318,  1045 
Leigh,  J.,  63,  1012 
Lein,  J.,  623,  1013 
Leloir,   L.   F.,   61,   144,   476,   594,   1005, 

1014,  1031 
Lembeck,  F.,  611,  1031 
Lenney,  J.  F.,  791,  793,  1031 
Lenti,   C,   712,   1031 
Lentz,  C.  P.,  677, 1042 
Lenz,  G.  R.,  739,  1031 
Lenzi,  F.,  215,  1031 
Leopold,  A.  C,  881,  967,  1041 
Le  Page,  G.  A.,  168,  1063 
Leppla,  W.,  518,  1066 
Lerner,  A.  B.,  304,  1031 
Leslie,  J.,  672,  1031 
Lester,  G.,  418,  1031 
Letnansky,  K.,  125,  399,  1031.  1052 
Leuschner,  F.,  750,  927,  1031 
Leuthardt,  F.,  157,  158,  1007,  1040 
Levaditi,  C,  984,  985,  1030 
Levenberg,  B.,  333,  1031 
Levey,  H.  A.,  31,  50,  176, 1031 
Levey,  S.,  458,  1031,  1036 
Levi,  A.  A.,  224,  995 
Levin,  G.,  281,992 
Levine,  H.  B.,  168,  1011 
Levine,  M.,  189,  1044 
Levine,  S.,  705,  714,  717,  718,  803,  805, 

831,  991 


Levinthal,C.,  439,  1011 

Levitt,  M.  F.,  920,  921,  1013 

Levvy,   G.   A.,   61,   272,   424,   426,   427, 

428,    429,    1000,    1008,    1024,    1031, 

1032,  1036 
Levy,  H.  M.,  868, 1032 
Levy,  H.  R.,  449,  712,  713,  1032 
Levy,  J.  F.,  210,  1021 
Levy,   R.   I.,   932,   933,   934,   935,   1032, 

1065 
Lewin,  J.,  758,  1032 
Lewis,  G.,  954,  994 
Lewis,  H.  B.,  236,  999 
Lewis,  S.  E.,  80,  85,  121,  1032 
Li,  C.  H.,  681,  1032 
Li,  S.  C,  838,  1032 
Li,  T.-K..  785,  1032 
Li,  Y.,  838,  1032 

Lichtenstein,  N.,  336,  340,  1017,  1032 
Lieb,  H.,  4,  1019 
Liebecq,  C,  63,  1032 
Lieben,  F.,  657,  658,  1032 
Lieber,  E.,  362,  1032 
Lieberman,  I.,  467,  1032 
Lieberman,  M.,  120,  1032 
Liener,  I.  E.,  770,  804,  1032 
Lifson,  N.,  228,  232,  1031,  1032 
Lilly,  J.  H.,  29,  1000 
Lin,  E.  C.  C,  263,  265,  394,  403,  1015, 

1067 
Lind,  C.  J.,  292,  1067 
Lindberg,  O.,  444,  445,  865,  872,  1033, 

1054 
Lindell,  S.  E.,  363,  1032 
Linderholm,  H.,  882,  912,  1032 
Linderdt,  T.,  461,  464,  1004,  1008 
Linderstrom-Lang,  K.,  646,  1032 
Lindner,  R.  C,  979,  1032 
Lindsay,    A.,    341,    347,    547,   549,    556, 

557,  558,  560,  780,  995,  1017 
Lindsten,   T.,    105,   213,   215,   217,   218, 

1009 
Lindstrom,   E.   S.,   26,   293,   1032,   1068 
Lineweaver,    H.,    227,    228,    667,    668, 

990,  1032 
Ling,  G.,  212,  1032 
Link,  G.  K.  K.,  74,  171,  189,  1032 
Linskens,  H.  F.,  967,  1044 


AUTHOR    INDEX 


1099 


Linstrom,  0.,  959,  960.  1059 

Lipke,  H.,  675,  70-32 

Lipmann,  F.,  326,  556,  887,  1023,  1032, 

1053 
Lipsett,  M.  N.,  741,  1032 
Lipten,  M.  A.,  75,  79,  115,  1057 
Lipton,  M.  M.,  385.  1061 
Lister,  A.  J.,  603,  610,  1032 
Litchfield,  J.  H.,  53,  74,  77,  79,  1032 
Lito,  E.,  975,  1033 
Littlefield,  J.  W.,  844,  1050 
Liverman,  J.  L.,  451,  1046 
Livemiore,  A.  H.,  518,  524,  1032,  1052 
Ljunggren,  M.,  708,  849,  1006 
Llinas,  J.  M.,  708,  837,  1045 
Lloyd,  D.  R.,  9,  1032 
Lockwoad,  W.  H.,  888,  1032 
Lodin,  Z.,  950,  1032 
Loevenhart,  A.  S.,  956,  1064 
Loblich,  H.  J.,  924,  1051 
Loefer,  J.  B.,  576, 1032 
Loevenhart,   A.   S.,   701,   707,   723,   724, 

725,  1032 
Levtrup,  S.,  76,  81,  152,  1033 
Low,  H.,  444,  445,  547,  548,  549,  551, 

555,  556,  864,  865,  872, 1014, 1033, 1054 
Logemann,  W.,  686,  693,  993 
London,  M.,  440,  441,  442,  1032 
Long,  W.  K.,  957,  1032 
Longley,  J.  B.,  926,  989 
Loomis,  W.  D.,  61,  1058 
Loomis,  W.  F.,  556, 1032 
Long,  J.  P.,  40,  1004 
Long,  M.  v.,  15,  1044 
Long,  W.  K.,  943,  1032 
Loofbourovv,  J.  R.,  657,  1015 
Lorand,  L.,  375,  1033 
Lord,  C.  F.,  196,  1033 
Lorenz,  B.,  461,  1033 
Lorenz,  R.,  461,  1033 
Loring,  J.  M.,  393,  399,  1063 
Lotspeich,  W.  D.,  81,  207,  210,  709,  844, 

1025,  1033,  1063 
Loureiro,  J.  A.,  de,  975,  1033 
Louwrier,  K.  P.,  409,  1043 
Loveless,  L.  R.,  195,  727,  972,  973,  1033 
Lovenberg,  W.,  310,  1033 
Low,  B.  W.,  757,  758,  1033 


Lowe,  H.  J.,  60,  341,  346,  1010 
Lowenstein,    J.    M.,    70,    72,    89,    147, 

1024,  1028,  1056 
Lowery,  D.  L.,  599,  1051 
LowTy,   O.   H.,   89,   286,   287,   288,   289, 

385,  407.   474.   540,   997,   1024,   1033, 

1044 
Lozano,  R.,  918,  923,  924,  929,  930,  935, 

1026 
Lucas,  D.  R.,  953,  1056 
Ludden,  C.  T.,  315,  1057 
Ludowieg,  J.,  385,  1004 
Luduena,  F.  P..  901.  1038 
Ludwig,  B.  J.,  297.  1033 
Ludwig.    G.    D..   544.    1060 
Luebering,  J.,  817,  818,  820,  1047 
Lubke,  M.,  891,  1002 
Liick,  H.,  623,  1033 
Liillmann,  H.,  937,  938.  1029 
Luh.  B.  S.,  421,  1033 
Lukomskaia,  I.  S.,  417,  1033 
Lumnis.  W.  L..  201,  577,  968,  1055 
Lumry,  R.,  60,  365,  1055 
Lund,  P.,  594,  1004 
Lundbom,  S.,  293,  1039 
Lundgren,  K.-D.,  959,  960,  1059 
Lusty,  C.  J.,  46,  1054 
Luteraan,  P.  J.,  195,  1033 
Lutwak-Mann,  C,  18,  25,  34.  35,  41,  179, 

661,  662,  664,  1020,  1033 
Luukkainen,  T..  349,  1027 
Luzzati,   M.,    552,    708,    994 
Luzzato,  R.,  726,  1033 
Lwoff,  A.,  62,  63,  81,  82,  1033 
Lyman,  C.  M.,  666,  667,  991 
Lynch,  J.  L.,  359,  360,  1041 
Lynen,  F.,  27,  77,  92,  187,  614,  751,  887, 

987,  995,  1033 
Lyon,  I.,  148,  150,  1033 


M 


Mc  AUan,  A.,  429,  1032 

Mc  Calla,  T.  M.,  974, 1033 

Mc  Clure,  F.  J.,  630,  631, 1070 

Mc  Coll,  J.  D.,  504,  505,  1023 

Mc  CoUister,  D.  D.,  627,  1056 

Mc  Comb,  R.  B.,  381,  382,  388,  395,  1033 


1100 


AUTHOR    INDEX 


Mc  Cormack,  B.  R.  S.,  175,  184,  1039 
Mc  Cormick,  D.  B.,  358,  475,  539,  542, 

564,  565,  578,  1033 
Mc  Cormick,  N.  G.,  675,  1033 
Mc  Cornack,  B.,  910,  913,  1050 
Mc  Crea,  F.  D.,  944,  945,  1033 
Mc  Curdy,  M.  D.,  Jr.,  169,  1033 
Mc  Daniel,  E.  G.,  495,  1033 
Mc  Devitt,  M.,  810,  835,  870,  1001 
Macdonald,  K.,  63,  439,  711,  778,  1034 
Mc  Donald,  J.  K.,  38,  1033 
Mac  Donnell,  L.  R.,  745,  754,  760,  1034 
McDougall,  B.  M.,  581,  994 
Macdowall,  F.  D.  H.,  891,  1034 
Mc  Elroy,  W.  D.,  845,  850,  1013,  1026 
Mc  Ewen,  B.  S.,  32,  189,  393,  622,  624, 

1034 
Mc  Fadden.  B.  A.,  61,  892,  1034 
Macfarlane,  E.  W.  E.,  964,  966,  1034 
Macfarlane,  W.  V.,  938,  1034 
Mc  Garrahan,  J.  F.,  382,  1034 
Mac  Gillavry,  C.  H.,  4,  1012 
Mc  Gilvery,  R.  W.,  1038 
Mc  Gowan,  J.  C,  618,  632,  1034 
Mc  Grath,  H.,  351,  1046 
Mc  Guire,  J.  S.,  Jr.,  449,  1034 
Machado,  A.  L.,  685,  1040 
Macheboeuf,  M.,  15,  939,  999 
Mc  Henry,  E.  W.,  556,  569,  573,  991,  997 
Machlis,    L.,    27,    116,    168,    170,    185, 

994,  1034 
Macht,  D.  I.,  944,  956,  957,  966,  1034 
Mc  Hugh,  R.,  440,  441,  442,  1032 
Mcllwain,   H.,   350,  485,  491.  497,  504, 

587,  637,  1034 
Mackay,  D.,  310,  1034 
Mc  Kee,  R.  W.,  396,  397,  1021 
Mackenzie,  C.  G.,  601,  1010 
Mc  Kinney,  G.  R.,  434,  594,  1034 
Mac  Kinnon,  J.  A.,  376,  381,  382,   383, 

842,  997 
Mackler,    B.,    549,    552,    553,    555,    783, 

832,  847,  848,  849,  1020,  1034,  1035, 

1044 
Mc  Lean,  P.,  708,  711,  838,  839,  1012 
Mac  Lean,  P.  D.,  499,  1000 
Mac  Lennan,  D.  H.,  274,  1034 
Mac  Leod,  J.,  699,  1034 


MacLeod,  R.  M.,  451,  452,  1034 

Mc  Mahon,  P.,  173,  882,  987 

Mc  Mahon,  R.  E.,  592,  1034 

Mc  Manus,  T.  T.,  149,  232,  887,  1039 

Mac  Nider,  W.  de  B.,  954,  956,  985,  1034 

Mc  Rae,  S.  C,  836,  1056 

Mc  Shan,  W.  H.,  29,  30,  31,  1000,  1034 

Madden,  T.  J.,  260,  489,  504,  1068 

Madinaveitia,  J.,  538,  546,  1034 

Madoff,    M.,   495,    992 

Madonska,   L.,   206,   988 

Madsen,   N.   B.,    19,   26,   547,   555,   648, 

649,  789,  803,  808,  811,  813,  853,  862, 

1034 
Maengwyn-Davies,  G.  D.,  444,  1034 
Magee,   P.   N.,   225,   990 
Magel,  T.  T.,  619,  997 
Mager,  J. ,848,  878,989 
Maggiolo,  I.  W.,  778,  1010 
Magin,  J.,  942,  945,  1057 
Mahadevan,  S.,  547,  554, 1035 
Mahler,  H.   R.,   16,  284,  511,  549,  555 

707,  710,  713,  781,  832,  849,  977,  981 

1034,  1035,  1054,  1063,  1064 
Mahowald,  T.  A.,  869,  1007 
Maio,  J.  J.,  394,  1035 
Maitre,  L.,  318,  1035,  1040 
Maitra,   P.   K.,   53,   228,   395,   838,   839, 

1035 
Maizels,  M.,  177,  209,  928,  1035 
Makino,  K.,  578,  1035 
Malachowski,  R.,  619,  1035 
Maley,  F.,  381,  382,  469,  474,  781,  837, 

1034,  1035 
Maley,  G.  F.,  469,  781,  837,  1035 
Malhotra,  0.  P.,  810,  1064 
Malkin,  A.,  485,  489,  492,  1035 
Mallus,  E.,  968,  1035 
Malmstrom,  B.  G.,  789,  810,  837,  1035 
Malvin,  R.  L.,  205,  924,  1035,  1062 
Manaker,  R.  A.,  978,  1015 
Manchol,  P.,  268,  340,  1048 
Mancilla,  R.,  406,  407,  839,  1035 
Mandel,  H.  G.,  478,  1016 
Mandelstam,  J.,  155,  156,  1035 
Mandelstam,  P.,  379,  380,  1008 
Manery,  J.  F.,  121,  1036 
Mangelsdorf,  P.  C,  64,  1026 


AUTHOR    INDEX 


1101 


Mann,  F.  C,  914,  1035 

Mann,  F.  D.,  914, 1035 

Mann,  K.  M.,  377,  1030 

Mann,  P.  J.  G.,  485,  1035 

Mannering,  G.  I.,  597,  1059 

Manners,  D.  J.,  429,  843,  1015 

Manning,  D.  T.,  271, 1035 

Manning,  G.  B.,  269,  833,  834,  1035 

Mano,  Y.,  518,  522,  523,  854,  1035,  1065 

Manoukian,  E.,  129,  1007 

Mansour,  J.  E.,  474,  1035 

Mapson,  L.  W.,  542,  547,  550,  662,  713. 

773,    774,    778,    781,    858,    859,    1035, 

1058 
Marchenko,  N.  K.,  194,  1050 
Marcus,  G.  J.,  121,  1036 
Marcus,  P.  I.,  449,  1036,  1059 
Mardashev,  S.  R.,  307,  1036 
Margulies,  S.  I.,  856,  1040 
Marinetti,  G.  V.,  151,  1036 
Marini,  M.,  458,  1036 
Maritz,  A.,  338,  348,  1070 
Mark,  H.  J.,  763,  1051 
Marks,  P.  A.,  166,  497,  1018,  1036 
Markwardt,  F.,  675,  707,  836,  1036 
Marre,  E.,  623,  850,  1036 
Marsh,   B.   B.,   704,  705,   721,   722,  865, 

876,  989 
Marsh,  C.   A.,  272,  426,  427,  428,  429, 

1008,  1031,  1032,  1036 
Marsh,  J.  B.,  178,  185,  1017 
Marsh,  M.  M.,  759,  1025 
Marshall,  J.  K.,  264,  1031 
Marshall,  M.,  781,  804,  1036 
Marshall,  P.  B.,  560,  1036 
Marsland,  D.,  964,  965,  1029,  1070 
Martell,  A.  E.,  13,  658,  739,  998,  1031, 

1036 
Martin,  A.  R.,  224,  236,  1003 
Martin,  D.  B.,  234,  1020 
Martin,  G.  J.,  259,  261,  586.  587,  1036 
Martin,  L.  E.,  63,  1012 
Martin,  R.  B.,  374,  988 
Martin,  R.  G.,  351,  861,  1036 
Martin,  S.  M.,  27,  677,  1036 
Martin,  S.  P.,  434,  1031 
Martonosi,  A.,  868,  939,  940,  1011,  1036 
Marui,  E.,  985,  1061 


Maruyama,  K.,  866,  939,  1011,  1042 

Maruyama,  M.,  892,  1022 

Maschmann,  E.,  685,  793,  1036 

Mason,  H.  L.,  670,  1036 

Mason,  J.,  972,  1001 

Mason,  M.,  64,  595,  600,  607,  608,  609, 

1036,  1068 
Mason,  ,S.  F.,  280,  1036 
Masoro,  E.  J.,  613,  1008 
Massart,   L.,   693.  1036 
Massen,  V..   16,   18,  48,  54,  60,  65,  79, 

80,  94.   174,   183,  243,  273,  275,  276, 

277,    774,    787,    788,    805,    856,    1036, 

1043,  1054 
Massini,  P.,  892,  1036 
Master,  R.W.  P.,  15,987 
Masuda,  T.,  535,  1036 
Masuoka,  D.  T.,  191,  215,  1036 
Matheson,  N.  A.,  593,  1025 
Mathews,  A.   P.,  963,  1036 
Mathews,  C.  K.,  479,  1036 
Matkovics,  B.,  503,  1037 
Matsuda,  R.,  618,  632,  1062 
Matsumoto,  H.,  33,  244,  1021 
Matsumura,  Y.,  78,  149,  179,  228,  1070 
Matsuo,  Y.,  709,  843,  1037 
Matsuoka,  M.,  611,  612,  1042 
Matsushima,  T.,  416,  417,  1025 
Matteucci,  W.  V.,  626,  1051 
Matthes,  K.  J.,  147,  273,  987 
Matthews,  J.,  387,  991 
Matthew,  M.,  238,  1037 
Matthews,  L.  W.,  114,  141,  1011 
Matthews,  R.  E.  F.,  194,  261,  1037 
Matthews,  R.  J.,  358,  1002 
Matthies,  H.,  946,  1045 
Maung,  K.,  429,  843,  1015 
Mauzerall,  D.,  859,  888,  1037 
Maver,  M.  E.,  688,  707,  1037 
Mavrides,  C,  612,  1004 
Maw,  G.  A.,  356,  1037 
Maxwell,  E.  S.,  858,  859,  1037 
May,  C.  E.,  383,  991 
Maybury,  R.  H.,  758,  759,  761,  1005 
Mayeda,  M.,  385,  1004 
Mayer,  A.  M.,  847,  1037 
Mayer,  R.  L.,  297,  304,  458,  459,  461,  462, 

1018,  1028 


1102 


AUTHOR    INDEX 


Mazelis,  M.,  151,  466,  1037 

Mazia,  D.,  969,  1037 

Mazur,  A.,  711,  1037 

Mazurova,  T.  A.,  224,  225,  1056 

Meares,  J.  D.,  938,  1034 

Mechler,  F.,  949,  1016 

Medina,    A.,    547,    553,    554,    782,    842, 

1018, 1037, 1041 
Medina,  H.,  550,  1001 
Meek,  W.  J.,  944,  945,  1033 
Mehl,  J.  W.,  691,  692,  1037 
Meinke,  M.,  145,  1002 
Meier,  R.,  875,  880,  883,  884,  968,  1037 
Meisel,  E.,  926,  927,  1064 
Meiss,  A.  N.,  831,  988 
Meister,  A.,  A.,  437,  1037 
Mello  Ayres,  G.  C,  778,  781,  1037 
Melnick,  I.,  333,  1031 
Melson,  G.  L.,  359,  1020 
Melville,  K.  I.,  952,  1009 
Mendel,  B.,  673,  677,  1037 
Mendelsohn,  M.  L.,  928,  1037 
Mendez,    R.,    669,    723,    724,    942,    944, 

1037 
Mendicino,  J.,  470,  1037 
Menegas,  R.,  429,  1013 
Meneghetti,  E.,  902,  1037 
Menke,  K.  H.,  875,  1051 
Menon,  G.  K.  K.,  228,  233,  234,  1037 
Menon,  V.  K.  N.,  64,  1040 
Mercker,  H.,  937,  1029 
Meridith,  C.  H.,  359,  1020 
Meriwether,  B.  P.,  409,  1060 
Meroney,  W.  H.,  925,  1065 
Merrett,  M.  J.,  237,  1037 
Merrifield,  R.  B.,  533, 1068 
Merritt,  P.,  221,  222,  993 
Mescon,  H.,  767,  1037 
Messina,  R.  A.,  770,  1037 
Metzenberg,  R.  L.,  781,  804,  838,  1036, 

1037 
Metzger,  R.  P.,  410,  839,  1037 
Meyer,  D.  K.,  909,  1037 
Meyer,    H.,    137,    868,    969,    1036,    1037 
Meyer,  J.  R.,  966, 1037 
Meyer,  R.  K.,  30,  31,  1034 
Meyer,  V.,  701,  1037 
Meyerhof,  O.,  292,  1037 


Michaelis,  L.,  259,  271,  416,  421,  1037 

Michel,  O.,  678,  689,  1047 

Michel,  R.,  678,  689,  1047 

Michelson,  C,  61,  1058 

Mickelson,  M.  N.,  77,  87,  104,  168,  1038 

Middlebrook,  M.,  689,  1038 

Mil,  S.,  463,  1038 

Mikulski,  P.,  988 

Mildvan,    A.    S,,    789,    793,    816,    819, 

1014,  1038 

Milholland,  R.  J.,  571,  574,  1049 

Millar,  D.  B.  S.,  787,  802,  1038,  1067 

Millbank,  J.  W.,  51,  1038 

Miller,  A.,  783,  1038 

Miller,  J.  J.,  195, 1038 

Miller,  D.  N.,  75,  1038 

Miller,  T.  B.,  917,  934,  935,  1007,  1038 

MiUer,  V.  L.,  957,  959,  961,  1038 

Millerd,  A.,   18,   19,  27,  53,  77,  78,   79, 

80,  82,  120,   172,   173,   189,  855,  994, 

995,  1017,  1038 
Millington,  P.  F.,  950,  1038 
Millington,   R.   H.,    137,    141,   273,   274, 

1065 
Mills,  G.  T.,  61,  427,  1038 
Mills,  J.,  592,  1034 
Mills,  J.  M.,  \\,2Z6,1044 
Mills,  R.  C,  18,  187,  543,  555,  840,  1024 

1047,  1064 
Mills,  R.  R.,  332,  1038 
Milofsky,  E.,  435,  1043 
Milstein,  C,  662,  705,  706,  711,  716,  717, 

780,  804,  810,  832,  1038,  1058 
Minakami,   S.,   163,  510,  798,  872,   888, 

1024,  1038,  1069 
Minatoya,  H.,  901,  1038 
Mirsky,   A.   E.,  32,   189,   393,   622,   624, 

661,    670,    671,    672,    690,    988,    1034, 

1038 
Mirsky,  I.  A,,  883,  884,  1015 
Misaka,  E.,  817,  1038 
Misra,  U.  K.,  421,  1038 
Mitchell,  H.  K.,  53,  1017 
Mitchell,  I.  L.  S.,  384,  1023,  1038 
Mitchell,  J.  H.,  Jr.,  585,  1055 
Mitchell,  M.  B.,  53,  1017 
MitcheU,  P.,  267,  690,  910,  912,  1038 
MitcheU,  R.  A.,  772,  782,  1038 


AUTHOR    INDEX 


1103 


Mitchell,  R.  B.,  221,  222,  223,  993 

Mitidieri,  E.,  289,  1063 

Mittermayer,  C,  708,  1038 

Mitz,  M.  A.,  367,  1069 

Miura,  N.,  985,  1038 

Miyachi,  S.,  892,  1038 

Mize,  C.  E.,  841,  1038 

Mizushima,  S.,  349,  1038 

Mochizuki,  Y.,   156,  1042 

Modell,  W.,  956,  957,  1038 

Modignani,  R.  L.,  394,  1038 

Mollering,   H.,   841,   993 

Moetsch,  J.  C,  957,  1046 

Mokrasch,  L.  C,  1038 

Molinari,  R.,  435,  511,  547,  551,  557,  558, 

709,    710,    717,    773,    843,    844,    849, 

1035,   1038,   1048 
Molnar,  D.  M.,  291,  292,  294,  1038 
Mommaerts,  W.  F.  H.  M.,  445,  939,  940, 

1014,  1025 
Momose,  G.,  2,  138,  1038 
Mondy,  N.  I.,  585,  1058 
Moniz,  R.,  519,  1005 
Monk,  C.  B.,  11,  13,  1007 
Monroy,  A.,   726,  1038 
Monson,  W.  J.,  36,  37,  41,  240,  288, 1004 
Montgomery,  C.  M.,  31,  35,  36,  46,  47. 

48,  51,  69,  70,  71,  75,  76,  81,  82,  83, 

88,  198,  240,  241,  1039,  1038 
Montgomery,  R.,  15,  996 
Mook,  W.,  943,  944,  945,  1007 
Mookerjea,  S.,  138,  144,  149,  218,  1039 
Moore,  B.  W.,  901,  902,  905,  1039 
Moore,  H.,  901,  905,  1024 
Moos,  C,  28,  446,  1023,  1039 
Moosburger,  A.,  750,  881,  884,  885,  1006 
Mora,  P.  T.,  459,  462, 1039 
Morales,    M.    F.,    446,    447,    1041,    1042 
Morawetz.  H.,  457,  1039 
Morgan,  E.  J.,  18,  25,  34,  35,  41,  661, 

662,  664,  1020,  1039 
Morgan,  H.  E.,  125,  263,  1039,  1045 
Morgan,  H.  R.,  194,  1039 
Morhara,  K.,  660,  1039 
Mori,  T.,  292,  293,  1018 
Morino,  Y.,  564,  1039 
Morisue,   T.,   564,   578,   676,   858,   1039, 

1064 


Morita,  T.  N.,  387,  1066 

Moriyama,  H.,  977,  980,  1039 

Morren,  L.,  747,  763,  1027 

Morris,  M.  P.,  244, 1039 

Morrison,  J.  F.,  707,  833, 1039 

Morrison,  J.  H.,  197,  1067 

Morrison,  J.  F,,  171,  185,  189,  467,  1014, 

1039 
Morrison,  J.  R.,  359,  1020 
Morrow,  P.  F.,  W.,  444,  1065 
Morse,  P.  A.,  Jr.,  481, 1022 
Morton,  H.  E.,  349,  1039 
Morton,    R.    K.,    20,    27,    63,   471,   481, 

485,    493,    510,    686,    787,    989,    995, 

1039,  1067 
Mosbach,  E.  H.,  15,  1044 
Moser,  J.  C,  780,  798,  865, 1050 
Moses,  F.  E.,  385,  1004 
Moses,  v.,  51,  53,  187,  236,  1039 
Motoyama,  H.,  969,  1025 
Mott,  J.  C,  962,  1046 
Moudgal,  N.  R.,  681,  1046 
Moudler,  J.  W.,  53,  73,  74,  78,  91,  93, 

124,  125,   194,  360,  1039,  1056 
Mounter,  L.  A.,  68,  662,  675,  684,  685, 

836,  1039 
Moussatche,  H.,  176,  212,  213,  1039 
Moyed,  H.  S.,  321,  481,  854,  1039,  1062 
Moyer,  I.  H.,  933,  1039 
Mozen,  M.  M.,  293,  1039 
Mozingo,  R.,  538,  1019 
Mudd,   J.   B.,    148,    149,   226,   232,   887, 

1039 
Mudd,  S.,  956,  972,  1055 
Mudge,   G.   H.,   78,   137,   179,  205,   623, 

626,  883,  917,  924,  928,  931,  932,  933, 

934,  935,  936,  1032,  1039,  1040,  1065 
Miiller,  A.  F.,  158,  1040 
Miiller,  E.,  664,  1040 
Miiller,  F.,  941,  951,  956,  957,  1040 
Miiller,  G.,  168,  228,  237,  504,  1040 
Miiller,  J.,  950,  1032 
Mueller,  J.  F.,  574,  577,  1040 
Miiller,  O.  H.,  929,  1065 
Mueller,  R.  T.,  852,  1064 
Muller-Rodloff,  I.,  664,  1040 
Muenst«r,   A.   M.,   23,   33,  46,   47,   1020 
Muir,  C,   133,  1040 


1104 


AUTHOR    INDEX 


Mullnos,  M.  G.,  724,  993 

Mullins,  L.  J.,  905,  1040 

Munier,   R.,    15,   999 

Murachi,  J.,  340,  810,  1040 

Murayama,  M.,  755,  1040 

Murphy,  B.,  946,  1027 

Murphy,  J.  V.,  209,  909,  1063 

Murray,  A.  W.,  471,  481,  989 

Murray,  D.  R.  P.,  259,  1040 

Murthy,  S.  K.,  547,  554,  1035 

Muscatello,  U.,  865,  989 

MuschoU,  E.,  318,  938,  1040 

Mushett,  C.  W.,  577,  1040 

Mustakallio,  K.  K.,  925,  1040,  1060 

Muus,  J.,  813,  1040 

Myant,  N.  N.,  149,  1022 

Mycek.  M.  J.,  375,  887,  1000,  1010 

Myerhof,  O.,  815,  866,  1045 

Myers,  C.  D.,  974, 1000 

Myers,  D.  K.,  60,  126,  816,  820,  867,  869, 

988,  1040 
Myers,  T.  C,  446,  474,  1039.  1054 
Myrback,   K.,   464,   660,   675,   683,   685, 

688,  691,  781,  791,  792,  838,  1040 


N 


Naber,  E.  C,  530,  1040 
Nachlas,  M.  M.,  856,  1040 
Nachmansohn,   D.,   683,   685,   1040 
Nadai,  Y.,  338,  1040 
Nadeau,  L.  V.,  964,  966,  1034 
Nadkarni,  S.  R.,  833,  1040 
Nadler,  J.  E.,  944,  1050 
Nagai,  S.,  26,  36,  62,  63,  855,  1029,  1069 
Naganna,  B.,  64,  1040 
Nagasawa,  M.,  662,  687,  1069 
Nagata,  Y.,  153,  1061 
Nagatsu,  I.,  544,  1068 
Nagatsu,  T.,  320,  1062 
Nagler,  H.,  947,  1050 
Naik,  K.  G.,  240, 1040 
Naite,  B.,  744,  1010 
Nair,  P.  M.,  547,  549,  1040 
Nair,  P.  V.,  437,  997 
Najjar,  V.  A.,  383,  1010 
Nakada,  H.  I.,  78,   152,   178,   187,  218, 
219,  228,  232,  233,  234,  387,  390,  392, 


398,   399,   404,   407,   857,   1040,  1052, 

1066,  1067 
Nakahara,  T.,  546,  988 
Nakamura,  H.,  847,  1040 
Nakamura,  K.,  225,  856,  1040 
Nakamura,  M.,  61,  196, 1040 
Nakamura,  M.,  61,  196,  1040 
Nakanishi,  K.,  817,  1038 
Nakao,  K.,  163,  1069 
Nakao,  M.,  817,  1017 
Nakata,  H.  M.,  236,  237,  1040 
Nakayama,  T.,  832,  898,  1041 
Nanninga,  L.  B.,  445,  1041 
Naoi,  M.,  537,  1005 
Naoi-Tada,  M.,  817,  1041 
Naono,  S.,  479,  1014 
Narrod,  S.  A.,  547,  550,  1041 
Narurkav,  M.  V.,  415,  1050 
Naschke,  M.  D.,  71,  1052 
Nason,  A.,  487,  490,  493,  551,  552,  553, 

843,  849,  851,  994,  1024,  1031,  1049, 

1068 
Nasu,  H.,  578,  676,  858,  1064 
Nath,  R.,  582,  1041 
Nathan,  H.,  283,  1006 
Nathans,  D.,  207,  265,  1041 
Natsume,  K.,  340,  1068 
Naurath,  H.,  365,  1006 
Navazio,  F.,  518,  1041 
Neame,  K.  D.,  266,  1041 
Needham,  D.  M.,  866,  869,  1014 
Negelein,  E.,  875,  1051 
Neidle,  A.,  887,  1000 
Neilands,  J.  B.,  437,  686,  710,  802,  825, 

845,  1041,  1067 
Neims,  A.  H.,  340,  1041 
Neish,  A.  C,  599,  1011 
Nelson,  C.  A.,  475,  1041 
Nelson,  E.  K.,  2,  225,  1041 
Nelson,  J.  M.,  297,  1014,  1033 
Nelson,  L.,  722,  864,  882,  991 
Nelson,  M.,  859,  1041 
Nelson,  M.  M.,  576,  1041 
Nelson,  W.  L.,  837,  1049 
Nerurkar,  M.  K.,  415,  1050 
Neuberger,  A.,  238,  751,  888,  1012,  1037 
Neufeld,  E.  F.,  507,  508,  409,  511,  512, 

842,  1007,  1041 


AUTHOR    INDEX 


1105 


Neuhaus,  F.  C,  270,  359,  360,  853,  1041 

Neuhaus,  H.,  981,  1041 

NeuhofF,  v.,  499,  1018,  1067 

Neuman,   W.   F.,  54,  58,   166,   175,  995 

Neumann,  H.,  457,  1024 

Neurath,    H.,   369,   370,   373,   735,    797, 

798,  810,  1014,  1025,  1040,  1041 
Newburgh,  R.  W.,  132,  1023 
Newcomb,  E.  H.,   136,   143,  1021,  1067 
Newey,  H.,  265,  1041 
Newhouse,  J.  P.,  953,  1056 
Newman,  M.  D.,  950,  1017 
Newmark,  M.  Z.,  773,  817,  1041 
Newton,  G.  G.  F.,  599,  987 
Neyman,  M.  A.,  540,  544,  993 
Ngai,  S.  H.,  532,  1041 
Nichol,  C.  A.,  571,  574,  582,  583,  1041, 

1049,  1069 
Nicholas,  D.  J.  D.,  547,  553,  554,  782, 

787,  842,  1008,  1037,  1041,  1064 
Nickerson,  M.,  217,  1041 
Nickerson,  W.  J.,  169,  880,  1041 
Niderland,  J.  R.,  954,  957,  1005 
Niedergang-Kamien,  E.,  881,  967,  1041 
Niederpruem,  D.  J.,  881,  1041 
Nielsen,  H.,  157,  1007 
Niemann,   C,   271,   349,   372,   373,   374, 

988,  1009,  1020,  1029,  1035,  1064 
Niemer,  H.,  291,  1041 
Nigam,  V.  N.,  63,  442,  1041 
Nihei,  J.,  446.  447,  1041,  1042 
Xilsson,   K.,   363,   1032 
Nilsson,  M.,  619,  1009 
Nimmo-Smith,  R.  H.,  772,  1041 
Ninomiya,  H.,  882,  1041 
Nirenberg,   M.   W.,   263,   389,   390,   391, 

1041 
Nishi,  A.,  368,  842,  1041,  1042 
Xishida,  J.,  439,  1062 
Nishimura,  J.  S.,  357,  787,  1042 
Nishimura,  S.,  463,  1042 
Nishiyama,  T.,  156,  1042 
Nistratova,  S.  N.,  947,  1042 
Niwa,  T.,  124,  135,  176,  184,  1026 
Nocito,  v.,  60,  338,  348,  994 
Noda,  H.,  939,  1042 
Noda,  L.,  60,  446,  447,  836,  847,  1028, 

1041,   1042 


Noe,  F.  E.,  952,  954,  1042 

Noguchi,  H.,  983,  1042 

Nogueira,  O.  C,  510,  552,  850,  1047 

Nohara,  H.,  816,  1042 

Noltmann,  E.,  406,  855,  996,  1042 

Nomaguchi,  G.  M.,  217,  1041 

Nomura,  M.,  51,  74,  77,  1059 

Nordlie,  R.  C,  357,  475,  1010,  1042 

Norris,  E.  R.,  289,  1028 

Norris,  J.  L.,  864,  1029 

Norris,  L.  C,  530,  1002 

Norton,  G.,  91,  105,  106,  111,  132,  149, 

154,  172,  182,  1048 
Nosoh,  Y.,  768,  1042 
Nossal,  P.  M.,  15,  62,  1029,  1042 
Nour  El  Dein,  M.  S.,  91,  678,  1053 
Novikoff,  A.  B.,  798,  865,  1042 
Novinger,  G.,  865,  912,  1030 
Novoa,  W.  B.,  433,  435,  436,  1042 
Novosel,  D.  L.,  360,  1039 
Nozzolillo,  C.  G.,  511,  553,  847,  1019 
Nuenke,  B.  J.,  681,  745,  750,  804,  812, 

1002,  1062 
Xiirnberger,  H.,  803, 1052 
Nukada,  T.,  611,  612,  1042 
Nutting,  M.-D.  F.,  657,  1023 
Nygaard,  A.  P.,  62,  435,  437,  683,  1042 
Nyhan,  W.  L.,  152,  156,  1042 
Nyteh,  P.  D.,  709,  718,  997 


Gates,  J.  A.,  315,  1042 

Ochoa,  S.,  62,  63,  64,  75,  83,  145,  175, 
181,  187,  224,  226,  234,  235,  463,  469, 
474,  597,  705,  830,  853,  990,  1008, 
1025,  1035,  1038,  1042,  1049,  1050 

O'Connor,  R.  J.,  270,  354,  674,  705,  831, 
1042 

Odeblad,  E.,  959,  1009 

Oesper,  P.,  711,  1047 

Officer,  J.  E.,  360,  1039 

Ogata,  K.,  156,  816,  1042 

Ogata,  M.,  156,  1042 

Ogilvie,  R.  F.,  954,  1042 

Oginsky,  E.  L.,  353,  1042 

Ogiu,  K.,  985,  1061 

Ogston,  D.,  560,  995 


1106 


AUTHOR    INDEX 


Ohashi,  S.,  977,  980,  1039 

Ohnishi,  T.,  868,  1061 

Ohno,  M.,  328,  1042 

Ohr,  E.  A.,  907,  1042 

Ohshima,  S.,  428,  987 

Ohta,  J.,  173,  1042 

Okada,  K.,  347,  1069 

O'Kane,  D.  J.,  848,  854,  855,  1007,  1039 

Okazaki,  R.,  470,  1042 

Okuda,  J.,  347,  1069 

Okui,  S.,  817,  1027 

Okunuki,    K.,    26,    27,    228,    229.    328, 

1020,  1022,  1042 
Olavarria,  J.  M.,  476,  1031 
Oleson,  J.  J.,  504,  505,  1016 
Oliphant,  J.  F.,  982,  1042 
Olivard,  J.,  575,  1042 
Oliver,  I.  T.,  603,  1042 
Olemucki,  A.,  779,  780,  1063 
Olsen,  N.  S.,  350,  1027 
Olson,  E.  J.,  714,  1042 
Olson,   J.   A.,   708,   709,   826,   840,   863, 

1042,  1043 

Olson,  J.  M.,  891,  1043 

Olson,  M.  E.,  336,  351,  1046 

Olson,  R.  E.,  75,  103H 

Omote,  Y.,  781,  827,  1001 

Onrust,  H.,  519,  774,  775,  854,  1043 

Ordal,  E.  J.,  675,  1033 

Ordal,  Z.  J.,  53,  74,  77,  79,  1032 

Ordin,  L.,  170,  209,  273,  274,  1043 

Oren,  R.,  435,  1043 

OrlofF,  J.,  917,  1043 

Ornstein,  N.,   199,  887,  964,  965,  1014, 

1043 
Orten,   J.    M.,    89,    91,    104,    109,    1002, 

1010,  1043 
Ortiz,  P.  J.,  64,  145,  224,  226,  234,  235, 

1008 
Osborn,  M.  J.,  581,  585,  1043,  1066 
Ota,  S.,  64,  792,  813,  1043 
Otsuka,  S.,  554,  1043 
Otsuka,  S.-I.,  880,  1053 
Ott,  J.  L.,  692,  1043 
Ott,  P.,  768,  1028 
Ott,  W.  H.,  562, 1043 
Ottolenghi,    P.,    55,    65,    179,   436,   485, 

1043,  1049 


Ouellet,  L.,  711,  772,  773,  1030 

Overall,  B.  T.,  15,  1043 

Overgaard-Hansen,  K.,  267,  1030 

Owens,  H.  S.,  15,  1043,  1057 

Owens,  R.  G.,  797,  1043 

Oxender,  D.  L.,  266,  1043 

Ozaki,  K.,  706,  709,  841,  1043 

Ozaki,  M.,  315,  316,  317,  318,  319,  320, 

611,  1012,  1018,  1062 
Ozawa,  G.,  573,  991 
Ozawa,  T.,  347,  772,  1069 


Paasonen,  M.  K.,  611,  7022 

Packer,  L.,  31,  154,  177,  1043,  1050 

Padilla,  A.  M.,  540,  997 

Padykula,  M.  A.,  927,  947,  1073 

PafF,  G.  H.,  215,  462,  995,  1043 

Pagan,  C.,  244,  1039 

Pahl,  J.  I.,  228,  232,  1045 

Paigen,  K.,  125,  1065 

Paige,  M.  F.  C,  224,  236,  1003 

Paine,  C.  M.,  912,  1009 

Palm,  D.,  359,  577,  1020 

Palmer,  G.,  772,  774,  805,  1043 

Palmer,  J.  F.,  918,  920,  997 

Palmer,  J.  K.,  91,  105,  106,  172,  190,  225, 

228, 1063 
Pan,  S.  F.,  90,  1069 
Pan,  Y.  L.,  984,  994 
Panagos,  S.  S.,  613,  1008 
Pantlitschko,  M.,  459,  695,  1043 
Papa,  M.  J.,  177,  1043 
Papacenstantinou,  J.,  433,  434,  435,  1043 
Pappas,   A.,    158,   1047 
Paquette,  L.  A.,  358,  1052 
Pardee,  A.  B.,  36,  38,  63,  67,  68,  73,  75, 

76,  82,  88,  105,  117,  177,  178,  326,  357, 

468,  480,  481,  816,  1011,  1043,  1068 
Park,  C.  R.,  125,  263,  1039,  1045 
Park,  J.  H.,  409,  714,  1042,  1060 
Park,  R.  B.,  409,  1043 
Parke,  D.  V.,  630,  990 
Parker,  A.  J.,  639,  1043 
Parker,  C.  A.,  292,  293,  294,  1043,  1044 
Parks,  R.  E.,  Jr.,  383,  478,  1000,  1002, 

1029,  1044 


AUTHOR    INDEX 


1107 


Parmar,  S.  S.,  38,  40,  42,  240,  1018 

Parpart,  A.  K.,  908,  1014 

Parr,  C.  W..  272,  405,  406,  998,  1044 

Partowi,  R.,  922,  1011 

Partridge,  A.  D.,  973,  983,  985,  1044 

Passey,  R.  D.,  225,  990 

Passonneau,  J.  V.,  385.   407,  474,  1024, 

1044 
Passow,    H.,    187,    188,    898,    899,    908. 

1012,  1044 
Pasternak,   C.   A.   472,   478,   1016,   1044 
Patel,  C,  464,  1021 
Patel,  R.  P.,  240,  1040 
Pattabiraman,  T.  N.,  816,  1044 
Patterson,  M.  K.,  Jr.,  804,  842,  1062 
Patterson,  P.  A.,  576,  1024 
Patterson,  W.  B.,  208,  1044 
Paul,  J.,   61,  428,   1038 
Paul,  M.  H.,  71,  140,  995.  1010 
Pauling,  L.,  43, 1044 
Payes,  B.,  616,  1044 
Pazur,  J.  H.,  389,  1044 
Pearse,  A.  G.  E.,  850,  1018 
Pearson,  A.  M.,  603,  1046 
Pearson,  D.  E.,  189,  1044 
Pearson,  J.  A.,  53,  77,  79,  80,  82,  173, 1017 
Pearson,  R.  G.,  11,  236,  1044 
Pease,  D.  C,  198,  678,  1044 
Pece,  G.,  623,  1036 
Pechstein,   H.,  259,  271,  421,  1037 
Peck,  H.  D.,  Jr.,  49,  79,  1000,  1011,  1044 
Pedersen,  T.  A.,  231,  1044 
Peel,  E.  W.,  538,  1019 
Pelczar,  M.  J.,  Jr.,  79,  1060 
Pelizza,  G.,  615,  1049 
Peluffo,  C.  A.,  203,  727,  1003 
Pendergast,   J.,    143,   1011 
Penefsky,  H.  S.,  548,  556,  705,  865,  869, 

873,  1044,  1046 
Penefsky,  Z.  J.,  713,  1061 
Pengra,  R.  M.,  292,  1012 
Penn,  N.,  552,  849,  1044 
Pennington,  R.  J.,  548,  556,  613,  1044 
Pentschew,  A.,  952,  1044 
Peralta,  B.,  669,  723,  942,  1037 
Pereira,  A.  S.  R.,  967,  1044 
Perez,  J.  E.,  978,  979,  980,  1027,  1044 
Perisutti,  G.,  870,  1001 


Perkins,  D.  J.,  737,  739,  748,  753.  1044 
Perkins,  M.  E.,  658,  676,  683,  686,  1017 
Perlmann,  G.  E.,  657,  1012 
Perova,  K.  Z.,  983,  984,  985,  1022 
Perri,  V.,  520,  526,  527,  528,  532,  578, 

1003,  1048 
Perry,  J.  J.,  78,  83,  1044 
Perry,   S.   V.,   684,   691,   692,   723.   816. 

938,  939,  989,  998 
Pershin,  G.  N.,  774,  976,  1044 
Person,  P.,  19,  29,  1044 
Persson,  B.,  464,  1040 
Perutz,  M.  F.,  755,  1014 
Peskoe,  L.  Y.,  574,  577,  1051 
Petering,  H.  G.,  287,  289,  538,  851,  1030, 

1044 
Peters,  E.  L..  834,  1061 
VeteTS,J.M.,  582, 1044 
Peters,  R.  A.,  36,  63,  75,  125,  181,  187, 

709,  844,  980,  1032,  1033,  1044 
Peters-Mayr,  T.,  368,  1013 
Peterson,  E.  A.,  156,  1044 
Petrucci,  D.,  174,  1044 
Pette,  D.,  1020 
Pfleiderer,  G.,  802,  810,  814,  857,  861, 

1014,  1015,  1044 
Phaff,  H.  J.,  421,  1033 
Phares,  E.  F.,  15,  1003,  1044 
Philips,  F.  S.,  698,  699,  700,  1012,  1045 
Philipsen,  L.,  976,  977,  979,  980,  981,  999 
Phillips,  A.  H.,  815,  817,  1045 
Phillips,  P.  H.,  77,  87,  124,  128,  176,  203, 

1030 
Philpot,  J.  St.  L.,  658,  1045 
Phizackerley,  P.  J.  R.,  845,  1009 
Pierpont,  W.  S.,  28,  79,  80,  82,  87,  1045 
Pietschmann,  H.,  1045 
Pihl,  A.,  639,  662,  663,  691,  697,  699, 

750,  751,  804,  1006,  1045,  1050 
Piloty,   O.,   664,  1045 
Pincus,  G.,  179,  183,  1009 
Pine,  M.  J.,  911,  1045 
Piquet,  J.,  835,  1010 
Fine,  A.,  596,  846,  1062 
Pirie,  N.  W.,  693,  712,  1019 
Pisanty,  J.,  723,  724,  1037,  1045 
Pitot,  H.  C,  593,  1028 
Pitts,   R.  F.,    917,    918,   919,   923,   924, 


1108 


AUTHOR    INDEX 


928,    929,    930,    934,    935,    960,    995, 
1014,  1026,  1045 

Pizer,  L.  I.  852,  1001 

Plant,  G.  W.  E.,  63,  509,  513,  539,  543, 
696,  705,  782,  844,  852,  872,  874, 
999,  1045,  1054,  1067 

Plass,  M.,  61,  987 

Piatt,  M.  E.,  689,  1051 

Piatt,  M.  H.,  359, 1020 

Pletscher,  A.,  318,  954,  1045,  1060 

Plummer,  D.  T.,  433,  1045 

Podber,  E.,  798,  865,  1042 

Poertzel,  H.,  750,  927,  1031 

Pogell,  B.  M.,  564,  1045 

Pogrund,  R.  S.,  314,  1045 

Pohle,  W.,  946,  1045 

Pointer,  N.  S.,  956,  988 

Polatnick,  J.,  176,  181,  194,  1045 

Polglase,  W.  J.,  W,  365,  1055 

Polimeros,  D.,  920,  921,  1013 

Polis,  B.  D.,  815,  866,  1045 

Pollock,  M.  R.,  697,  699, 1045 

Polonovski,  M.,  308,  1045 

Polynovskii,  O.  L.,  675,  857,  1045 

Pon,  N.  G.,  409,  1043 

Pontremoli,  S.,  407,  412,  995,  1013 

Ponz,  F.,  708,  837,  1045 

Pope,  H.,  228,  993 

Popinigis,  J.,  988 

Popjak,  G.,  146,  147,  711,  852,  886,  887, 
1004,  1013,  1018,  1045 

Porter,  C.  C,  315,  316,  317,  318,  562, 
572,  664,  1045,  1057 

Porter,  J.  W.,  234,  886,  988,  1057 

Porter,  W.  L.,  15,  996 

Portzehl,  H.,  938,  1045,  1065 

Post,  R.  L.,  1045 

Postgate,  J.  R.,  699,  1045 

Potter,  V.  R.,  3,  15,  22,  29,  30,  31,  35, 
36,  38,  41,  42,  63,  67,  68,  73,  75,  76, 
82,  88,  91,  98,  100,  101,  102,  103,  104, 
105,  112,  117,  126,  128,  138,  139,  175, 
177,  178,  201,  217,  218,  228,  232, 
309,  473,  478,  479,  481,  659,  676,  687, 
784,  856,  987,  994,  997,  1022,  1043, 
1045,  1046,  1047,  1058 

PoweU,  G.  W.,  444,  1004 

Prado,  J.  L.,  272,  273,  274,  1050 


Prairie,  R.  L.,  121,  1000 

Pratt,  E.  A.,  462,  1031 

Preiss,  B.,  137,  1037 

Pressman,  D.,  683,  783,  845,  1005 

Prestidge,  L.  S.,  326,  1043 

Price,  C.  A.,  61,  84,  117,  1046 

Price,  J.  M.,  1003 

Prince,  R.  H.,  9,  1032 

Pringle,  A.,  589,  1068 

Proctor,  C.  H.,  194,  1037 

Prouvost-Danon,  A.,  176,  212,  213,  1039 

Psychoyos,  S.,  387,  1062 

Pubols,  M.  H.,  854, 1046 

Putter,  J.,  695,  1046 

Pugh,   D.,  419,  1046 

Pujarniwcle,  S.,  225,  226,  1009 

Pullman,  M.  E.,  548,  556,  705,  865,  869, 

873, 1044, 1046 
Pupilli,  G.,  228,  1046 
Purpura,  D.  P.,  574,  1046 
Pursiano,  T.  A.,  623,  1013 
Purvis,  S.  E.,  589,  989 
Putnam,  F.  W.,  635,  1046 
Pyefinch,  K.  A.,  962,  1046 


Quaestel,  J.  H.,  2,  18,  20,  21,  26,  30,  31, 
32,  34,  35,  36,  40,  52,  55,  62,  76,  78, 
87,  113,  115,  135,  138,  144,  151,  152, 
153,  156,  176,  177,  184,  207,  237,  238, 
261,  305,  349,  381,  382,  432,  485, 
574,  589,  593,  613,  699,  991,  995, 
1002,  1009,  1013,  1016,  1019,  1023, 
1026,  1027,  1035,  1046,  1048,  1069 

Quesnel,  V.  C.  J.,  529,  1046 

Quinn,  J.  R.,  603,  1046 


Raaflaub,  J.,  210,  1046 
Raaschou,  F.,  918,  923,  996 
Rabin,  B.  R.,  787,  996 
Rabin,  R.  S.,  595, 1046 
Rabinovitch,  M.,  461,  712,  1046 
Rabinovitz,  M.,  336,  351,  1046 
Rabinowitch,  E.,   163,  1005 


AUTHOR    INDEX 


1109 


Rabinowitz,  J.  C,  575,  1046 

Racker,    E.,    121,    408,    409,    519;    548, 

556,    705,    837,    855,    865,    869,    873, 

1000,   1002,    1044,    1046,   104S,    1059 
Rafter,  G.  W.,  439,  714,  783,  1046 
Raghupathy,  E.,  681,  1046 
Ragland,  J.  B.,  451,  1046 
Rahatekar,  H.  I.,  673,  674,  830, 1046 
Rahman,  M.  A.,  81,  176,  183,  997 
Rahn,  O.,  1046 
Raiziss,  G.  W.,  957,  1046 
Rajagopalan,  K.  V.,  549,  711,  772,  803, 

1046,  1058 
Rail,  J.  E.,  678,  689,  1047 
Ram,  D.,  383,  396,  1047 
Ramachandran,  S.,  125,  1047 
Ramakrishnan,    C.    V.,    677,    830,    1047, 

1052 
Ramasarma,  T.,  839,  1047 
Ramel,  A.,  807,  990 
Ramsdell,  P.  A.,  657,  701,  1017 
Rand,  M.  J.,  318,  1003 
Randa,  V.,  391,  1057 
Randle,  P.  J.,  264,  376,  911,  991,  1001 
Ranzi,  S.,  726,  1047 
Rao,  D.  R.,  711,  1047 
Rao,  M.  R.  R.,  274,  673,  674,  830,  988, 

1046,  1047 
Rao,  N.  A.,  660,  685,  854,  1012,  1047 
Rao,  S.,  389,  1064 

Rapkine,  L.,  661,  662,  683,  687,  1047 
Rapoport,   S.,  489,  493,   817,   818,   820, 

1019,  1047,  1051 
Rapp,  G.  W.,  406,  1047 
Rapport,  D.,  613,  1008 
Raska,  S.  B.,  513,  1047 
Rassaert,  C.  L.,  333,  1012 
Rassweiler,  C.  F.,  618,  1047 
Rathbun,  R.  C,  626, 1047 
Ratner,  S.,  60,  116,  158,  338,  348,  994, 

1047 
Rauschke,  J.,  18,  991 
Raval,  D.  N.,  603,  610,  1053 
Raw,   I.,  510,  511,   552,   710,   849,   850, 

1035,  1047 
Ray,  T.,  928,  997 
Razzel,  W.  E.,  473,  1047 
Read,  C.  P.,  28,  173,  228, 1047 


Recknagel,  R.  0.,  138,  139,  1047 

Redetzki,  H.  M.,  505,  1047 

Redfearn,  E.  R.,  16,  1047 

Redfern,    S.,    659,    660,    662,    674,    683, 

684,  792,  833,  1003 
Rees,  K.  R.,  80,  81,  111,  114,174,  175, 

999,  1047 
Rege,  D.  V.,  478,  993 
Regen,  D.  M.,  125,  1039 
Reichard,   P.,   707,   711,   779,   780,   783, 

851, 1047 
Reichel,  G.,  314,  1003 
Reichert,  E.,  751,  1033 
Reichmann,  M.  E.,  980,  1047 
Reif,  A.  E.,  100,  1046 
Reinafarje,  B.,  126,  128,  987 
Reinbothe,  I.  H.,  15, 1047 
Reiner,  L.,  546,  1047 
Reinhardt,  F.,  1045 
Reisberg,  R.  B.,  707,  813,  835,  1047 
Reiss,  0.  K.,  38,  40,  42,  76,  240,  702,  816, 

878,  1017,  1018,  1047 
Rem,  L.  T.,  552,  710,  1005 
Remberger,  U.,  836,  1005 
Remington,  M.,  177,  209,  928,  1035 
Remy,  C.  N.,  966,  471,  1047 
Remy,   E.,  974,  1015 
Rendina,  G.,  840,  1047 
Rene,  R.  M.,  32,  204,  205,  626,  1053 
Rennels,  E.  G.,  925,  1047 
Rennick,  B.  R.,  204,  205,  626,  921,  923, 

1007,  1065 
Renson,  J.,  611,  989 
Repaske,  R.,  18,  26,  292,  293,  547,  553, 

847,  1047 
Repasky,  W.,  574,  1040 
Resch,  H.,  611,  1031 
Resnick,  H.,  650,  1047 
Resnik,  R.  A.,  833,  1067 
Revel,  H.  R.,  865,  1048 
Reynard,  A.,  480,  1049 
Reznikoff,  P.,  982,  1048 
Rhoads,  C.  P.  261,  1048 
Rhoads,  W.  A.,  225,  1048 
Rice,  B.,  120,  873,  878,  1015 
Rice,  L.  I.,  13,  75,  131,  216,  228,  1048 
Rice,  M.  S.,  886,  988 
Rich,  A.  E.,  973,  983,  985,  1044 


1110 


AUTHOR    INDEX 


Richards,  0.  C,  407,  1048 

Richert,  D.  A.,  287,  614,  783,  814,  859, 

1004,  1018,  1066 
Richter,  D.,  259,  296,  1048 
Rickenberg,  H.  V.,  394,  1035 
Ricketts,  C.  R.,  388,  389,  391,  996,  998 
Ridgway,  L.  P.,  576,  1024 
Rieken,  E.,  219,  220,  991 
Rieser,  P.,  267,  1048 
Rigbi,  M.,  457,  1048 
Riggs,  A.,  756,  757,  1048,  1067 
Riggs,  A.  F.,  756, 1048 
Riggs,   T.   R.,   155,   399,   575,   908,   999, 

1048 
Riker,  A.  J.,  171,  197,  1005,  1019 
Riklis,  E.,  207,  676,  1028,  1048 
Rimington,  C,  678,  1048 
Rimon,  S.,  816,  992 
Rindi,  G.,  520,  522,  526,  527,  528,  532, 

578,  1003,  1048 
Ringler,  R.  L.,  500,  798,  872,  1038 
Ritchie,  J.  L.,  838,  1012 
Rittenberg,  D.,  61,  293,  294,  676,  1008 

1019,  1028 
Rittenberg,   S.   C,    137,   228,    230,    231, 

1064,  1067 
Riva,  F.,  808,  1061 
Rivera,  G.  F.,  173,  882,  987 
Robbins,  W.  J.,  516,  528,  1048 
Robert,  B.,  772,  773,  836, 1048 
Robert,  L.,  772,  773,  836, 1048 
Robert,  M.,  772,  773,  836,  1048 
Roberts,  E.,  64,  269,  327,  332,  336,  569, 

772,    782,   857,   989,   991,   1048,    1051 
Roberts,  E.  R.,  292,  294,  1067 
Robertson,  A.  E.,  Jr.,  53,  77,  78,  81,  146, 

1000 
Robertson,  D.  H.,  625,  992 
Robertson,  M.  E.,  971,  1048 
Robertson,  R.  N.,  20,  27,  53,  63,  77,  79, 

80,  82,  173,  1017,  1067 
Robertson,  W.  van  B.,  686,  1048 
Robie,  C.  H.,  574,  577,  1051 
Robins,  R.  K.,  281,  282,  1007 
Robinson,  B.,  353,  1048 
Robinson,  J.  B.  D.,  966,  994 
Robinson,  J.  C.,  551,  709,  843,  1048 
Robinson,  J.  D.,  614,  1048 


Robinson,  J.  R.,  883,  928,  1048 

Robinson,  R.  J.,  76,  1001 

Robinson,  W.  G.,  782,  1048 

Robson,  J.,  965,  1048 

Robson,  J..S.,  921,  1029 

Rocca,  E.,  336,  709,  718,  840,  1048 

Roche,    J.,    268,    340,    678,    689,    1047, 

1048 
Rodnight,   R.,   485,  1034 
Rogach,  Z.,  297,  1018 
Rogers,  E.  F.,  530,  1048 
Rogers,  H.  J.,  453,  459,  1048,  1056 
Rogers,  K.  S.,  789,  816,  819,  1048 
Rogers,  L.  L..  588,  1048 
Rogers,  W.  I.,  640,  645,  992 
Rogers,  W.  P.,  54,  79,  80,  94,  174,  183, 

1036 
Rogulski,  J.,  206,  988 
Rohdenburg,  E.  L.,  751,  1015 
Roholt,  D.  A.,  Jr.,  660,  1014 
Rohutt,  O.,  683,  783,  845,  1005 
Rolf,  D.,  937,  1066 
RoHnson,  G.  N.,  169,  1048 
Romanchek,  L.,  873,  878,  987 
Romberger,  J.  A.,  91,  105,  106,  111,  132, 

149,  154,  172,   182,  1048 
Rona,  P.,  259,  271,  416,  1037 
Ronzoni,  E.,  766,  1063 
Roos,  B.-E.,  363,  1032 
Ropes,  M.  W.,  686,  1048 
Rosa,  N.,  122,  1049 
Roscoe,   H.   E.,   837,   1049 
Rose,  C.  L.,  999 
Rose,  F.  L.,  224,  236, 1003 
Rose,  I.  A.,  705,  830,  1049 
Rose,  W.  C.,  2,  219, 1001, 1049 
Rosell-Perez,  M.,  391,  1049 
Roseman,  S.,  356,  385,  1004,  1012 
Rosen,  F.,  571,  574,  1049 
Rosen,  S.  M.,  208,  909,  914,  1051 
Rosen,  S.  W.,  32,  35,  243,  1049 
Rosenberg,  A.  J.,  52,  64,  74,  81,  195,  989, 

1026,  1049 
Rosenberg,  H.,  685,  704,  707,  710,  836, 

845,  1006,  1010 
Rosenberg,  L.  E.,  264,  1049 
Rosenberg,  T.,  262,  461,  464,  1004,  1008, 

1049,  1066 


AUTHOR    INDEX 


nil 


Rosenthal,  A.,  595,  1049 

Ross,  C.  A.,  315,  1057 

Ross,  H.  E.,  336,  1032 

Ross,   J.    E.,   207,    208,    265,   911,   913, 

1018, 1041 
Ross,  R.  T.,  224,  1012 
Ross,  W.  F.,  15,  999,  1049 
Rossi-Fanelli,  A.,  518,  1041 
Rostorfer,  H.  H.,  845,  1001 
Roth,  J.  S.,  461,  462,  815,  1043,  1049 
Rothberg,  S.,  853,  1017 
Rothemund,  E.,  18,  30,  1070 
Rothschild,  A.  M.,  353,  1049 
Rothschild,  H.  A.,  60,  119,  121,  707,  717, 

781,  873,  1049 
Rothschild,  Lord,  175,  875,  882,  1000 
Rothstein,  A.,  876,  883,  893,  894,  895, 

897,  898,  899,  900,  902,  903,  904,  905, 

906,  907,  908,  958,  959,  960, 1003, 1044, 

1049,  1065 
Roughton,  F.  J.  W.,  757,  1012 
Roussos,    G.    G.,    462,    552,    553,    1031, 

1049 
Rovery,  M.,  649,  1003 
Rowan,  K.  S.,  22,  97,  171,  181,  182,  185, 

189,  190,  1016 
Rowe,  A.  W.,  803,  1049 
Rowe,  V.  K.,  627,  1056 
Rowland,  R.  L.,  744,  1049 
Roy,  A.,   177,   178,  1002 
Roy,  A.  B.,  443,  444,  1049 
Roy,  S.  C,  53,  153,  228,  838,  839,  1002, 

1035 
Rozenfel'd,  E.  L.,  417,  1033 
Rubbo,  S.  D.,  972,  973,  98S 
Rubin,  B.  A.,  170,  1049 
Rubin,  R.  J.,  480,  1049 
Rubino,  R.,  877,  1010,  1049 
Rubinstein,  D.,  55,  179,  485,  1049 
RuefF,  L.,  751,  1033 
RuflFo,  A.,  164,  615,  616,  1049 
Ruiz-Amil,  M.,  387,  388,  400,  1056 
Rule,  N.  G.,  375,  1033 
Rulon,  O.,  198,  1049 
Rumpf,  P.,  236,  1049 
Rundles,  R.  W.,  283,  1006 
Runnstrom,    J.,    726,    964,    1028,    1038, 

1049 


Ruska,  H.,  975,  1049 

Ruskin,  A.,  856,  883.  925,  941,  943,  948, 

1047,  1049,   1050 

Ruskin,  B.,  856,  883,  948,  1050 

Russell,  D.  S.,  951,  952,  953,  954,  1021 

Russell,  P.,  983,  1050 

Russo,  H.  F.,  264,  993 

Rust,  J.  H.,  Jr.,   177,  1043 

Rutter,  W.  J.,  407,  710,  779,  783,  846, 

1048,  1050 

Ryan,  C.  A.,  33,  38,  552,  1050 
Ryan,  E.  M.,  868,  1032 
Ryan,  J.,  798,  865,  1042 
Ryan,  K.-J.,  555,  713,  1050 
Rydon,  H.  N.,  321,  1016 
Ryley,  J.  F.,  173,  1050 
Ryzhkov,  V.  L.,  194,  1050 


Sabato,  G.  di,  210,  1050 

Sabine,  J.  C,  545,  549,  559,  1068 

Sable,  H.  Z.,  413,  1015 

Sacerdote,  F.   L.,  1058 

Sachs,  G.,  751,  1050 

Sachs,  H.  W.,  926,  1050 

Sacktor,  B.,  154,  446,  780,  798,  865,  1050, 

1054 
Sadhu,  D.  P.,  138,  144,  149,  218,  1039 
Saelhof,  C,  399,  400,  1056 
SafFran,   M.,    150,   272,   273,   274,   1050, 

1051 
Sage,  H.,  457,  1039 
Sahasrabudhe,  M.  B.,  415.  1050 
Saiga,  Y.,  676,  1053 
St.  John,  E.,  699,  700,  1012 
Saito,  K.,  435,  1043 
Sajgo,  M.,  817,  827,  1059 
Sakai,  H.,  727,  965,  1050 
Sakami,  W.,  570,  1050 
Sakamoto,   Y.,   64,   564,   578,   676,   858, 

1022,  1039,  1043,  1064 
Sakata,  K.,  401,  1050 
Salama,  A.  M.,  911,  1060 
Salaman,  M.  H.,  225,  990 
Salant,  W.,  942,  944,  946,  947,  948,  1050 
Salle,  A.  J.,  683,  685,  691,  692,  737,  968, 

972,  973,  1027,  1050 


1112 


AUTHOR    INDEX 


Salles,  J.  B.  V.,  63,  1050 

Saltman,   P.,   383,   842,    852,    910,   913, 

989,  1050 
Salvatore,  G.,  546,  972,  1007 
Salvin,  E.,  98,  99,  104,  110,  1028 
Samborski,  D.  J.,  196,  529,  1050 
Samiy,  A.  H.,  265,  1015 
Samorodin,  A.  J.,  939,  1058 
Sanabria,  A.,  924,  1050 
Sanadi,  D.  R.,  844,  856,  873, 1022,  1050 
Sanborn,  R.  C,  21,  29,  1050 
Sanders,  C.  R.,  388,  990 
Sanderson,  P.  H.,  920,  921,  1002 
Sankar,  D.  V.  S.,  520,  529, 1050 
Sanner,  T.,  691,  750,  751,  804,  1050 
San  Pietro,  A.,  490,  510,  553,  850,  851, 

891,  1003,  1025,  1030,  1050 
Santi,  R.,  983,  984,  992 
Santilli,  V.,  741,  980,  1025 
Sanwal,  B.  D.,  332,  840,  845,  1050 
Sagir,   D.,   924,   1065 
Sargent,  J.  R.,  693,  842,  1013 
Sarkar,  N.  K.,  710,  849,  1063 
Sarma,  P.   S.,  522,  676,  681,   711,   712, 

1046,   1054,   1058 
Saroff,  H.  A.,  763, 1051 
Sarreither,  W.,  125,  1029 
Sasaki,  A.,  794,  1051 
Sasaki,  S.,  78,  79,  86,  349,  844,  1024,  1051 
Sato,  R.,  842,  1051 
Sato,  T.  R.,  447,  1060 
Sato-Asano,  K.,  817,  1041 
Satta,  G.,  726,  1033 
Sauer,  G.,  818,  1051 
Sauermann,  G.,  397,  1051 
Saunders,  P.  R.,  55,  56,  76,  81,  178,  181, 

183,  191,  214,  215,  993,  1036,  1065 
Sawai,  T.,  420,  1051 
Sawatzky,  H.,  744,  1054 
Sayre,  F.  W.,  332,  1051 
Saz,  A.  K.,  599,  1051 
Saz,  H.  J.,  26,  168,  1051 
Scala,  A.  R.,  193,  1005 
Scala,  R.  A.,  539,  1051 
Scarano,  E.,  469,  675,  706,  1051 
Schaal,  R.,  236,  1049 
Schachter,  D.,  208,  355,  909,  914,  1004, 

1051 


Schacter,  H.,  694,  1051 

Schaefer,  M.  A.,  870,  1028 

Schaeg,  W.,  776,  1065 

Schanker,  L.  S.,  911,  1051 

Schapira,  G.,  308,  1045,  1051 

Scharff,  T.  G.,  263,  1051 

Schatz,  V.  B.,  257,  261,  1051 

Schatzberg,  G.,  878,  989 

Schauer,  R.,  910,  1051 

Schayer,  R.  W.,  353,  363,  1049,  1051 

Scheffer,  R.  P.,  27,  1068 

Schellenberg,  K.  A.,  673,  702,  703,  816, 

1017,   1051 
Schenker,  H.,  208,  909,  914,  1004,  1051 
Scheraga,  H.  A.,  683,  998 
Schimmel,  N.  H.,  626,  1051 
Schlegel,  D.  E.,  194,  1051 
Schlegel,  V.,  29,  1034 
Schleyer,  H.,  485,  487,  490,  1017 
Schlieselfeld,  L.  H.,  706,  718,  1030 
Schmid,  C,  120,  873,  878,  1015 
Schmidt,  G.,  439,  1051 
Schmidt,  S.,  297,  1051 
Schmitt,  A.,  707,  1054 
Schmitt,  J.  A.,  287,  289,  1044 
Schneider,    G.,    297,    1051 
Schneider,  K.  C,  293,  1019 
Schneider,  S.,  432,  817,  1020 
Schneider,  W.  C.,  3,  1045 
Schneiderman,  H.  A.,  199,  1051 
Schoeller,  W.,  941,  951,  956,  957,  1040 
Schom,  R.,  875,  1051 
Schonbaum,  E.,  150,  1051 
Schoniger,  W.,  4,  1019 
Schoepke,  H.  G.,  214,  1063 
Schorcher,  C,  924,  1051 
Scholefield,  P.  G.,  62,  122,  177,  200.  238, 

265,    266,    267,    395,    432,    596,    987, 

1001,  1046,  1051 
Schollmeyer,  P.,  178,  1027 
Schopfer,  W.  H.,  529, 1051 
Schor,  J.  M.,  325,  1010 
Schramm,  M.,  408,  409,  1046 
Schrauth,  W.,  941,  951,  956,  957,  1040 
Schrecker,  A.  W.,  582,  1051 
Schrodt,  G.  R.,  574,  577,  1051 
Schroeder,  E.  A.  R.,  837,  1046 
Schroeder,  E.  F.,  689,  1051 


AUTHOR    INDEX 


1113 


Schueler,  F.  W.,  44,  261,  1051 

Schiller,   H.,  670,  1051 

Schiitte,  H.  R.,  803,  1052 

Schuler,  M.  N.,   77,  87,   104,   168,  1038 

Schuler,  W.,  362,  1051 

Schulman,  M.  P.,  831,  1055 

Schultz,  G.,  1020 

Schultz,  S.  G.,  387,  1052 

Schulz,  D.  W.,  407,  474,  1024 

Schulz,  H.,  875,  1052 

Schuize,  H.  O.,  32,  388,  389,  391,  400, 

990 
Schumann,  E.  L.,  358,  1052 
Schutz,  B.,  210,  1021 
Schwarz,  K.,  878,  1001 
Schwartz,  W.,  195,  1052 
Schwarz,  D.  R.,  1000 
Schwarzenbach,    G.,   732,    739,   994 
Schweet,  R.  S.,  307,  1052 
Schweisfurth,  R.,  195,  1052 
Schwerin,  B.  G.,  664,  1045 
Schwert,  G.  W.,  373,  432,  433,  435,  436, 

782,  787,  802,  811,  1015,  1038,  1041, 

1042,  1054,  1059,  1067 
Scott,  J.  J.,  600,  998 
Scott,  R.  L.,  722,  909,  914,  1052 
Scott,  J.  W.,  77,  1052 
Scutt,  P.  B.,  292,  294,  1044 
Seal,  U.  S.,  802,  1052 
Sealock,  R.  R.,  518,  524,  1032,  1052 
Seaman,  G.  R.,   18,  28,  37,  71,  74,  91, 

173,  241,  242,  243,  1052 
Sebrell,  W.  H.,  495,  1033 
Seeley,  H.  W.,  591,  830,  1052 
Seehch,  F.,    125,  399,  695,  1031,  1043, 

1052 
Seevers,  M.  H.,  620,  622,  623,  624,  625, 

627,   628,   629,   630,   631,   1052,   1053, 

1068 
Segal,   H.   L.,   334,   358,   412,   472,   641, 

714,    760,    766,    806,    810,    856,    995, 

1012,  1020,  1052 
Segal,  R.,  816,  992 
Segal,  S.,  264,  1049 
Seibert,  M.  A.,  835,  870,  871,  1028,  1052, 

1056 
Seibert,  R.  A.,  933,  1016,  1039 
Sekiya,  K.,  868,  1061 


Sekuzu,  I.,  26,  843,  1042,  1052 

Sela,  M.,  457,  459,  462,  1048,  1052 

Seligman,  A.  M.,  856,  1040 

Selim,  A.  SS.  M.,  357,  712,  1052 

Sellinger,  0.  Z.,  335,  1052 

Semina,  L.  A.,  307,  1036 

Sen,  D.  K.,  338,  1052 

Senthe   Shanmuganathan,   S.,   783,   857, 

1052 
Serif,   G.   S.,   392,  393,  404,   1052,  1066 
Servettaz,  O.,  850,  1036 
Sery,  T.  W.,  978,  981,  1052 
Severac,  M.,  957,  1046 
Severens,  J.  M.,  972.  983,  984,  985,  1052 
Severin,  E.  S.,  359,  1026 
Sgaros,  P.  L.,  168,  225,  1052 
Shacter,    B.,    722,    879,    880,    883,    1052 
Shaffer,  C.  F.,  944,  957,  998 
Shah,  V.  K.,  830,  1052 
Shaner,  G.  A.,  264,  993 
Shanes,  A.  M.,  211,  1053 
Shannon,  L.  M.,  225,  226,  227,  228,  232, 

1053,  1069 
Shapiro,  D.  M.,  288,  505,  538,  569,  570, 

577,  1004,  1010,  1053 
Sharon,  N.,  326,  868,  1032,  1053 
Sharp,  A.  G.,  213,  747,  1065 
Sharpensteen,  H.  H.,  750,  1010 
Shaw,  E.,  281,  282,  1053 
Shaw,  P.  D.,  453,  1053 
Shaw,    W.    H.    R.,    64,    603,    610,    976, 

1026,  1053 
Shcherbakova,  L.  I.,  774,  976,  1044 
Sheets,  R.  F.,  902,  905,  1053 
Sheinfeld,  S.,  458,  1031 
Sheinin,  R.,  418,  1053 
Shemin,  D.,  160,  1053,  1068 
Shepherd,  D.  M.,  310,  353,  1034,  1048 
Sherman,  I.  R.,  273,  1053 
Sherman,  R.,  367,  1014 
Shetlar,  M.  R.,  838,  1032 
Shibasaki,  I.,  633,  851,  1053 
Shichi,  H.,  553,  1053 
Shideman,  F.  E.,  32,  204,  205,  212,  620, 

622,  623,  624,  625,  626,  627,  628,  629, 

630,  631,  1004,  1047,  1052,  1053,  1058, 

1065,  1068 
Shifrin,  S.,  500,  784,  1053 


1114 


AUTHOR    INDEX 


Shigeura,  H..T.,  467,  1053 

Shiio,  I.,  154,  880,  1053,  1061 

Shilo,  M.,  601,  1053 

Shils,  M.  E.,  505,  538,  569,  570,  1053 

Shimazono,  H.,  63,  1053 

Shimi,  I.  R.,  91,  678,  1053 

Shimizu,  T.,  179,  1059 

Shimomura,  O.,  676,  1053 

Shine,  H.  J.,  372,  1009 

Shinohara,  K.,  747,  1054 

Shiraki,  M.,  551,  555,  1022 

Shive,  W.,  354,  588,  1001,  1048 

Shkol'nik,  M.  Y.,  15,  1053 

Shonk,  C.  E.,  587,  995 

Shookhoff,  M.  W.,  5,  1066 

Shore,  B.,  873,  922,  926,  927,  1053 

Shore,  V.,  873,  922,  926,  927,  1053 

Shorr,   E.,   166,  1018 

Shrago,  E.,  846,  851,  1053 

Shrivastava,  D.  L.,  692,  831,  987 

Shug,   A.   L.,   293,   294,   854,   977,   981, 

1053,  1054,  1066 
Shukuya,   R.,  811,  1054 
Shull,  K.  H.,  772,  846, 1054 
Shulman,  A.,531,i05-^ 
Shwartzman,  G.,  569,  1054 
Sibly,  P.  M.,  834,  1054 
Sie,  H.-G.,  428,  1054 
Siebert,  G.,  707,  844,  852,  1054 
Siegel,  I.  A.,  216,  1065 
Siegel,  L.,  804, 1054 
Siegel,  S.  M.,  750,  1010 
Siegelman,  H.  W.,  173,  182,  1054 
Siekevitz,  P.,  444,  445,  865,  872,  1033, 

1054 
Sights,  W.  P.,  878,  883,  991 
Sih,  C.  J.,  708,  837,  870,  1054 
Silber,  R.,  582, 1054 
Silber,  R.  H.,  562,  572,  1045 
Silberman,  H.  R.,  281,  1054 
Sihprandi,  N.,  518,  1041 
Sillen,  L.   G.,   732,   733,   734,   736,   739, 

994,  1018,  1054 
Silva,  O.  L.,  693,  1056 
Silva,  R.  B.,  745,  754,  760,  1034 
Silverman,  J.  L.,  382, 1054 
Silverman,  M.,  582,  1010 
Simola,  P.  E.,  109,  1054 


Simon,  E.  W.,  19,  28,  60,  182,  837,  855, 
870, 1054 

Simon,  L.  N.,  474, 1054 

Simonds,  J.  P.,  924,  926,  1018,  1054 

Simonsen,  D.  G.,  772,  782,  1048 

Simonsen,  D.  H.,  50,  1054 

Simpson,  F.  J.,  713,  833,  857,  859,  993, 
1012,  1054 

Simpson,  J.  R.,  450,  1031 

Simpson,  R.  B.,  739,  740,  744,  745,  748, 
759,  1054 

Singer,  B.,  741,  1054 

Singer,  T.  P.,  16,  18,  29,  32,  38,  41,  46, 
48,  49,  65,  435,  437,  510,  541,  542, 
547,  549,  668,  673,  675,  676,  693,  703, 
706,  710,  713,  716,  718,  720,  721,  783, 
798,  803,  825,  833,  840,  845,  846,  857, 
866,  870,  877,  878,  991,  994, 1006, 1014, 
1025, 1036, 1038, 1054,  1065 

Singer,  Altbeker,  R.,  532,  992 

Sipos,  J.  C,  744,  1054 

Sirsi,  M.,  490,  847,  1013 

Siu,  P.  M.  L.,  852, 1054 

Sivaramakrishnan,  V.  M.,  522,  1054 

Sizer,  I.  W.,  64,  334,  551,  610,  657,  658, 
660,  675,  676,  685,  686,  692,  709,  843, 
1015,  1023,  1048,  1054,  1055 

Sjoerdsma,  A.,  315,  1042 

Skarnes,  R.  C,  459,  1055 

Skeggs,  H.  R.,  264,  993 

Skipper,  H.  E.,  585,  1055 

Skoda,  J.,  474,  1055 

Skou,  J.  C,  865,  869,  1055 

Slater,  E.  C,  33,  60,  61,  80,  83,  85,  121, 
122,  713,  715,  718,  810,  816,  820,  825, 
826,   867,  869,  871,  1032,  1040,  1055 

Slater,  G.  G.,  585,  1066 

Slater,  T.  F.,  553,  850,  1055 

Slaughter,  C,  357,  1008 

Slechta,  L.,  476,  1055 

Slein,  M.  W.,  376,  1055 

Slingerland,  D.  W.,  883,  910,  1055 

Sliwinski,  R.  A.,  406,  1047 

Sloane,  N.  H.,  356,  1055 

Sloboda,  A.,  504,  505,  1016 

Slotin,  L.,  63,  1007 

Sloviter,  H.  A.,  401,  1050 

SmaU,  P.  A.,  658,  1045 


AUTHOR    INDEX 


1115 


SmaUey,  H.  M.,  231,  997 

Smalt,  M.  A.,  744,  870,  1055 

Smeby,  R.  R.,  225,  992 

Smiley,  J.  D.,  61,  329,  998 

Smiley,  R.  L.,  363,  1051 

Smillie,  R.  M.,  27,  91,  1055 

Smith,  A.  H.,  89,   104,   109,  1010,  1043 

Smith,  C.  G.,  201,  413,  577,  968,  1055 

Smith,  D.  E.,  956,  972,  1055 

Smith,  E.  E.  B.,  61,  427,  1038 

Smith,  E.  L.,  60,  365,  375,  667,  668,  769, 

770,  783,  804,  1008,  1019,  1026,  1055, 

1057 
Smith,  F.,  15,  1055 
Smith,  G.  B.  L.,  362,  1032 
Smith,  G.  N.,  231,  997 
Smith,  H.  M.,  949,  1055 
Smith,  J.  E.,  840,  1055 
Smith,  J.   T.,   598,   615,   676,   711,   717, 

1016,    1055 
Smith,  L.,  163,  168,  892,  1055 
Smith,  L.  C,  434,  1034 
Smith,   L.   H.,  Jr.,  467,  470,  479,  1055 
Smith,  M.  E.,  507,  709,  717,  841,  1055 
Smith,  O.  H.,  49,  1044,  1065 
Smith,  P.  F.,  592,  1055 
Smith,  P.  H.,  953,  1064 
Smith,  P.  J.  C,  382,  832,  1055 
Smith,  S.  E.,  309.  310,  315,  318,  1055 
Smith,   W.,   550,   1001 
Smyrniotis,  P.   Z.,  413,  855,  1021 
Smyth,  D.  H.,  77,  87,  94,  95,  176,  265, 

387,  991,  1023,  1041,  1056 
Sneed,  M.  C,  736,  1055 
Snell,  E.  E.,  358,  452,  475,  539,  543,  544, 

561,  564,  565,  569,  575,  576,  578,  833, 

1001,   1003,   1020,   1026,   1033,   1034, 

1042, 1046, 1064 
Snodgrass,  P.  J.,  743,  785,  789,  825,  831, 

1055 
Snoswell,  A.  M.,  18,  547,  551,  1055 
Snow,  N.  S.,  756,  998 
Snyder,  R.,  831,  1055 
Snyder,  S.  H.,  693,  1056 
Soars,  M.  H.,  582,  589,  1018 
Sorbo,  B.,  803,  1056 
Sorbo,  B.  H.,  713,  1056 
Sohler,  M.  R.,  835,  1056 


Sohonie,  K.,  833,  1040 
Sokoloff,  B.,  399,  400,  1056 
Soldatenkov,  S.  V.,  224,  225,  1056 
Solomon,  A.  K.,  435,  1066 
Solomon,  J.  B.,  686,  845,  846,  1056 
Sols,   A.,   376,   379,   381,   382,   383,   387, 

388,  389,  391,  400,  414,  782,  824,  843, 

1001,  1056 
Solvonuk,  P.  F.,  825,  836,  854,  1056 
Somers,  E.,  976,  1056 
Sondheimer,  E.,  595,  1056 
Soodak,  M.,  516,  523,  530,  533,  998,  1056 
Sophianopoulos,  A.  J.,  610,  1056 
Sorkin,  E.,  518,  530,  1006 
Sorm,  F.,  474,  1055 
Sormova,  Z.,  474,  1055 
Sorof,  S.,  675,  703,  706,  716,  718,  833, 

1006 
Sorsby,  A.,  953,  1056 
Sourkes,    T.    L.,    308,    325,    544,    816, 

1029,  1056,  1067 
Southwick,  P.  L.,  518,  531,  1006 
Spackman,  D.  H.,  783,  1055 
Sparrow,  B.  W.  P.,  967,  994 
Speakman,  J.  B.,  762,  1019 
Speck,  J.  F.,  53,  73,  74,  78,  91,  93,  124, 

1056 
Spector,  A.,  662,  1056 
Spector,  W.  G.,  663,  909,  914,  1056 
Speer,  H.  L.,  169,  1056 
Spencer,  A.  F.,  147,  1056 
Spencer,   B.,  427,  444,   684.   1004,  1056 
Spencer,  D.,  615,  1056 
Spencer,  H.  C,  627,  1056 
Spencer,  R.  P.,  267,  268,  686,  1056 
Spensley,  P.  C.,  453,  459,  1048,  1056 
Speyer,  J.  F.,  273,  1056 
Spiegelman,  S.,  326,  351,  875,  1016,  1056 
Spiro,   R.   G.,   1056 
Spizizen,  J.,   194,  1056 
Spoeri,  E.,  195,  727,  972,  973,  1033 
Spooner,  D.  F.,  168,  179,  693,  882,  1010 
Spragg,  S.  P.,  852,  1021 
Spatt,  N.  T.,  199,  1056 
Spriestersbach,  D.,  15,  1055 
Sprinson,  D.  B.,  413,  1056 
SpjTopoules,  C.  S.,  950,  995 
Sreenivasan,   A.,   478,   993 


1116 


AUTHOR    INDEX 


Sreenivasaya,  M.,  15,  1024 

Srere,  P.  A.,  855,  1059 

Srinivasan,  P.  R.,  413,  969,  1056,  1057 

Stachiewicz,  E.,  435,  437,  1029 

Stadtman,  E.  R.,  356,  452,  1000,  1057 

Staehelin,  M.,  318,  1035 

Staemmler,  M.,  924,  1057 

Stafford,  H.  A.,  225,  1057 

Stafford,  J.  E.,  358,  1002 

Stahmann,  M.  A.,  456,  457,  458,  1003, 

1027,  1061 
Stalder,  K.,  98,  225,  226,  228,  235,  1057, 

1060 
Stamer,  J.  R.,  709,  1030 
Stanbury,  F.  A.,  730,  735,  962,  990 
Standen,  O.  D.,  772,  1041 
Stanley,  R.  G.,  27,  74,  80,  82,  1057 
Stanley,  W.  M.,  979,  980,  98S 
Stansly,  P.  G.,  233,  992 
Staple,  E.,  150,  1066 
Stare,  F.  J.,  55,  56,  75,  79,  81,  114,  115, 

141,    164,    175,    176,   991,    1011,   1057 
Stark,  J.  B.,  15,  1043,  1057 
Starr,  J.  L.,  817,  820,  988,  1057 
Stauff,  J.,  760,  1057 
Stauffer,  J.  F.,  171,  1018 
Steam,  A.  E.,  373,  1057 
Stebbins,  R.  B.,  577,  1040 
Steberl,  E.  A.,  234,  1057 
Stedman,  R.  L.,  74,  77,  632,  1057 
Steel,  K.  J.,  971,  972,  974,  1001,  1057 
Steel,  R.,  677,  1047 
Steele,  A.  B.,  620,  1057 
Steele,  R.,  401,  988 
Stegner,  H.-E.,  923,  1067 
Stein,  E.,  942,  945,  946,  1027,  1057 
Stein,  W.  D.,  906,  990 
Steinberg,  D.,  351,  614,  1057 
Steiner,  D.  F.,  391,  1057 
Steiner,  L.  A.,  459,  1052 
Stengle,  J.,  388,  401,  WL^9 
Stent,  H.  B.,  225,  226,  1058 
Stephenson,  M.  L.,  156,  887,  1057 
Stern,  J.  R.,  80,  228,  233,  234,  235,  708, 

751,  774,  837,  1037,  1057 
Stetten,  D.  Jr.,  208,  1044 
Stevens,  C.  0.,  709,  1057 
Stevenson,  T.  D.,  574,  577,  1051 


Stewart,  C.  J.,  392,  393,  394,  403,  404, 

996,  999,  1052,  1066 
Stewart,  H.  B.,  518,  1057 
Steyn-Parve,  E.  P.,  521,  525,  526,  1025, 

1027,  1057 
Stickland,  R.  G.,  65,  238,  597,  1057 
Stjernholm,  R.,  224,  235,  1057,  1067 
Stock,  C.  C,  576,  1024 
Stockell,  A.,  375,  497,  1057 
Stoeken,  L.  A.,  956,  1057 
Stoerk,  H.  C,  566,  576,  577,  587,  995, 

1057 
Stolen,  J.  A.,  228,  232,  1032 
Stolzenbach,  F.  E.,  497,  510,  1024,  1025 
Stone,  B.  A.,  132,  1057 
Stone,  C.  A.,  315,  1057 
Stone,  J.  E.,  478,  1058 
Stone,  R.  W.,  26,  36,  1012 
Stoneman,  F.,  626,  1058 
Stoppani,  A.  0.  M.,  18,  32,  38,  41,  74, 

77,  79,  81,  92,  104,  116,  117,  169,  687, 

705,  706,  708,  711,  712,  713,  716,  717, 

775,  778,  779,  780,  781,  783,  810,  832, 

838,  852,  853,  998,  1003,  1007,  1058 
Storey,  I.  D.  E.,  713,  817,  859,  860,  1058 
Stotz,  E.,  143,  151,  1036,  1067 
Stracher,  A.,  792,  866,  867,  1058 
Straessle,  R.,  681,  759,  761,  1005,  1021, 

1058 
Strait,  L.  A.,  694,  1062 
Strandskov,  F.  B.,  690,  1068 
Straub,  F.  B.,  29,  1058 
Straub,  R.  W.,  211,  949,  1014 
Strauss,  N.,  33,  38,  1058 
Strayhorn,  W.  D.,  745,  750,  1002 
Strecker,  H.  J.,  336,  410,  553,  850,  1012, 

1016,  1058 
Streicher,  J.  A.,  155,  999 
Strelitz,  F.,  677,  1037 
Strength,  D.  R.,  585,  1058 
Strickland,   K.   P.,   1058 
Strickler,  J.  C.,  206, 1058 
Stricks,  W.,  747,  748,  763,  1027,  1058 
Strittmatter,  C.  F.,  547,  711,  787,  847, 

1058 
Strittmatter,    P.,    685,    804,    809,    870, 

1058 


AUTHOR    INDEX 


1117 


Strohman,  R.  C,  939,  1058 
Strominger,  J.   L.,   359,   662,   713,   858, 

859,  1020,  1037,  1058 
Strong,  F.  M.,  260,  489,  504,  1068 
Stulberg,  M.  P.,  812,  1027 
Stumpf,  P.  K.,  61,  64.  136,  137,  138,  145, 

148,  151,  226,  228,  231,  684,  830,  832, 

1012,    1017,    1037,    998,    1021,    1039, 

1058 
Subrahmanyan,  V.,  407,  684.  692,  833, 

1018 
Subramaniam,    V.,    2,    225,    226,    228, 

998,  1058,  1064 
Suda,  M.,  857,  1021 
Siipfle,  K.,  975,  1058 
Sugiura,  H.  T.,  462,  1043 
Sullivan,  L.  P.,  924,  1062 
Sullivan,  M.,  467,  470,  479,  1055 
Sullivan,  R.  D.,  538,  1030 
Sullivan,  W.  J.,  929,  1014 
Sumizu,  K.,  859,  1058 
Summers,  W.  A.,  576,  1058 
Summerson,  W.  H.,  304,  1031 
Sund,  H.,  508,  785,  788,  792,  810,  831, 

1064 
Sundaram,  S.,  676,  712,  1058 
Sundaram,  T.  K.,  711,  1046,  1058 
Surtshin,  A.,  959,  1058 
Sussman,  A.,  195,  1058 
Sussman,  M.,  875,  1056 
Sutherland,  V.  C,  179.  181,  1006 
Sutton,  C.  R.,  780,  783,  810,  814,  1059 
Sutton,  W.  B.,  552,  1059 
Suzuoki,  T.,  173,  596,  924,  1059 
Suzuoki,   Z.,    173,   596,   882,   1041,  1059 
Svanberg,  0.,  813,  837,  1063 
Svedberg,  A.,  928,  997 
Svennerholm,  L.,  76,  81,  152,  1033 
Swan,  A.  A.  B.,  961,  1010 
Swan,  J.  M.,  694,  1059 
Swenson,  A.  D.,  643,  780,  788,  803,  833, 

1059 
Swensson,  A.,  959,  960,  1059 
Swim,  H.  E.,  89,  1003,  1059 
Swingle,  K.  F.,  35,  36,  1059 
Syrett,  P.  J.,  237,  1037 
Szabolcsi,   G.,   649,   788,   803,   812,   817, 

827,  1059 


Szego,  C.  M.,  31,  50,  176,  1031 
Szeinberg,  A.,  497,  1036 
Szent-Gyorgui,  A.,  689,  1038 
Szep,  O.,  960,  1059 
Szulmajster,  J.,  52,  60,  74,  81,  989 


Tabachnik,  M.,  855,  1059 

Tabor,  H.,  585,  1059 

Tachibana,  S.,  543,  554,  1059 

Tada,  M.,  537,  1005 

Taeger,  H.,  950,  1015 

Taufel,  K.,  168,  228,  237,  504,  1040 

Tager,  J.  M.,  153,  1059 

Taggart,  J.  V.,  204,  355,  921,  928,  1002, 

1009,  1051 
Tahmisian,  T.  N.,  876,  991 
Takagaki,  G.,   126,   127,   135,   153,   176, 

1059,  1061 
Takagi,  Y.,  471,  1059 
Takahashi,  H.,  51,  74,  77,  1059 
Takahashi,  M.,  880,  1053 
Takahashi,  N.,  444,  675,  816,  1059 
Takamiya,  A.,  787,  1025 
Takebe,  I.,  444,  1059 
Takemori,  A.  E.,  597,  1059 
Takemori,  S.,  842,  1051 
Takenaka,  Y.,  782,  811,  1059 
Takeuchi,  T.,  603,  610,  1059 
Talalay,    P..   447,   449,    712,    713,   1032, 

1036,  1059 
Tamari,  M.,  285,  992 
Tamura,  T.,  618,  632,  1062 
Tanada,  T.,  909,  1059 
Tanaka,  K.,  747,  973,  1059 
Tanaka,  M.,  542,  1059 
Tanaka,  R.,  518,  522,  523,  1035 
Tanaka,  S.,  179,  443,  859,  1058,  1059 
Tanaka,  Y.,  151,  676,  678,  887,  1059 
Tangen,   0.,   857,   1008 
Tanner,  F.  W.,  972,  983,  984,  985,  1052 
Tapley,   D.   F.,   90,   207,   208,   210,   265, 

681,   703,    762,   819,   911,   913,    1014, 

1018,  1023,  1024,  1041,  1059 
Tappel,  A.  L.,  75,  80,  91,  121,  137,  996, 

1015 
Tarr,  H.  L.  A.,  411,  1059 


1118 


AUTHOR    INDEX 


Tasaki,  I.,  950,  995 

Tashian,  R.  E.,  329,  1059 

Tashiro,  M.,  340,  1040 

Tata,  J.  R.,  555,  778,  1059,  1060 

Taube,  H.,  943,  944,  956,  957,  1031 

Tauroq,  A.,  209,  1060 

Taylor,  A.,  494,  987 

Taylor,  B.  B.,  670,  690,  1006 

Taylor,  E.  L.,  409,  1060 

Taylor,  E.  S.,  660,  845,  1060 

Taylor,   F.   J.,   379,   391,   1060 

Taylor,  H.  D.,  956,  1060 

Taylor,  J.,  601,  995 

Taylor,  J.  J.,  859,  1060 

Taylor,  K.  B.,  55,  179,  228,  988 

Teal,  J.  M.,  169,  1019 

Teitell,  L.,  228,  993 

Tekman,   S.,   645,   1017 

Telegdi,  M.,  409,  1025 

Telkka,  A.,  925,  1040,  1060 

Tenebaum,  L.  E.,  333,  1012 

Terner,  C,  75,  153,  176,  1060 

Terui,  G.,  633,  851,  1053 

Testa,  E.,  615,  1049 

Thannhauser,  S.  J.,  439,  1051 

Theis,  F.  V.,  696,  1060 

Theorell,    H.,   508,   544,    688,    779,   780, 

784,   785,  1060,  1069 
Therattil- Antony,  T.,  704,  723,  939,  990 
Thienes,  C.  H.,  55,  56,  76,  81,  178,  181, 

183,  214,  1065 
Thiessen,  C.  P.,  678,  992 
Thimann,  K.  V.,  156,  196,  887,  966,  968, 

1000,  1057,  1060 
Thoai,  N.,  268,  340,  1048 
Thoelen,  H.,  954,  1060 
Thomas,  C.  A.,  741,  1066 
Thomas,  G.  M.,  780,  798,  865,  1050 
Thomas,  G.  W.,  972,  973,  1060 
Thomas,  K.,  98,  225,  226,  228,  235,  1060 
Thompson,  C.  C,  168,  169,  190,  999 
Thompson,  E.  O.  P.,  770,  1055,  1060 
Thompson,  J.  W.,  667,  688,  707,  1014, 

1037 
Thompson,  R.  H.  S.,  854,  1060 
Thompson,  R.  L.,  193,  1060 
Thompson,  T.  E.,  789,  816,  819,  1048 
Thompson,  W.,  858,  1003,  1060 


Thomson,  J.  F.,  34,  38,  240,  447,  1060 
Thorn,  M.  B.,  18,  19,  22,  23,  24,  25,  32, 

33,  42,  717,  718,  1060 
Thome,  C.  J.  R.,  807,  846, 1060 
Thorne,  K.  J.  I.,  886,  1060 
Torsell,  W.,  464,  465,  989 
Threefoot,  S.,  928,  997 
Thunberg,  T.,  2,  32,  35,  37,  40,  175,  185, 

1060 
Thurlow,  S.,  280,  1004 
Thyagarajan,  B.  S.,  672,  1060 
Ticha,  M.,  389,  1027 
Tieckelmann,   H.,   531,   990 
Tietz,  A„   146,  147,  228,  233,  234,  887, 

1045,  1060 
Tietze,  F.,  29,  37,  242,  243,  1027,  1060 
Tilak,  B.  D.,  415,  1050 
Tinacci,  F.,  218,  219,  1060 
T'ing-seng,  H.,  662,  857,  1060 
Tin-Sen',  S.,  360,  995 
Tishler,  M.,  538,  1006 
Tissieres,  A.,  53,  1017 
Titus,  E.  O.,  281,  282,  285,  1005 
Tjutjunnikowa,  A.  W.,  835,  999 
Tobari,  J.,  845,  1026 
Tobie,  E.  J.,  91,  127,  995 
Tocco,  D.  J.,  911,  1051 
Todd,  A.  R.,  516,  992 
Tokuyama,  K.,  673,  675,  692,  717,  782, 

1060 
Tolba,  M.  K.,  911,  973,  978,  1060 
Tolbert,  N.  E.,  61,  999 
Tolman,  L.,  586,  1036 
Tomchick,  R.,  843,  996 
Tomkins,  G.  M.,  449,  1034 
Tong,  W.,  209,  1060 
Tonhazy,  N.  E.,  79,  1060 
Tonomura,  Y.,  816,  820,  866,  867,  868, 

939,  940,  1060,  1061 
Tooth,  B.  E.,  678,  1048 
Topper,  Y.  J.,  385,  1061 
Torda,  C,   165,  1061 
Tosi,   L.,   54,    114,    174,   552,   708,   870, 

994,  1011 
Tosteson,  D.  S.,  937,  1066 
Totaro,  J.  A.,  315,  316,  317,  318,  1045, 

1057 


AUTHOR    INDEX 


1119 


Totter,  J.  R.,  585,  845,  1001,  1061 
Tower,  D.   B.,  391,  392,  394,  395,  399, 

834,  1061 
Towsend,   E.,  325,  1056 
Trazler,   G.,   707,   1054 
Traniello,  S.,  708,  837,  1061 
Traub,  A.,  20,  29,  61,  80,  844,  848,  989, 

1013 
Trembley,  R.,  864,  997 
Trim,  A.  R.,  741,  1061 
Troger,  R.,  975,  1061 
Trudinger,  P.  A.,  321,  323,  1061 
Truscoe,  R.,  177,  209,  1035 
Tsao,  T.-C,  939,  1061 
Tschudy,  D.  P.,  591,  1061 
Tsen,    C.    C,    833,    845,    859,    901,    906, 

1061 
Tseng,  N.  S.,  513,  1062 
Tsuboi,  K.  K.,  452,  711,  713,  788,  839, 

842,  855,  1061,  1062 
Tsukada,  Y.,   126,    127,    135,    153,    176, 

1059,  1061 
Tsukaraoto,  H.,   196,  1061 
Tsunoda,  T.,  154,  1061 
Tsurumaki,   T.,   985,   1061 
Tsuyuki,  E.,  457,  1061 
Tsuyuki,  H.,  457,  1061 
Tubbs,  P.  K.,  61,  552,  1061 
Tuck,  L.  D..  694,  1062 
Tull,  F.  A.,  400,  1006 
Tuppy,  H.,  641,  1067 
Turano,  C,  808,  827,  1061 
Turba,   F.,   641,   938,   939,   1007,   1029, 

1061 
Turner,  A.  W.,  707,  1031 
Turner,  D.  H.,  831,  1061 
Turner,  J.  F.,  831,  853,  1017,  1061 
Turner,  J.  S.,  22,  97,  171,  181,  182,  185, 

189,  190,  1016,  1061 
Turner,  W.  A.,  2,  225, 1061 
Turpaev,  T.  M.,  947,  951,  958,  959,  1010, 

1042, 1061 
Turrian,  H.,  956,  1061 
Turrian,  V.,  956,  1061 
Tuttle,  L.  P.,  459,  993 
Tutunji,  B.,  95,  128,  1007 
Tyler,  D.  B.,  133,  1061 
Tytell,  A.  A.,  610,  1054 


u 

Uchida,  M.,  64,  1043 
Udaka,  S.,  481,  1062 
Udenfriend,  S.,  266,  310,  315,  316,  317, 

318,    319,    320,    354,    611,  999,  1002, 

1018,  1033,  1042,  1062 
tjhlein,  E.,  760,  1057 
Ukita,  T.,  618,  632,  1062 
Ulbricht,  T.  L.  V.,  530,  1062 
Ullberg,  S.,  958,  993 
Ulmer,  D.  D.,  785,  1032 
Ulrich,  F.,  209,  909,  1062 
Ulrich,  J.  A.,  529,  1062 
Umbreit,   W.    W.,    168,    350,    564,    566, 

1062,  1063 
Umeraura,  Y.,  439,  1062 
Ungar,  G.,  387,  1062 
Ungar-Waron,  H.,  285,  992,  993 
Unna,  K.,  562,  1062 
Urata,  G.,  674, 1062 
Uritani,  I.,  439,  1062 
Urivetzky,  M.,  855,  1062 
Usami,  S.,  78,  79,  86,  349,  1024 
Utter,  M.  F.,  676,  853,  1025 
Utzinger,  G.  E.,  694,  1062 


Vagelos,  P.  R.,  232,  234,  751,  774,  847, 

1020,  1062 
Vahlhaus,  E.,  855,  996 
Vaidyanathan,    C.    S.,    490,    547,    549, 

615,   660,   685,   832,   847,   1013,  1029, 

1047 
Valentine,  R.  J.,  268,  1056 
Vallee,   B.   L.,   743,   746,   770,  780,   781, 

785,    789,    806,    818,    825,    827,    831, 

1000,  1001,  1019,  1032,  1055,  1062 
Van  Aken,  G.  M.  F.  A.,  745,  998 
Van  Arsdel,  P.  P.,  725,  949,  995,  1062 
Van  Baerle,  R.  R.,  332,  550,  551,  1013, 

1062 
Van  Bibber,  M.  J.,  475,  992 
Vandemark,  P.  J.,  886,  1006 
Vandendriessche,    L.,    462,    693,    1036, 

1062 
Vander,  A.  J.,  920,  924,  1062 


1120 


AUTHOR    INDEX 


Van  der  Linden,  A.  C,  5ie,  774,  775,  854, 

1043 
Van  der  Schoot,  J.  B.,  320,  1062 
Van  Duuren,  A.  J.,  15,  1062 
Van  Eys,  J.,  434,   497,   498,    508,    513, 

784,  804,  842,  1062,  1064 
Van  Grembergen,  G.,  54,   173,  1062 
Van  Heyningen,  R.,  596,  662,  712,  846, 

1062,  1064 

Van  Oorschot,  J.  L.  P.,  597,  1062 

Vanov,  S.,  611,  1062 

Van  Pilsun,  J.  F.,  285,  1063 

Van  Rheenen,  D.  L.,  521,  525,  1025,  1027 

Van  Thoai,  N.,  779,  780,  1063 

Van  Vals,  G.  H.,  130,  149,  150,  156,  1063 

Van  Wagtendonk,  W.  J.,  50,  1054 

Vardanis,  A.,  474,  1063 

Varela,  877,  1049 

Vargas,  R.,  918,  1063 

Varner,  J.  E.,  269,  336,  662,  841,  1063 

Varrone,  S.,  678,  689,  1047 

Vasarhely,  F.,  470,  1037 

Vasilevskis,  J.,  739,  1008 

Vasington,   F.  D.,  209,   909,  910,   1063 

Vaslow,  F.,  373,  1063 

Vasta,  B.  M.,  122,  1022 

Vaughan,  M.,  351,  1057 

Vavra,  J.,  334,  1063 

Veeger,  C,  18,  38,  1003 

Veitch,  F.  P.,  238,  1063 

Velick,   S.   F.,   334,   766,   770,   785,   786, 

787,  802,  824,  870,  1058,  1063 
Velluz,  L.,  518,  1063 
Vely,  V.  G.,  26,  989 
Venditti,  J.  M.,  496,  1024 
Vennesland,    B.,    63,    226,    470,    1007, 

1010,  1063 
Ventura,  U.,  578,  1048 
Venturi,  V.  M.,  214,  1063 
Vercauteren,  R.,  676,  1063 
Vernberg,  W.  B.,  173,  183,  1063 
Vernon,  L.  P.,  557,  710,  847,  849,  1004, 

1063,  1066 

Vescia,  A.,  594,  708,  837,  1061,  1063 

Vesthing,  C.  S.,  437,  990 

Vickery,  H.  B.,  91,   105,   107,   172,   190, 

225,  228,  1063 
Villee,  C.  A.,  393,  399,  1063 


Villela,  G.  G.,  288,  289,  1063 

Vilter,  R.  W.,  577,  1040 

Vincent,  P.  C,  902,  906,  908,  1063 

Vining,  L.  C,  547,  549,  1040 

Virtanen,    A.    I.,    293,    706,   1029,    1039 

Vishniac,  W.,  675,  1026,  1063 

Vishwakarma,  P.,  207,  1063 

Visscher,  M.  B.,  910,  1022 

Vitale,  J.  J.,  208,  1063 

Vogel,  A.  I.,  3,  1063 

Vogel,  W.,  664,  1045 

Vogel,  W.  H.,  831,  1055 

Vogler,  K.  G.,  168,  1063 

Volk,  W.  A.,  411,  1063 

Von  Boventer-Heidenhain,  B.,  195,  1058 

von  Brand,  T.,  18,  25,  28,  91,  127,  173, 

203,  882,  987,  995 
von  Bruchhausen,  F.,  505,  1063 
von   Euler,   H.,   61,  500,  504,  685,  813, 

837,  987,  1063 
von  Euler,  L.,  901,  1038 
Von  Holt,  C,  520,  1063 

w 

Wachi,  T.,  64,  1043 

Wachstein,  M.,  926,  927,  1064 

Wachter,  W.,  701,  1037 

Wada,  E.,  225,  1064 

Wada,  H.,  561,  564,  578,  676,  858,  1039, 

1064 
Waddell,  J.  G.,  564,  1062 
Wadkins,  C.  L.,  18,  543,  555,  865,  872, 

874,  1031,  1064 
Wadso,  I.,  9,  1005 
Wadzinski,  I.  M.,  325,  1009 
Waelsch,  B.,  783,  1038 
Waelsch,  H.,  887,  1000 
Wagner,  R.  H.,  817,  818,  820,  1047 
Wagreich,  H.,  298,  299,  301,  1029 
Wahl,  R.,  979,  980,  1064 
Wainio,  W.  W.,  60,  61,  550,  1005 
Waisman,  H.  A.,  325,  1009 
Wajda,  I.,  363,  364,  994 
Wakelin,  R.  W.,  36,  1044 
Wakerlin,  G.  E.,  956,  1064 
Wakil,  S.  J.,  146,  224,  234,  707,  781,  887, 

996,  1014,  1064 


AUTHOR    INDEX 


1121 


Walaas,  E.,  541,  545,  1064 

Walaas,  O.,  541,  545,  1064 

Wald,   G.,   953,   1064 

Waldi,  I).,  518,  530,  1006 

Waley,  S.  G.,  594,  662,  712,  1000,  1062, 

1064 
Walker,  A.  D.,  77,  78,  1019 
Walker,  D.  G.,  389,  1064 
Walker,  E.,  982,  1064 
Walker,  G.  C,  787,  1064 
Walker,  J.,  960,  1014 
Walker,  L.  M.,  908,  1048 
Walker,  M.,  428,  1030 
Walker,  P.  G.,  419,  420,  1046,  1064 
Walker,  T.   K.,   2,   225,   226,   228,  399, 

998,  1048,  1068,  1064 
Wallace,  A.,  225,  226,  852,  1021,  1048, 

1064 
Wallace,  D.  M.,  428,  995 
WaUace,  R.  A.,  373,  1064 
Wallace,   R.   H.,   74,   77,   190,  228,  999 
Wallach,  D.  P.,  358.  1052,  1064 
Wallenfels,  K.,  508,  785,  788,  792,  810, 

831, 1064 
Wallgren,  H.,  1064 
Walsh,  E.  O'F.,  837,  1064 
Walsh,  G.,  837,  1064 
Walter,  C,  869,  1064 
Walter,  P.,  508,  1064 
Waltman,  J.  M.,  137,  1064 
Wanczura,  T.,  619,  1035 
Wang,  T.  P.,  509,  1064 
Warburg,  0.,  695,  768,  1064 
Waring,  G.  B.,  582,  1009 
Warkany,  J.,  953,  1064 
Warmke,  H.  E.,  244,  1039 
Warner,  C,  297,  298,  1064 
Warner,  R.  C,  463,  1064 
Warnock,  L.  G.,  434,  1064 
Warringa,  M.  G.  P.  J.,  33,  38,  49,  773, 

855,  1064,  1065 
Warzecha,  K.,  814,  861,  1015 
Wase,  A.,  960,  997 
Washio,  S.,  854,  1065 
Wasson,  G.  W.,  234,  1057 
Watanabe,  K.,  675,  1065 
Watanabe,  S.,  439,  1022 
Waterhouse,  D.  F.,  18,  1065 


Watertor,  J.  L.,  197,  1019 

Watkins,  W.  M.,  417,  418,  419,  1065 

Waton,  N.  G.,  352,  362,  1065 

Watson,  D.  W.,  459,  1055 

Watt,  D.  D.,  430,  1065 

Watts,  R.  W.  E.,  165,  1001 

Way,  J.  L.,  478,  1002 

Weakley,  D.  R.,  412,  1030 

Weatherall,  M.,  901,  905,  1024 

Weaver,  M.  E.,  200,  1065 

Webb,   E.   C,  444,  1065 

Webb,  J.  L.,  31,  35,  36,  46,  47,  48,  51, 

55,  56,  69,  70,  71,  75,  76,  81,  82,  83, 

88,  178,  181,  183,  213,  214,  217,  240, 

241,    625,    896,   943,   944,   945,   1039, 

1065 
Webb,  J.  L.  A.,  747.  1065 
Weber,  G.,  446,  772,  847,  997 
Weber,  H.  H.  938,  1065 
Weber,  J.  F.,  927,  1003 
Webster,  G.  C,  662,  841,  1063 
Weed,  L.  A.,  974, 1000 
Weed,  R.  I.,  902,  903,  904,  905,  906,  907, 

1065 
Weeks,  J.  R.,  212,  620,  622,  623,  624, 

625,  627,  628,  630,  631,  1052,  1065, 

1068 
Weeks,  T.  E.,  979,  1032 
Wehrli,  S.,  953,  1070 
Weibull,  C,  548,  864,  1014 
Weichel,  E.  J.,  757,  1033 
Weigert,  M.  G.,  281,  282,  285,  1005 
Weil-Malherbe,  H.,  75,  80,  95,  97,  379, 

1065 
Weiler,  P.,  Jr.,  335,  1052 
Weill,  C.  E.,  658,  660,  662,  673,  674,  683, 

684,  803,  833,  1049,  1065 
Wein,  J.,  38,  40,  42,  240,  1018 
Weinbach,  E.  C,  174,  273,  349,  704,  843, 

882,  987,  1010,  1065 
Weinberger,  R.,  926,  927,  993 
Weiner,  I.  M.,  917,  923,  924,  929,  931, 

932,  933,  934,  935,  1032,  1040,  1065 
Weinhouse,  S.,  137,  141,  149,  151,  152, 

178,  273,  274,  1023,  1040,  1065 
Weisman,    T.    H.,    195,    727,    972,   973, 

1033 
Weiss,  B.,  128,  176,  1065 


1122 


AUTHOR    INDEX 


Weiss,  L.,  614,  1066 

Weissbach,  A.,  413,  855,  1021 

Weissbach,  H.,  310,  103S 

Weisz-Tabori,  E.,  597,  1042 

Weitzel,  G.,  776,  852,  875,  881,  884,  1065 

Welch,  A.  D.,  261,  582,  583,  1041,  1065 

Welch,  K.,  209,  1065 

Wellman,  H.,  798,  869,  1030 

Wells,  H.  G.,  920,  1065 

Wells,  I.  C,  290,  1065 

Welt,  L.  G.,  925,  1065 

Wenger,  B.  S.,  773,  817,  1041 

Wenger,   H.   C,  315,   1057 

Wenner,  C.  E.,  125,  273,  274,  396,  397, 

1065 
Wenzel,  D.  b.,  216,  1065 
Wenzel,  F.,  15,  1004 
Werkheiser,  W.  C,  582.  583,  584,  1066 
Werkman,  C.  H.,  26,  78,  79,   164,   187, 

242,  430,  852.  987,  991,  996,  1065,  1068 
Werle,  E.,  352,  1066 
Wessels,  J.  G.  H.,  53,  74,  78,  79,  81,  104, 

1066 
Wessels  J.  S.  C,  445,  548,  864,  891,  892, 

1066 
Wesson,  L.  G.,  921,  1066 
Westcott,  W.  L.,  592,  1003 
Westerfeld,  W.  W.,  287,  614,  783,  814, 

859,  1004,  1018,  1066 
Westennan,  M.  P.,  834,  1006 
Westermann,    E.,    314,    315,    937,    1029, 

1066 
Westead,  E.  W.,  351,  1066 
Westheimer,  F.  H.,  5,  1066 
Westlake,  D.  W.  S.,  854,  1066 
Westling,  H.,  363,  1032 
Westmark,  G.  W.,  325,  1010 
Westveer,    W.  M.,    619,    632,    633,    1067 
Wetter,  L.  R.,  599,  706,  709,  841,  1011, 

1043 
Whatley,  F.  E.,  892,  989 
Wheatley,  A.  H.  M.,  2,  20,  31,  32,  34, 

52,  55,  113,  115,  144,  1046 
Whetham,   M.   D.,  2,   18,  35,  228,  1046 
White,  A.  G.  C,  504,  516,  528,  530,  1068 
White,  F.  G.,  847,  1066 
White,  H.  L.,  937,  1066 
White,  H.  S.,  524,  1052 


White,  I.  G.,  77,  1052 
Whitehead,  B.  K.,  422,  1019 
Whitehouse,  M.  W.,  150,  1066 
Whiteley,    H.    R.,    585,    675,    783,    847, 

1020,  1033,  1066 
Whitley,  R.  W.,  459,  998 
Whitmore,  F.  C.,  745,  1066 
Whittaker,  V.  P.,  662,  675,  685,  836,  854, 

1039.  1060 
Whittingham,  C.  P.,  891,  1066 
Wick,  A.  N.,  78,  187,  218,  219,  228,  232, 

233,  234,  387,  390,  392,  393,  394,  398, 

399,  403,  404,  410,  435,  839,  996,  999, 

1037,  1040,  1052,  1066 
Wickson-Ginzburg,   M.,   435,   1066 
Wiebelhaus,  V.  D.,  377,  1030 
Wieland,  O.,  614,  841,  993,  1066 
Wieland,   T.,    122,   814,   861,   1015 
Wien,  R.,  952,  956,  957,  1066 
Wiesmeyer,  A.,  415,  416,  780,  1066 
Wiethoff,  E.  O.,  518,  1066 
Wight,  K..  392,  400,  401,  1030 
Wigler,  P.  W.,  274,  277,  278,  1066 
Wilbrandt,  W.,  262,  900,  906,  907,  1049, 

1066 
Wilbur,  K.  M.,  54,  174,  1023 
Wilcox,  P.  E.,  739,  759,  1015 
Wilcox,  S.  S.,  410,  839,  1037 
Wilde,  C.  E.,  836,  1066 
Wilde.  W.  S.,  924,  1062 
Wilder,  V.,  878,  883,  991 
Wildman,  S.  G.,  58,   170,  191,  297,  995 
Wiley,  C.  E.,  461,  462,  712,  1008 
Wilkin,  G.  D.,  77,  78,  1019 
Wilkinson,  J.  H.,  433,  436,  1006,  1045 
Will,  J.  J.,  574,  1040 
Will,  L.  W.,  504,  505,  1016 
Willard,  H.  H.,  15,  1066 
Williams,  A.  D.,  585,  1066 
Williams,  A.  K.,  389,  1066 
Williams,  C.  H.,  Jr.,  817,  850,  1066 
Williams,  C.  M.,  21,  29,  199,  1050,  1051 
Williams,  D.  C.,  428,  995 
Williams,  D.  L.,  1031 
Williams,  G.  R.,  709,  994 
Williams,  H.  L.,  567,  568,  990 
Williams,  J.  N.  Jr.,  36,  37,  41,  240,  288, 

503,  509,  513,  1004,  1007,  1066 


AUTHOR    INDEX 


1123 


Williams,  K.,  428,  444,  684,  995,  1004 
Williams,  R.  H.,  391,  1057 
Williams,  R.  J.  P.,  658,  770,  1062,  1067 
Williams,  R.  T.,  225.  427,  630,  990,  1056 
Williamson,  D.  H.,  550,  662,  711,  855, 

1003,  1021 
Williamson,  S.,  335,  989 
Willing,  F..  499,  1067 
Wills,   E.   D.,   676,   686,   694,   710,   719. 

1067 
Wilson,  A.  N.,  518,  1067 
Wilson,  B.  R.,  224,  236,  1003 
Wilson,  G.  B.,  197,  1067 
Wilson,  J.  B.,  168,  1011 
Wilson,  J.  E.,  520,  1067 
W^ilson,  L.,  695,  990 
Wilson,  L.  Cx..  543,  554.  989 
Wilson,  P.  W.,  17,  26,  53,  291,  292,  293, 

294,  854,  996,  997,  1032,  1038,  1047, 

1053, 1066, 1067 
Wilson,  R.  H.,  601,  995 
Wilson,   T.   G.   G.,  292,  294,   404,   1067 
Wilson,  T.  H.,  263,  265,  387,  394,  403, 

1015,  1067 
Wilson,  W.  L.,  965,  1017 
Winder,  F.  G.,  77,  78,  1019 
Winer,  A.  D.,  433,  435,  436.  783,   787. 

845,  846,  1042,  1067 
Wingerson.  F.,  358,  1002 
Wingo.  W.  J..  328,  1067 
Winnick.  J..  156,  684,  685,  687,  1011, 1014 
Winestock,  C.  H.,  539.  543,  1067 
Winzler,  R.  J.,  585,  1066 
Winzor,  D.  J.,  682,  1067 
Wiseman,  G.,  265,  1067 
Wiseman,  M.  H.,  544,  1067 
Wiskich,  J.  T.,  20,  27,  63,  1067 
Wislicenus,  J.,  1067 
Withycombe,  W.  A.,  436,  1006 
Witkop,  B.,  316,  611,  1002,  1062 
Witlin,  B.,  690,  1011 
Witter,  A.,  641,  784,  1067 
Witter,  R.  F.,  143,  1067 
WMxon,  R.  L.,  708,  1024 
Wlodawer,  P.,  887,  1067 
Wockel,  W.,  923,  1067 
Wohl,  A.,  259,  421,  1067 
Wojtczak,  L.,  887,  1067 


Wolbach,  R.  A.,  756,  1048,  1067 
Wold,  F.,  409,  803,  1020,  1067 
Wolf,  P.  A.,  619,  632,  633,  1067 
Wolfe,  J.  B.,  78.  145,  187,  218,  219,  228, 

230,  231,  232,  233,  234,  387,  390,  399, 

407,  1040,  1066,  1067 
Wolfe,  R.  G.,  802,  1067 
Wolfe,  S.  J.,  522,  530,  1067 
Wolif,  H.  G.,  165,  1061 
Wolff,  J.  B.,  833,  1067 
Wood,  H.  G.,  164,  224,  235,  852,  1054, 

1057,    1067,    1068 
Wood,  J.  D.,  676.  1068 
Wood.  J.  G.,  834,  1054 
Wood,  N.  P.,  676,  693,  1004 
Wood,  R.  C.  584,  1068 
Wood,  W.  A.,  885,  1028 
Woodbury,  R.  A.,  921,  1011 
Woodruff,  L.  L.,  981,  1068 
Woods,  D.  D.,  260,  836,  988,  1068 
Woods,  L.  A.,  620,  622,  623,  624,  625, 

627,  628,  629,  630,  631,  1052,  1053, 

1068 
Woodward.   G.   E.,   385,   386,   389.   391, 

392,  689,  745,  1001,  1020,  1051,  1066, 

1068 
Woody,  B.  R.,  26,  1068 
Wooldridge,   W.   R.,   2,   21,   26,   35,   36, 

40,   62,   152,  237,  1046 
Woollen,  J.  W.,  419,  420,  1064 
Woolley,  D.  W.,  259,  260,  261,  281,  282, 

489,  504,  516,  518,  519,  522,  523,  528, 

529,    530.    533,    537,    538,    589,    1053, 

1068 
Wooltorton,  L.  S.  C,  120,  1023 
Worgan,  J.  T.,  427,  1032 
Work,  E.,  261,  593,  708,  1003,  1068 
Work,  T.  S.,  261,  1068 
Wormser,  E.  H.,  357,  1068 
Woronkow,  S.,  207,  1033 
Wortman,  B.,  707,  816,  1068 
Wosilait,  W.  D.,  713,  1068 
Wright.  C.  I.,  545,  549,  559,  1068 
Wright,  E.  M.,  387,  991 
Wright,  G.  F.,  744,  1054 
Wright,  L.  D.,  264,  993 
Wright,  N.  C,  658,  1068 
Wriston,  J.  C,  Jr.,  160,  1068 


1124 


AUTHOR    INDEX 


Wu,  H.  L.  C,  600,  1068 

Wu,  L.  C,  27,  1068 

Wurtz,  E.,  538,  1006 

Wyatt,  A.,  574,  1046 

Wyatt,  H.  v.,  453,  1068 

Wylie,  D.  W.,  611,  612,  1068 

Wyman,  J.,  638,  1005 

Wyngaarden,  J.  B.,  281,  282,  467,  474, 

480,  1054,  1068 
Wyngarden,  L.,  585,  1059 
Wynn,  J.,  602,  1068 
Wyss,  0.,  690,  1068 


Yoshimatsu,  H.,  578,  676,  858,  1064 

Yoshimura,  J.,  868,  939,  940,  1061 

Young,  B.  G.,  459,  1039 

Young,  G.  A.,  358,  1002 

Young,    L.    C.    T.,    80,    119,    120,    1000 

Young,  P.,  15,  1066 

Young,  R.  H.,  225,  226,  227,  228,  232, 

1053,  1069 
Younger,  F.,  569,  1048 
Yphantis,  D.  A.,  64,  334,  1023 
Yushok,  W.  D.,  381,  382,  385,  388,  395, 

396,  397,  1033,  1069 


Yagi,  K.,  340,  347,  537,  541,  544,  772, 

1005,  1060,  1068,  1069 
Yagi,  T.,  551,  555,  1022 
Yall,  I.;  237,  1013 
Yamada,  E.  W.,  60,  1069 
Yamada,  H.,  543,  554,  1059 
Yamada,  K.,  845,  1069 
Yamada,  T.,  880,  973,  975,  1069 
Yamafuji,  K.,  675,  1065 
Yamamura,  Y.,  26,  36,  62,  63,  148,  228, 

855,  1029,  1069 
Yamane,  T.,  739,  741,  1004,  1008,  1069 
Yamashita,  J.,  844,  1069 
Yamauchi,  M.,  224,  1069 
Yanagita,  T.,  880,  973,  975,  1069 
Yanari,  S.,  367,  1069 
Yang,  W.  C.  T.,  877,  1069 
Yanofsky,  C,  321,  1012 
Yashimatsu,  H.,  676, 1022 
Yates,  J.  R.,  657,  1019 
Yeas,  M.  F.,  175,  181,  1069 
Yee,  R.  B.,  90, 1069 

Yonetani,  T.,  508,   779,   780,   785,   1069 
Yoneya,  T.,  547,  549,  1069 
Yoneyama,  T.,  163,  888,  1024,  1069 
York,  J.  L.,  842,  996 
Yoshiba,  A.,  163,  1069 
Yoshida,  F.,  662,  687,  1069 
Yoshida,  H.,  151,  184,  1069 
Yoshihara,  I.,  194,  999 
Yoshikawa,  H.,  817,  1017 
Yoshikawa,  M.,  859,  1058 


Zahl,  P.  A.,  585,  1069 

Zakrzewski,  S.  F.,  582,  583,  1069 

Zalik,  S.,  122,  1049 

Zalusky,    R.,   387,   1052 

Zambonelli,  C,  984,  985,  1069 

Zamcheek,  N.,  208,  1063 

Zamecnik,  P.  C,  156,  887,  1057 

Zannoni,  V.  G.,  272,  306,  595,  1029,  1069 

Zarnitz,  M.  L.,  810,  1064 

Zarudnaya,  K.,  61,  64,  1058 

Zatman,  L.  J.,  485,  487,  490,  496,  498, 

504,  1069 
Zborowski,  J.,  887,  1067 
Zeliteh,  I.,  61,  438,  439,  782,  842,  1069 
Zeller,  A.,  335, 1005 
Zeller,    E.    A.,    60,   338,   348,   362,   365, 

1089,    1070 
Zellweger,  H.,  953,  1070 
Zewe,  v.,  376,  1010 
Zhanley,  J.  C,  854,  1046 
Ziegler,  D.  M.,  18,  1070 
Ziff,  M.,  684,  691,  692,  1070 
Zimmerman,  A.  M.,  964,  965,  1029,  1070 
Zimmerman,  S.  B.,  446,  1070 
Zink,  M.  W.,  845,  1051 
Zipkin,  I.,  630,  631,  1070 
Zittle,  C.  A.,  475,  1070 
Zollner,  N.,  18,  30,  461,  462,  1033,  1070 
Zubrod,  C.  G.,  538,  1030 
Zygmunt,  W.  A.,  359,  1070 
Zweifach,  B.  W.,  879,  895,  998 


SUBJECT  INDEX 

When  a  substance   is   named  in   connection  with  an  enzyme  or  process,  the  effect 
of  this   substance   on  the  enzyme  or  process,  usually  an  inhibition,  is  designated. 


Accumulation  of  substances,  see  specific 

substances  accumulated 
Acetaldehyde,    p\Tuvate    decarboxylase, 

432,  600 
Acetaldehyde    dehydrogenase,    o-iodoso- 

benzoate,  705 
Acetamide, 

A'-acetyl-/S-glucosaminidase,   419 

pyruvate  decarboxylase,  430 

urease,  603 
2-Acetamido-2-deoxygalactonolactone, 

xV-acetyl-yS-galactosaminidase,  420 

A'^-acetyl'/^-glucosaminidase,  419 
2-Acetamido-2-deoxygluconolactone, 

iV-acetyl-/3-glucosaminidase,  419 

carcinostasis,  428 
6-Acetaniidolevulinate,     aminolevulinate 

dehydrase,  592 
Acetanilide  deacetylase,  ferricyanide,  674, 
Acetate, 

accumulation  of,   benzoate,   349 

iV^-acetyl-/?-glucosaminidase,   419 

glutamate  decarboxylase,  328 

glyoxylate  transacetase,  594 

incorporation  into  lipid, 
malonate,    146-149 
mercurials,  887 
propionate,  613-614 

kynurenine:a-ketoglutarate     transami- 
nase, 608 

lactate  dehydrogenase,  436 

metabolism  in  rabbit,  2-deoxyglucose, 

399 

metaboUsm  of, 

fraris-cyclopentane-l,2-dicarbocylate, 
241 


6-deoxy-6-fluoroglucose,   404 
propionate,  613-614 
oxidation  of,  see  also  Respiration  (ace- 
tate) 

2-deoxyglucose,  397 
malonate,  77-78 
malonic  diethyl  ester,  236-237 
mercurials,  878,  883 
pantoate:/?-alanine  ligase,  597 
tyrosinase,  300-301 
Acetate  kinase  (acetokinase), 
o-iodosobenzoate,   705 
mercurials,  830 
Acetoacetate, 

accumulation  of,  tartronate,  237 
formation  of, 

from  butyrate,  613 
from  malonate,  234 
malonate,  138-144 
pathways,   139 
)?-hydroxybutyrate  dehydrogenase,  594 
metabolism  of,  malonate,   144 
Acetoacetate  carboxy-lyase,  see  Acetoace- 
tate decarboxylase 
Acetoacetate  decarboxylase, 
acetopyruvate,  591 
mercurials,  830 
Acetoacetyl-CoA,  splitting  by  mercurials, 

751 
Acetobacter 

ethanol  oxidation,  mercurials,  898 
tartronate  occurrence  in,  225 
Acetobacter  melanogenum,  pyruvate   oxi- 
dation, 

mercurials,  878 
Acetobacter  pasteurianutn,  succinate  oxi- 
dation, 
malonate,  52 


1125 


1126 


SUBJECT  INDEX 


Acetobacter  xylinum.,   cellulose  synthesis, 

malonate,  132 
Acetoin,   formation  from  pyruvate, 

oxythiamine-PP,  519 

phenylpyruvate,  430 
Acetokinase,  see  Acetate  kinase 
Acetopyruvate, 

acetoacetate   decarboxylase,   591 
Acetylation,  folate  analogs,  586-587 
Acetylcholine, 

biosynthesis  of,  malonate,  165-166 

cardiac  response  to, 
malonate,  217 
mercurials,  946-947 

ganglionic  response  to,  mercurials,  949 

intestinal    response    to,    hydrogen    pe- 
roxide, 696 
Acetylcholinesterase,  see  Cholinesterase 
Acetyl-CoA, 

formation  of,  propionate,  613 

splitting  by  mercurials,  751 
Acetyl-CoA  carboxylase, 

acyl-CoA  analogs,  614 

mercurials,  830 
Acetyl-CoA  kinase,   mercurials,  830 
Acetyl-CoA  synthetase,  propionate,  613 
Acetylene-carboxylate    hydrase,    malon- 
ate, 60 
Acetylene-dicarboxylate, 

intercharge  distance,  6 

pyruvate  oxidation,  240-241 

succinate   dehydrogenase,   34,   36,   38, 

240-241 
iV-Acetylgalactosamine, 

iV-acetyl-/^-galactosaminidase,   419-420 

A'^-acetyl-/3-glucosaminidase,  419 

a-galactosidase,  417 

/?-galactosidase,  419 
iV^-Acetyl-/5-galactosaminidase,   analogs, 

419-420,  429 
iV-Acetylglucosamine, 

iV-acetyl-/3-halactosaminidase,  419-420 

iV-acetyl-/S-glucosaminidase,  419 

a-galactosidase,  417 

jS-galactosidase,  418 

glucokinase,  390 

glucosamine  kinase,  593 

glucosamine  phosphorylation,  382 


glycogen  formation,  382 

hexokinase,   381-382,   390 
iV-Acetylglucosamine-6-phosphate,      glu- 

cose-6-P  dehydrogenase,  411 
iV-Acetyl-^-glucosaminidase,  analogs,  419 

429 
iV-Acetylglucosaminolactone,  iV-acetyl-/S- 

glucosaminidase,  429 
Acetylindoxyl  oxidase,  analogs,  591 
iV-Acetylisatin,     acetylindoxyl     oxidase, 

591 
iV-Acetylleucine,   leucine    decarboxylase, 

352 
Acetyllipoate,  biosynthesis  of, 

lipoate  analogs,  590 
Acetyl-D-phenylalaninamide,  chymotryp- 

sin,  372 
Acetyl-L-phenylalanine,  cathepsin  C,  375 
Acetyl-D-phenylalanine  methyl  ester,  chy- 

motrypsin,  372 
Acetylphosphatase,  o-iodosobenzoate,  705 
3-Acetylpyridine, 

alcohol  dehydrogenase,  498 

central  nervous  system,  499 

chick  embryo,  494 

heart, 

conduction,  494 
contractility,  499-500 
membrane  potentials,  499-500 

Lactobacillus  growth,  494 

lethal  doses,  499 

mechanisms  of  action,  494 

metabolism  of,  494-495 

iV-methylnicotinamide  formation  from, 

495 

NAD  levels  in  tissues,  495-496 

NAD  metabolism,  489-500 

NAD  nucleosidase,  491,  498 

nicotinamide  deamidase,  512 

nicotinamide  deaminase,  498 

nicotinate  deficiency,  489 

toxicity,  489,  494,  499 
4-Acetylpyridine,  toxicity,  499 
3-Acetylpyridine-NAD, 

brain  levels  of,  499 

formation  of  496 

function  as  coenzyme,  497 

glucose-6-P   dehydrogenase,   497 


SUBJECT  INDEX 


1127 


NADH:ferricyanide  oxidoreductase, 

510 

nicotinamide  deamidase,  512 

0-Acetylthiamine, 

pyruvate  oxidation,  519 
structure  of,  517 
thiamine  kinase,  523 

iV-Acetylthyroxine,  thyroxine  deiodinase, 
602 

Acetyltryptamine,  chymotrypsin,  371 

Acetyltryptophan, 
chymotrypsin,  374 
L-tryptophan-sRNA  ligase  (AMP),  326 

Acetyl-D-tryptophanamide,     chymotryp- 
sin, 371 

Acetyl  tryptophanates,  chymotrypsin,  371, 
373 

Acetyltryptophanmethylamides,    chymo- 
trypsin, 371 

Acetyl-D-tyrosinamide,  chymotrypsin,  371 

Acetyl-L-tyrosinate,  chymotrypsin,  371 

Acetyl-L-tyrosine, 

carboxypeptidase,  367 
tyrosinase,  304-305 

Ace  tyl-D- tyrosine    ethyl    ester,     chymo- 
trypsin, 371 

Acetyl  -  D  -  tyrosinehydroxamide,    chymo- 
trypsin, 371 

Acetyl  -  L  -  tyrosinemethylamide,    chymo- 
trypsin, 371 

Achro7nobacter,  nitrogen  fixation, 
oxygen,  292 

Achromobacter  fischeri, 

luminescence,    mercurials,   888-891 

mercurial  uptake,  897 

respiration  (glucose),   mercurials,   889- 

891 

SH  groups  in,  897 

Achromobacter  guttatum, 

malonate  occurrence  in,  225 
respiration  (endogenous),  malonate,  168 

Acid  phosphatase,  see  Phosphatase  (acid) 

Aconitase, 

frans-aconitate,  272-273 
ferricyanide,  674 
hydrogen  peroxide,  692,  694 
y-hydroxy-a-ketoglutarate,  615-616 
malonate,  60 


mercurials,  830 

propane-tricarboxylate,   240 
cis-Aconitate, 

glutamate   decarboxylase,   328 

oxalosuccinate  decarboxylase,  597 

oxidation  of,  malonate,  79 

phosphofructokinase,   385 
(rans-Aconitate, 

aconitase,  272-273 

citrate  accumulation,  273-274 

fatty  acid  synthesis,  273 

fumarase,  273,  275-276,  279 

glutamate  decarboxylase,  328 

glutamate    dehydrogenase,    331 

ion  transport  in  barley  roots,  274 

metabolism  of,  274 

occurrence  in  plants,  274 

paramecia,   274 

respiration,  273-274 

Venturia  ascosporulation,  195-196 
Aconitate  hydratase,  see  Aconitase 
Acridines,  chymotrypsin,  373 
Acriflavine   (Trypaflavine,   Euflavine), 

structure  of,  537 
Acrodynia, 

deoxypyridoxol,  566 

mercurials,  953-954 
Actin, 

association  with  myosin,  o-iodosoben- 

zoate,  723 

ATP  binding,  mercurials,  938-940 

Ca++  binding, 

o-iodosobenzoate,  723 
mercurials,  939-940 

o-iodosobenzoate,  704-723 

polymerization  of,  mercurials,  938-940 
Active  transport,  see  also  specific  organ- 
isms or  tissues 

analogs,  261-268 

malonate,   203-210 

mercurials,  205,  907-921,  928 

phlorizin,  205 

pyridoxal  analogs,  574-575 
Actomyosin, 

hydrogen  peroxide,  691 

o-iodosobenzoate,  723 

mercurials,   938-940 
Acylase,  mercurials,  831 


1128 


SUBJECT  INDEX 


iV-Acylated  glucosamines, 

carcinostasis,  382 

hexokinases,  381-382 
Acyl-CoA  dehydrogenase,   crotonyl-CoA, 

591 
Acyl  5'-nucleotidase,  AMP,  466 
Acyl  transfer,  lipoate  analogs,  590 
Adenine, 

adenosine  hydrolase,  466 

adenylosuccinate  synthetase,  467 

D-amino  acid  oxidase,  545 

ATPase,  445 

glucose  dehydrogenase,  501 

inosine  phosphorylase,  471 

malate  dehydrogenase,  509,  513 

NAD  nucleosidase,  488-490,  492-493 

nicotinate  deficiency,  513 

5'-nucleotidase,  472 

pyridoxal   kinase,   475,   477 

pyrophosphatase,  475 

xanthine  oxidase,  282 
Adenine    aminohydrolase,    see    Adenine 

deaminase 
Adenine  deaminase,  analogs,  466 
Adenocarcinoma, 

dehydrogenase   inhibition   in    vivo,    6- 

aminonicotinamide,  505 

growth  of, 

6-aminonicotinamide,   505 
riboflavin  analogs,  538 

pyridoxine  levels  in,   deoxypyridoxol, 

568 
Adenosinase, 

hydrogen  peroxide,  692 

mercurials,   831 
Adenosine, 

adenylosuccinate  lyase,  466 

adenylosuccinate  synthetase,  467 

alcohol   dehydrogenase,   506,   508 

D-amino  acid  oxidase,  545 

ATPase,  445 

creatine   kinase,   446 

glutamate  dehydrogenase,  508 

glutamate  semialdehyde  reductase,  507 

lactate  dehydrogenase,   501 

malate  dehydrogenase,  509,  513 

NAD  kinase,  509 

NADH  pyrophosphatase,  506,  511 


NAD  nucleosidase,  488,  490,  492 

5 '-nucleosidase,  472 

phosphodiesterase,  473 

pyridoxal  kinase,  475,  477 

pyrophosphatase,  475 

thiamine  kinase,  475 
Adenosine    aminohydrolase,    see    Adeno- 
sine deaminase 
Adenosine  deaminase,  analogs,  466,  477 
Adenosinediphosphate  (ADP), 

adenylosuccinate  synthetase,  467 

alcohol  dehydrogenase,  506,  508 

D-amino  acid  oxidase,  545 

AMP-ATP  transphosphorylase,  446 

arginine  kinase,  467 

ATPase,  444-445 

ATP-P,   exchange  reaction,  444 

creatine  kinase,  447 

deoxycytidylate  kinase,  469 

fructose- 1,6-disphosphatase,  470 

glutamate  semialdehyde  reductase,  507 

glutamine  synthetase,  471 

GTPase,  446 

hexokinase,  383 

isocitrate  dehydrogenase,  509 

ITPase,  446 

malate  dehydrogenase,  509 

NADH:menadione  oxidoreductase,  510 

NADH  oxidase,  506,  511 

NADH  pyrophosphatase,  506,  511 

NAD  kinase,  506,  510 

NAD:NADP    transhydrogenase,    507, 

510 

NAD  nucleosidase,  490,  492-493 

phosphodiesterase,  473 

phosphoribosyl-PP  amidotransferase, 

474 

pyridoxal  kinase,  475,  477 

pyrophosphatase,  475 

thiamine  kinase,  475 

UTPase,  446 

yeast  levels,  mercurials,  885 
Adenosine  hydrolase,  adenine,  466 
Adenosinemonophosphate    (adenosine-5'- 

phosphate,  adenylate,  AMP), 

acyl  5 '-nucleotidase,  466 

adenylosuccinate  lyase,  466 

adenylosuccinate  synthetase,  467 


SUBJECT  INDEX 


1129 


alcohol  dehydrogenase,  506,  508 

D-amino  acid  oxidase,  545 

arginine  kinase,  467 

deoxycytidylate  deaminase,  469 

flavokinase,  545 

fructose- 1,6-diphosphatase,   470 

glucose-6-P  dehydrogenase,   508 

glutamate    dehydrogenase,    508 

glutamate  semialdehyde  reductase,  507 

guanosine-5'-P   reductase,    471 

isocitrate  dehydrogenase,   509 

malate   dehydrogenase,   509 

NADH:menadione  oxidoreductase,  510 

NADH  oxidase,  511 

NADH  pyrophosphatase,  506,  511 

NAD  kinase,  509-510 

NAD:NADP     transhydrogenase,     507, 

510 

NAD  nucleosidase,  492 

NADPH:cytochrome  c  oxidoreductase, 

511 

NADPH  :  glutathione    oxidoreductase, 

512 

phosphatase,  439 

phosphodiesterase,   473 

pyridoxal  kinase,  475,  477 

pyrophosphatase,   475 

ribonuclease,  475 

thiamine  kinase,  475 

yeast  levels,  mercurials,  885 
Adenosinemonophosphate  -  3'  -  phosphate, 

phenol  sulfokinase,  473 
Adenosinemonosulfate,  AMP-ATP  trans- 

phosphorylase,  446 
Adenosine-2 '-phosphate,    NAD   nucleosi- 
dase, 490 
Adenosine-3 '-phosphate,  lactate  dehydro- 
genase, 501 
Adenosine-5'-phosphate,    see    Adenosine- 

monosphosphate 
Adenosine-5'-phosphsulfate  reductase,  io- 
dine, 684 
Adenosinetriphosphatase  (ATPase), 

ADP  inhibition, 

effect  of  Ca++  and  Mg++,  445 
effect  of  pH,  445 

analogs,  444-447 

Ca++,    453 


cystine,  663-664 

dehydroacetate,  623 

dithioglycolate,  663-664 

ferricyanide,  673 

hydrogen  peroxide,  691-692 

iodine,  682,  684 

o-iodosobenzoate,  705 

mercurials,  789,  863-869,  876,  905,  912 

aggregation,   789 

Ca++  and  Mg++  binding,   868 

configurational    changes,    868 

effects  of  Ca++  and  Mg++,  798 

effects  of  2,4-dinitrophenol,  869 

effects  of  pH,  792,  867 

initial  phosphate  release,  867-868 

kinetics,  869 

number  of  Hg++  ions  required,  912 

potentiation  by  substrate,  807 

protection  by  Mg++,  780 

rates  of  SH  group  reaction,  814 

relation    to    SH    groups,     803-804, 

806-807,  868-869 

stimulation,   815-816,   819-821 
porphyrindin,  668 
quinacrine,  547-548,  556 
Adenosinetriphosphate  (ATP), 
adenylosuccinate   lyase,   466 
adenylosuccinate  synthetase,  467 
D-amino  acid  oxidase,  545 
arginine  kinase,  467 
aspartate  carbamyltransferase,  468 
fructose- 1,6-diphosphatase,   470 
glutamate  semialdehyde  reductase,  507 
GMP  reductase,  471 
IMP  dehydrogenase,  471 
isocitrate  dehydrogenase,   509 
lactate  dehydrogenase,  501 
levels  in  nuclei, 

dehydroacetate,    624 

malonate,   189 
levels  in  tissues, 

aminopterin,  585 

2-deoxyglucose,  394-396 

o-iodosobenzoate,  721-722 

malonate,  157-158 

mercurials,  876 

quinacrine,  560 
levels  in  tumors, 


1130 


SUBJECT  INDEX 


2-deoxyglucose,  395 
D-glucosamine,  383 

levels  in  yeast,  mercurials,  884-885 

malate  dehydrogenase,  509,  513 

NADH:menadione  oxidoreductase,  510 

NADH  oxidase,  511 

NADH  pyrophosphatase,  511 

NAD:NADP  transhydrogenase,  510 

NAD  nucleosidase,  492 

phosphatase,  439 

phosphodiesterase,  473 

phosphoglucose  isomerase,   474 

phosphoribosyl  -  PP    amidotransferase, 

474 

polynucleotide  phosphorylase,  474 

pyrophosphatase,  475 

succinyl-CoA  deacylase,  475 
Adenovirus  type  5,  infectivity  of, 

mercurials,  977 
Adenylate,  see  Adenosinemonophosphate 
Adenylate  deaminase, 

ferricyanide,  674 

hydrogen  peroxide,  692 

mercurials,  774,  831 
Adenylate  kinase  (myokinase,  ATP:AMP 

phosphotransferase ) , 

adenosinemonosulfate,  446 

ADP,  446 

mercurials,  772,  847,  860 
Adenylmethylenediphosphonate,  ATPase, 

445-446 
Adenylmethylphosphonate,     polynucleo- 
tide phosphorylase,  474 
Adenylosuccinate  lyase, 

analogs,  466,  481 

mercurials,   831 
Adenylosuccinate     synthetase,     analogs, 

467 
Adipate, 

aspartate  -  a  -  ketoglutarate      transami- 
nase, 334 

carcinostasis,  201 

fumarase,  275 

glutamate  decarboxylase,  328 

glutamate  dehydrogenase,  330,  332 

intercharge  distance,  6 

ionization  constants,  8 

kynurenine:a-ketoglutarate     transami- 


nase, 595,  607-609 
lethal  dose,  201 

pyridoxamine :  oxalacetate  transami- 
nase, 600 

succinate  dehydrogenase,  35 
urinary  citrate,   109 

Adipose  tissue,  seealso  Epididymal  fat  pad 
fructose    metabolism,    2-deoxyglucose, 
391 
glucose  metabolism, 

6-deoxy-6-fluoroglucose,  404 
2-deoxyglucose,  391 
6-deoxyglucose,  403 
glycolysis,    2-deoxyglucose,    392 
methylmalonate  occurrence  in,  224 
respiration  (endogenous),  malonate,  178 
respiratory  quotient,  malonate,  185 

ADP,  see  Adenosinediphosphate 

Adrenals, 

catecholamine  release,  mercurials,  947 
corticosteroid  synthesis,  malonate,  150 
mercurial  levels  in,  959 

Aedes  aegypti, 

a -ketoglutarate  oxidation,  malonate,  80 
succinate  dehydrogenase,  malonate,  29 

Aerobacter  aerogenes, 

growth  of,  dehydroacetate,  632 
malonate  metabolism  in,  227-229 
respiration,  mercurials,  984 

Aerobacter  indologenes,  citrate  oxidation, 
malonate,  78 

Aerobic  glycolysis,  see  Glycolysis  (aero- 
bic) 

A-esterase,  iodine,  684 

Agmatine, 

cadaverine  oxidation,  363 
structure  of,  361 

Agrobacterium    turnefaciens,    growth    of, 
tungstate,  615 

Alanine,  phosphatase  (acid),  441 

/?- Alanine,  zl i-pyrroline-5-carboxylate  de- 
hydrogenase, 336 

D -Alanine, 

L-alanine  dehydrogenase,  354 

L-amino  acid  oxidase,  340 

Bacillus  cereus  germination  induced  by 

L-alanine,   270 

dipeptidase,  368 


SUBJECT  INDEX 


1131 


DL-Alanine, 

L-amino  acid  oxidase,  340 
phenylalanine  hydroxylase,  354 
L-Alanine,  arginase,  337 
L-Alanine  dehydrogenase, 
D-alanine,  270 
analogs,  354 
ferricyanide,  674 
o-iodosobenzoate,  705 
malonate,  60 
mercurials,  831 
Alanine:a-ketoglutarate  transaminase,  see 

Transaminases 
Alanine:pyriivate  transaminase,  see  Trans- 
aminases 
Alanine  racemase,  pyridoxal  analogs,  575 
D-Alanyl-D-alanine    synthetase,    D-cyclo- 

serine,  360 
Alcaligenes  faecalis, 

growth   of,    dehydroacetate,    632 
succinate  oxidation,  malonate,  52 
Alcohol,  see  Ethanol 
Alcohol  dehydrogenase, 

6-aminonicotinamide-NAD,  505 
o-iodosobenzoate,  705,  714-715,  717 
protection  by  ethanol,  717 
protection  by  NAD,  717 
reversal  by  GSH,  718 
titration  of,  714-715 
mercurials,  784-785,  788-789,  792,  803- 
806,  810-812,  825,  831 

coenzyme  displacement,  784-785 
NADH  binding,  788 
pH  effects,  792 

protection   by   ethanol,   779-780 
protection  by  NAD,  779-780 
rate  of  inhibition,  810-812 
relation  to  SH  groups,   803-806 
reversal  with  GSH,  825 
rotatory  dispersion  changes,  788 
sphtting  into  subunits,  789 
temperature  effects,  810-811 
NAD  analogs,  513-514 
nicotinyl-hydrazide-NAD,  497 
nucleotides,  506-508 
porphyrexide,  668 
pyridine  derivatives,  498 
pyridine-3-sulfonate,  504 


quinacrine,    548 
thionicotinamide-NAD,  497 
Aldehyde  dehydrogenase,  see  also  Acetal- 
dehyde   dehydrogenase 
o-iodosobenzoate,  704-706,  717 

protection  by  NAD,  717 
mercurials,  780,  832 

protection   by  NAD(P),   780 
Aldehyde:NAD    oxidoreductase,    GSSG, 

662 
Aldehyde  oxidase, 
mercurials,  772,  803 

relation  to  SH  groups,  803 
type  of  inhibition,  772 
quinacrine,  547,  549 
Aldolase, 

analogs,  407-408 
ferricyanide,  673-674,  677 
p-fluorophenylalanine  incorporation  in- 
to, 351 

hydrogen  peroxide,  692 
iodine,   682,   684 

mercurials,   643,   649,   780,    788,    803, 
832-833 

protection   by  fructose- 1,6-diP,   780 
rate  of  SH  group  reaction,  812 
relation  to  SH  groups,  803 
susceptibility  to  trypsin,  788 
p-mercuribenzoate,  643,  649 
SH  groups  of, 
pk„'s  of,  638 

titration     with     p-mercuribenzoate, 
643,  649 
sulfate,  414 
Aldose  1-epimerase,  see  Mutarotase 
Alfalfa,  malonate  occurrence  in,  224 
Alginate, 

phosphatase  (acid),  443,  464 
ribonuclease,  462 
Aliesterase,  quinacrine,  547,  549 
A'-Alkyl-2-amino-2-methylpropanediols, 

choline  oxidase,  290 
iV-Alkyl-2-amino-2-methylpropanols,  cho- 
line oxidase,  290 
iV-Alkyl-3-aminopropanols,    choline    oxi- 
dase, 290 
iV-Alkylethanolamines,    choline    oxidase, 
290 


1132 


SUBJECT  INDEX 


Alkylmalonates,    succinate    dehydrogen- 
ase, 37-38,  40 
Alkylnormorphines,    morphine    iV-deme- 

thylase,  590,  604 
AlHinase  (alHin  lyase),  o-iodosobenzoate, 

706 
Alhin  lyase,  see  Alliinase 
Allohydroxy  -  d  -  proline    oxidase,    quina- 

crine,  547,  549 
Allomyces  macrogynus, 

respiration  (endogenous),  malonate,  168 

succinate  dehydrogenase,  malonate,  27 
Allopurinol  (Zyloprim), 

urinary  excretion  of  xanthines,  283 

xanthine  oxidase,  283 
Allose,  fructokinase,  376-377 
Allose  kinase,  mercurials,  833 
Allose-6-phosphate, 

hexokinase,  379-380 

structure  of.   378 
Allothreonine  aldolase,  o-iodosobenzoate, 

706 
Alloxazine,  L-amino  acid  oxidase,  541 
Allylglycine,    DL-methionine    uptake    by 

ascites  cells,   265 
iV^-Allylnormorphine,  see  Nalorphine 
Alternaria  solani,  growth  of, 

dehydroacetate,  633 
Altrose,  fructokinase,  376 
Altrose-6-phosphate,  hexokinase,  380 
Ameboid  movement,  mercurials,  982 
Amethopterin   (Methotrexate), 

acetyl  transfer  in  liver,  586 

dihydro folate  reductase,  581-584 

folate  deficiency,  581 

formate  incorporation  into  proteins,  585 

purine  biosynthesis,  585 

resistance   to,   583 

structure  of,  580 

sulfonamide  acetylation,  586 

tetrahydrofolate   biosynthesis,    581 

tissue   distribution,    583 
Amidines, 

diamine  oxidase,  360-365 

structures  of,  361 
Amidophosphoribosyltransferase,  see 

Phosphoribosylpyrophosphate     amido- 

transferase 


Amidosulfonate,  see  Sulfamate 

Amine  oxidase,  see  Monoamine  oxidase 

Aminoacetone    synthetase,    ferricyanide, 

674 
Amino  acid  activating  enzymes,  see  spe- 
cific L-amino  acid:sRNA  ligases 
Amino  acid  decarboxylase,  see  also  spe- 
cific enzymes 
analogs,  317 
mercurials,  780 

protection  by  pyridoxal-P,  780 

protection   by  substrates,  780 
a-methyl-w -tyrosine  in  vivo,  315-317 
permanganate,  660 
D-Amino  acid  oxidase, 
adenine  analogs,  545 
amino  acid  analogs,  340-348 
analogs,   347-348,   545 

competition  with  FAD,  347-348 

competition    with    substrates,    347- 

348 

FAD  complexes,  347-348 
benzoates,   341-348 
o-iodosobenzoate,  704,  706,  717-718 

protection  by  FAD,  717 

protection  by  substrates,   717 

reversal  by  cysteine,  718 
kojic  acid,   349-350 
maleate,  343 
malonate,  60 
mercurials,  780,  804 

protection  by  alanine,  780 

protection  by  FAD,  780 

relation  to  SH  groups,  804 
p-mercuribenzoate,  competitive  nature 
of  inhibition,  771-772 
nucleotides,  545 
quinacrine,  547,  549,  557,  560 
riboflavin  analogs,  540-541,  544 
L-Amino  acid  oxidase, 
analogs,  338-340 

pH  effects,  340 
benzoate,  348 
D-leucine,  268 
malonate,  60 
D-phenylalanine,  268 
quinacrine,  547,  549 
riboflavin  analogs,  540-542 


SUBJECT  INDEX 


1133 


Amino  acid-sRNA  ligase,  stimulation  by 

mercurials,   816 
Amino  acids, 

active  transport  of,  p>Tidoxal  analogs, 

574-575 

deamination  of,  D-alanine,  270 

liver  levels  of,  mercurials,  954 

metabolism  of,  aminosulfonates,  350 

oxidation  by  hypochlorite.  658 

oxidation  by  permanganate,  657 

oxidation  of, 
kojic  acid,  350 
malonate,  151-154 

synthesis  from  glucose,  2-deoxyglucose, 

399 

transport  of,  analogs,  264-267 
Aminoacylase,  o-iodosobenzoate,  706 
2-Aminoadenine, 

adenine  deaminase,  466 

adenosine  deaminase,  466 
2-Amino-8-azapurine,    conversion    to    8- 

azaguanine,  281 
7n-Aminobenzoate, 

glucose  dehydrogenase,  501 

lactate  dehydrogenase,  501 
p-Aminobenzoate, 

acetylation  of,  benzoate,  349 

D-amino  acid  oxidase,  341 

glucose  dehydrogenase,  501 

lactate  dehydrogenase,  501 

transport    in    Flavobacterium,    p-ami- 

nosalicylate,  613 

tyrosine: a-ketoglutarate  transaminase, 

306 
Aminobenzoates,   D-amino   acid   oxidase, 

341,  348 
3-o-Aminobenzyl-4-methylthiazole 

(ABMT), 

structure  of,  517 

thiaminase,  524-525 
Aminobenzylthiazoles,    thiaminase,    524- 

525 
DL-a-Aminobutyrate,  homoserine  kinase, 

357 
L-a-Aminobutyrate, 

arginase,  337 

leucine  decarboxylase,  352 
y-Aminobutyrate  (GABA), 


brain  levels  of, 

aminooxyacetate,  358 
pj-ridoxal  analogs,  573-574 
toxopyrimidine,  578 
carboxypeptidase,  367 
oxidation  of,  malonate,  154 
A  ^-pyrroline-5-carboxylate  dehydrogen- 
ase, 336 
y-Aminobutyrate:a-ketoglutarate     trans- 
aminase, see  Transaminases 
f-Aminocaproate,  carboxypeptidase,  367 
a-Amino-/S-chlorobutyrate,   valine  incor- 
poration into  protein,  351 
1-Aminocyclopentane-carboxylate, 
glycine  uptake  by  ascites  cells,  265 
uptake  by  ascites  cells,  266 
uptake  by  brain,  266 
2-Amino-D-glucose,  see  D-Glucosamine 
Aminoguanidine, 

cadaverine  oxidation,  363 
diamine  oxidase,  362-363 
histamine  metabolism  in  man,  363 
histidine  decarboxylase,  352 
structure  of,  361 
p-Aminohippurate  (PAH), 
glycine  iV-acylase,  355 
renal  transport  of, 
azide,  205 
cyanide,  205 

dehydroacetate,  205,  625-626 
2,4-dinitrophenol,  205 
fluoride,  205 
fluoroacetate,  205 
iodoacetate,   205 
malonate,  203-205 
mercurials,  205,  920-921,  928 
phlorizin,  205 
2-Amino-4-hydroxy-6-formylpteridine, 

see  Pterin-6-aldehyde 
2-Amino-4-hydroxy-6-pteridyl    aldehyde, 

see  Pterin-6-aldehyde 
a-Aminoisobutyrate,   leucine   decarboxy- 
lase, 352 
^-Aminolevulinate    dehydrase,     analogs, 

591-592 
<5-Aminolevulinate  synthetase,  aminoma- 

lonate,  238-239 
Aminomalonate, 


1134 


SUBJECT  INDEX 


5-aminolevulinate  sjoithetase,  238-239 

condensation  with  aldehydes,  239 

formation  in  tissues,  239 

metabohsm  of,  239 
2-Amino-iV-methyladenine, 

adenine  deaminase,  466 

inosine  phosphorylase,  471 
2- Amino  -  2  -  methyl  - 1  -  propanol,     methio- 
nine biosynthesis,  291 
6-Aminonicotinamide,  504-505 

ADP-ATP  levels  in  tissues,  505 

carcinostasis,   505 

lactate  oxidation,  504 

lethal  doses,  504 

NAD  analog  of,  504-505 

NAD  levels  in  tissues,  505 

respiration   (endogenous),   504 

sulfanilamide  acetylase,  601 

toxicity,  504 
6-Aminonicotinamide-NAD,  504-505 

alcohol  dehydrogenase,  505 

creatine  kinase,  505 

formation  in  tissues,  504-505 

pyruvate  kinase,  505 
6-Aminonicotinate, 

NAD  nucleosidase,  488 

toxicity,  505 
Aminooxyacetate,  358-359 

anticonvulsant  activity,  358 

central  nervous  system,  358 

GABA  :  a-ketoglutarate  transaminase, 

358 

GABA  levels  in  brain,  358 

oxime  formation  with  pyridoxal-P,  358 

pyridoxal  deficiency,  358 

pyridoxal  kinase,  358-359 
Aminopeptidases, 

analogs,  367-368 

cysteine,  663 

ferricyanide,    674 

GSH,  663 

GSSG,  662 

mercurials,  787,  833 

displacement  of  Mn++,  787 
TO-Aminophenol,    D-amino   acid   oxidase, 

348 
p-Aminophenol,    D-amino    acid    oxidase, 

344 


Aminophenols,  tyrosinase,  304 

p-Aminophenylalanine,  L-phenylalanine: 
sRNA  hgase  (AMP),  355 

Aminopterin, 

acetyl  transfer,  586 

ATP  level  in  tissues,  585 

dihydrofolate  reductase,  581-584 

folate  deficiency,  581 

glycine-serine  inter  con  version,  585 

NAD  level  in  liver,  585 

nucleic  acid  biosynthesis,  585 

purine  biosynthesis,  585 

respiration,  585 

structure  of,  580 

tetrahydrofolate  synthesis,  581 

uptake  by  Bacillus  suhtilis,  mercurials, 

911 

uptake  by  bacteria,  584 

6-Aminopurimidine  deoxyribonucleoside- 
5 '-phosphate  deaminase,  o-iodosoben- 
zoate,  706 

Aminoquinolines,  chymotrypsin,  373 

p- Aminosalicylate, 

D-amino  acid  oxidase,  348 
p-aminohippurate  transport,  613 
protocatechuate  transport,  267 
sulfanilamide  acetylase,  601 

a-Aminosulfonates,  amino  acid  metabo- 
hsm, 350 

Aminotransferases,  ses  Transaminases 

3-Amino-L-tyrosine, 
tyrosinase,  304-305 
tyrosine: a-ketoglutarate  transaminase, 
306 

f5-Aminovalerate, 

carboxypeptidase,  367 
A  i-pyTroline-5-carboxylate   dehydroge- 
nase, 336 

^-Aminovalerate :  a  -  ketoglutarate  trans- 
aminase, see  Transaminases 

Amiphenazole,  see  2,4-Diamino-5-phenyl- 
thiazole 

Ammonium  chloride,  renal  refractoriness 
to  mercurials,  933-934 

Ammonium  ion, 

L-amino  acid  oxidase,  338 
formation  in  kidney, 
benzamide,  348 


SUBJECT  INDEX 


1135 


benzoate,  348 

phenylacetate,  348 

jS-phenylpropionate,  348 
glutaminase,  332 
AMP,  see  Adenosineraonophosphate 
Amphetamine  (Benzedrine), 
dopamine  /3-hydroxylase,  320 
phenylalanine  /S-hydroxylase,  600 
Amylamine,  diamine  oxidase,  361-362 
Amylase,  succinyl  peroxide,  694 
a-Amylase, 
analogs,  420 

bios3mthesis  of,  D-asparagine,  269 
cystine,   662 
dehydroacetate,  622 
ferricyanide,  674 
iodine,  683-684 
macroions,  464 
mercurials,  792,  795-797,  813,  833 

pH  effects,  792,  795-797 

rate  of  inhibition,  813 
nitrite,  660 
permanganate,  659 
/S-Amylase, 
analogs,  421 
cystine,  662 
dichromate,  660 
ferricyanide,  673-675 
GSSG,  662 

hydrogen  peroxide,  692 
iodine,  683-684 
o-iodosobenzoate,  703,  706,  716,  718 

failure  to  reverse,  718 

kinetics,  716 

pH  effects,  716 
macroions,  464 
mercurials,  772,  803,  833 

relation  to  SH  groups,  803 

type  of  inhibition,  772 
nitrite,  660 
permanganate,  659 
Amylo- 1 ,6-glucosidase,  o-iodosobenzoate, 

706,  718 
Amylomaltase   (maltose   4-glucosyl trans- 
ferase), 

analogs,  415,  422 

mercurials,  protection  by  maltose,  780 
Amylopectin  sulfate,  hyaluronidase,  459 


Anaerobic  glycolysis,  see  Glycolysis  (an- 
aerobic) 
Analog   inhibition, 

annual  number  of  publications,  260 

development  of  concept,  259-261 

expression  of  results,  252-255 

kinetics,    248-252 

mechanisms  of,   246-248 

membrane  transport,  261-268 

substrate   isomers,    268-274 
Analogs,  see  also  specific  substances 

definition  of  different  types  of,  257 

formed  by  fluorine  substitution,  258-259 

metabolism    to   active   inhibitor,    247, 

251-252,  258 

molecular  alterations  producing,   255- 

259 
5a-Androstane-3,17-dione,     J*-3-ketoste- 

roid  reductase,  449 
Androst  - 1  -  ene  -  3, 17  -  dione,   A  *  -  3-ketoste- 

roid  reductase,  449 
Anemonin,  617 
1,5-Anhydro-D-glucitol,  glucose  transport 

by  intestine,  264 
l,5-Anhydro-D-glucitol-6-phosphate, 

glucose-6-P  dehydrogenase,  379 

hexokinase,  377-380 

hydrolysis    by    glucose-6-phosphatase, 

379 

phosphoglucomutase,  379 

phosphoglucose  isomerase,   379 

respiration    (glucose),    379 

structure  of,  378 
Aniline,  D-amino  acid  oxidase,  348 
Anodonta  eggs,  mercurials,  965 
Anomers,   as  analog   inhibitors,   271-272 
Anthranilate, 

derivatives   of,    tryptophan   synthesis, 

321 

3-hydroxyanthranilate  oxidase,   594 

kynurenine  formamidase,  595 

metabolism  of,  tryptophan  analogs,  321 
Anthriscus,  malonate  occurrence  in,  225 
Antimetabolite,  definition  of,  246 
Antimycin  A,  gastric  acid  secretion,  915- 

916 
Antiseptic  mercurials,  see  also  Mercurials 

structures  of,  970 


1136 


SUBJECT  INDEX 


Aorta,  respiration  (endogenous), 
malonate,  179 

Apium,  malonate  occurrence  in,  225 

Aplysia  depilans  (sea  hare)  muscle, 
respiration  (endogenous),  malonate,  174 
succinate  oxidation,  malonate,  54 
effect  of  fumarate  on,  114 

Apparent   inhibitor   constant,    definition 
of,  252 

Apple,  see  also  Malus 

Apple  fruit, 

citrate  oxidation,   malonate,   79 
a-ketoglutarate  oxidation,  malonate,  80 
malate  oxidation,  malonate,  82 
respiration  (endogenous),  malonate,  173 
succinate  oxidation,  malonate,  53 

Apple  skin,  respiration  (endogenous), 
resorcinol,  296 

Apyrase, 

o-iodosobenzoate,  707 
mercurials,  860 
quinacrine,    549 

D-Arabinose, 

/S-galactosidase,  418 

glucose    uptake   by   lymph   node,  263 

a-glucosidase,    423 

a-mannosidase,   422 

phosphoarabinose   isomerase,   411 

uptake  by  heart,  glucose,  263 

L-Arabinose,    phosphoarabinose    isomer- 
ase, 411 

L-Arabinose  dehydrogenase,  ferricyanide, 
675 

Arabinose-5-phosphate,   hexokinase,   379 

Arabinose-phosphate  isomerase,  see  Phos- 
phoarabinose isomerase 

Arabitylflavin,    phosphorylation   of,    539 

Arabono-l,4-lactone,  a-glucuronidase,  426 

Arbacia  punctulata  aggs, 
cleavage, 

malonate,  117,  198 
mercurials,  963-965 
differentiation, 

o-iodosobenzoate,  726 
malonate,  198 

Arbacia  punctualata  spermatozoa, 
motility,  mercurials,  964 


respiration    (endogenous),    mercurials, 

882 
Arcaine, 

diamine  oxidase,  364 

structure  of,  361 
Arginase  (L-arginine  ureohydrolase), 

analogs,  335-338 

permanganate,  660 
Argininate,  carboxypeptidase  B,  367 
L-Arginine, 

dipeptidase,  367 

homoserine  kinase,  357 

uptake  by  ascites  cells,  ornithine,  338 
Arginine   decarboxylase, 

mercurials,    protection    by    substrate, 

779-780 

toxopyrimidine,  578 
Arginine    deiminase    (L-arginine    imino- 

hydrolase), 

canavanine,  353 

mercurials,  833 
L-Arginine  iminohydrolase,  see  Arginine 

deiminase 
Arginine  kinase, 

o-iodosobenzoate,  707 

mercurials,  833 

nucleotides,  467 
L-Arginine  ureohydrolase,  see  Arginase 
Argon,  nitrogen  fixation,  291 
Arsenate, 

phosphatases,  439-441 

phosphate  transport  in  S.  aureus,  267 
Arsenite,  porphyrin  biosynthesis,  162 
Arsenite  oxidase,   o-iodosobenzoate,   707 
Arsonoacetate, 

intercharge  distance,  7 

ionization  constant,  242 

succinate  dehydrogenase,  243 
Arum  maculatum  (cuckoo  pint)  spadix, 

cytochrome  oxidase, 
malonate,  60 
mercurials,  836 

respiration  (endogenous),  malonate,  171 

182 

succinate  dehydrogenase, 
malonate,  19,  28 
mercurials,  855 

succinate  oxidase,  malonate,  19 


SUBJECT  INDEX 


1137 


Arylesterase  (aryl  ester  hydrolase),  mer- 
curials, 834 

Arylsulfatase  (aryl  sulfate  sulfohydrolase), 
analogs,  443-444 
ferricyanide,  675 
iodine,  684 
mercurials,  816 

Arylsulfatase  b, 

o-iodosobenzoate,  707 
mercurials,  816 

Ascaridia  galli, 

a-ketoglutarate  oxidation,  malonate,  80 
respiration  (endogenous),  malonate,  174 
succinate  accumulation,  malonate,  94 
succinate  oxidation,  malonate,  54 

Ascaris  muscle, 

succinate  decarboxylation  to  propio- 
nate, malonate,  165 

Ascites  carcinoma  cells, 

acetate  oxidation,  2-deoxyglucose,  397 
amino  acid   uptake,   analogs,   265-266 
arginine  uptake,  ornithine,  338 
citrate  levels,  2-deoxyglucose,  399 
fructose- 1,6-diP   levels,  hydrogen  per- 
oxide, 695 

fructose  utilization,  galactose,  263 
glucose  utilization,  oxamate,  435 
glutamate  metabolism,  malonate,   152 
glycine  uptake,   malonate,   155 
glycolysis, 

2-deoxyglucose,  392-393 
ferricyanide,  677 
hydrogen   peroxide,   695 
malonate,  126,  209 
mercurials,  875 
oxamate,  434 
growth  of,   2-deoxyglucose,   400 
hydrogen  peroxide  formation  in,   695 
K+  uptake,   mercurials,   908 
malonate  inhibition, 

citrate  accumulation,   104 
a-ketoglutarate  accumulation.  111 
succinate  accumulation,  91 
NAD  levels,  hydrogen  peroxide,  695 
Na+  transport,  malonate,  209 
oxidative   phosphorylation,    malonate, 
122 
phosphate    incorporation    into    ADP- 


ATP,  2-deoxyglucose,  395 

protein  biosynthesis, 

o-fluorophenylalanine,  351 
malonate,  156 

pyridoxine  levels,  deoxypyridoxol,  568 

pyruvate  oxidation, 
2-deoxyglucose,  397 
D-glucosamine,  383 
oxamate,  434 

pjTuvate    utilization,    2-deoxyglucose, 

399 

respiration  (endogenous),  malonate,  177 

respiration   (glucose), 
o-iodosobenzoate,  722 
malonate,  125,  209 
mercurials,  883 

transplantability,    hydrogen   peroxide, 

695 

valine  incorporation  into  proteins,  2- 

deoxyglucose,  399 
Asclepain  m,  iodine,  684 
Ascophyllum  nodosum, 

iodide  uptake,  malonate,  209 

respiration  (endogenous),  malonate,  169 
Ascorbate  oxidase, 

hydrogen  peroxide,  692 

malonate,  21 

mercurials,  protection  by  amino  acids, 

778 

tetrathionate,  698 
Ashbya  gossyjni, 

acetate  oxidation,  malonate,  77,  87 

citrate  accumulation,  malonate,  104 

respiration  (endogenous),  malonate,  168 
L-Asparaginase    (L-asparagine    amidohy- 

drolase,  L-asparagine  deamidase), 

D-asparagine,  269 

hydrogen  peroxide,  692 

mercurials,  834 
D-Asparagine, 

a-amylase  biosynthesis,  269 

L-asparaginase,  269 
L-Asparagine  deamidase,  see  Asparaginase 
Asparagine:a-ketoglutarate  transaminase, 

see  Transaminases 
Asparagine:   pyruvate  transaminase,   see 

Transaminases 
Aspartase    (L-aspartate    ammonia-lyase), 


1138 


SUBJECT  INDEX 


analogs,  355 
iodine,  685 
o-iodosobenzoate,  707 
mercurials,  834 

Aspartate, 

cysteine  desulfurase,  357 
glutamate  decarboxylase,  327-328 
succinate  dehydrogenase,  36 

D-Aspartate, 
aspartase,  355 
L-glutamate  dehydrogenase,  332 

L- Aspartate, 

L-amino  acid  oxidase,  340 
arginase,  337 

L-glutamate  dehydrogenase,  332 
phosphatase  (acid),  441 
pyridoxamine  -  oxalacetate    transamin- 
ase, 600 

Aspartate  carbamyltransferase, 
analogs,  467-469 
o-iodosobenzoate,  707 
mercurials,  834 

protection  by  substrates,  779-781 
stimulation,  816 

Aspartate:a-ketoglutarate    transaminase, 
see  Transaminases 

Aspartate  kinase,  analogs,  356 

Aspartate  transcarbamylase,   see   Aspar- 
tate carbamyltransferase 

Aspergillus, 

citrate    accumulation,   hydrogen    per- 
oxide, 694 
succinate  accumulation,  malonate,  90 

Aspergillus  glaucus,  resistance  to  mercu- 
rials, 983 

Aspergillus  niger, 

citrate  accumulation,  ferrocyanide,  677 
growth  of, 

dehydroacetate,  632-633 
ferrocyanide,  677 
malonate,  195 
tungstate,  614 
malonate  formation  in,  226 
malonate  metabolism  in,  228 
respiration   (sucrose),    mercurials,    880 
sporulation,  malonate,   195 
succinate  dehydrogenase,  malonate,  27 

Aspergillus  oryzae. 


2-deoxyglucose  utilization,  387,  400 
succinate  dehydrogenase,  malonate,  27 

Aspergillus  tereus, 

glucose  utilization,  ferrocyanide,  678 
itaconate  formation,  ferrocyanide,  678 

Atabrine,  see  Quinacrine 

Atebrin,   see   Quinacrine 

ATP,  see  Adenosinetriphosphate 

ATP-ADP  exchange  reaction, 
ADP,  444 
mercurials,  874 

ATP:AMP  phosphotransferase,  see  Ade- 
nylate kinase 

ATP:creatine      phosphotransferase,      see 
Creatine  kinase 

ATP  diphosphohydrolase,  see  Apyrase 

ATP-P,  exchange  reaction, 
ADP,  444 

o-iodosobenzoate,  705 
mercurials,  872-873 

Atrium,   see   Heart 

Australorbis  glabratus,  respiration  (endo- 
genous), 

iraws-aconitate,    273 
malonate,  174 
mercurials,  882 

Auxin,  transport  of, 
mercurials,  967 

Avena  sativum  coleoptile, 

fumarate  oxidation,  malonate,  81 
growth  of, 

malonate,    196-197 
mercurials,  966-968 
indoleacetate  uptake,   mercurials,   911 
malonate  metabolism  in,  228 
respiration  (endogenous),  malonate,  167 
169,  182,  184 

succinate  accumulation,  malonate,  93 
succinate  oxidation,  malonate,  53 

Avocado, 

cis-aconitate  oxidation,  malonate,  79 
citrate  oxidation,   malonate,   78 
fumarate  oxidation,  malonate,  81 
a-ketoglutarate  oxidation,  malonate,  79 
malate  oxidation,  malonate,  82 
malonate  formation  in,  226 
pyruvate  oxidation,  malonate,  74 
succinate  accumulation,  malonate,  91 


SUBJECT  INDEX 


1139 


Avocado  mesocarp,  fatty  acid  formation 

from  acetate, 

malonate,  148 
2-Azaadenine, 

structure   of,   280 

xanthine  oxidase,   280-282 
8-Azaguanine, 

adenosine  deaminase,  466,  477 

deamination  to  8-azaxanthine,  478 

enzyme  induction  in  liver,  478 

guanine  deficiency,  281 

protein   biosynthesis,   478 

structure  of,  280 

xanthine  oxidase,  281-282,  477 
8-Azaguanosinetriphosphate, 

competition  ^vith  GTP,  478 

formation  from  8-azaguanine,  478 
5-Azaorotate,  orotate  conversion  to  oro- 

tidylate,  480 
8-Azapurines,  oxidation  by  xanthine  oxi- 
dase, 281 
Azaserine, 

formylglycinaniide  ribonucleotide  ami- 

dotransferase,    333 

glutaminase,  356 

inosinate   biosynthesis,   333 

phosphoribosyl-PP  amidotransferase, 

333 

structure  of,  333 
Azatryptophan, 

/S-galactosidase  synthesis,  326 

E.  coli  growth,  326 

incorporation  into  enzymes,  326 

L-tryptophan:sRNA  ligase  (AMP),  326 
6-Azauracil,  carcinostasis,  478 
6-Azauridine, 

conversion  to  6-azauridylate,  480 

orotidylate  decarboxylase,  472 
6-Azauridinediphosphate,   polynucleotide 

phosphorylase,  474 
6-Azauridinemonophosphate, 

orotidylate  decarboxylase,  472,  478 

polynucleotide  phosphorylase,  474 
8-Azaxanthine, 

adenosine  deaminase,  466 

xanthine  oxidase,  282 
Azelaeate,     kynurenine  :  a  -  ketoglutarate 

transaminase,  608 


Azide,  renal  transport  of  PAH,  205 
Azotohacter  agilis,   malonate  metabolism 

in,  227-228 
Azotohacter  vinelandii, 

growth  of,  tungstate,  614 

nitrogen  fixation,  gas  analogs,  292-294 

succinate  dehydrogenase,  malonate,  26 

B 

Bacillus  anthracis, 

growth  of,  dehydroacetate,  632 

mercurial  uptake,  975 
Bacillus  brevis,  motility  of, 

o-iodosobenzoate,   727 

malonate,  203 
Bacillus  cereus, 

acetate    utilization,     malonic    diethyl 

ester,  237 

germination  induced  by  L-alanine,  d- 

alanine,  270 

growth  of,   dehydroacetate,   632 

malonic  ethyl  esters,  236 
Bacillus  coagulans,  germination  of, 

quinacrine,  546 
Bacillus  megaterium,  growth  of, 

dehydroacetate,  632 
Bacillus  mesentericus,  growth  of, 

dehydroacetate,  632 
Bacillus  metiens,  iodine  killing  of,  690 
Bacillus  pumilus,  y-aminobutyrate  oxida- 
tion, 

malonate,  154 
Bacillus  pyocyaneus,  o-iodosobenzoate  kil- 
ling of,  727 
Bacillus  suhtilis, 

aminopterin  uptake,  mercurials,  911 

germination  of, 
malonate,  195 
mercurials,  972 
quinacrine,  546 
growth  of, 
dehydroacetate,  632 
malonate,  195 
trifluorothiamine,  531 

succinate  oxidation,  malonate,  52 
Bacteria, 

growth  of. 


1140 


SUBJECT  INDEX 


cyanocobalamin  analogs,  589-590 
D-cycloserine,  359-360 
dehydroacetate,  631-633 
desmethyldesthiobiotin,  588 
iodine,  690 
ion  antagonisms,  452 
kojic  acid,   349 
malonate,  195 
mercurials,  774,  970-976 
pantothenate  analogs,  587 
pyridine-3-sulfonate,  504 
pyridoxal  analogs,  575-576 
riboflavin  analogs,  537-538 
thiamine  analogs,  528-530 
unsaturated  lactones,  617-618 
infections  by,  malonate,  221-224 
Bacterium  lactis,  respiration  (glucose), 

mercurials,  880 
Bacterium  succinicum , 

acetate  oxidation,  malonate,  77 
pyruvate  oxidation,  malonate,  74 
succinate  oxidation,  malonate,  51 
Balantidium  coli, 

motility,  malonate,  203 
respiration  (endogenous),  malonate,  173 
Balanus   balanoides,    mercurial    toxicity, 

961-962 
Barley,  malonate  occurrence  in,  224 
Barley  roots, 

bromide  uptake,  malonate,  116-117 
ion  transport,  fraws-aconitate,  274 
K+  uptake,  malonate,  209 
phosphate  uptake  in  mitochondria,  ma- 
lonate, 122 

Rb+  uptake,  malonate,  209 
respiration     (endogenous),     malonate, 
170,  181,  210 

respiration  (ion-linked),  <raws- aconitate, 
273 

respiratory  quotient,  malonate,  185 
succinate  accumulation,  malonate,  91 
succinate  dehydrogenase,  malonate,  27 
succinate  oxidation,  malonate,  51 
Beech  (Fagus)  roots, 

respiration  (endogenous),  malonate,  172 
succinate  oxidation,  malonate,  51 
Bees,  mannose  toxicity,  414 
Beets,   succinate   dehydrogenase, 


malonate,  27 

Benzamide, 

ammonia  formation  in  kidney,  348 
chymotrypsin,   372 
glucose  dehydrogenase,  501-502 
lactate   dehydrogenase,   501-502 
NAD  nucleosidase,  488 
tyrosinase,  300-301 

Benzedrine,  see  Amphetamine 

Benzenesulfonate, 
arylsulfatase,  444 
glucose  dehydrogenase,  501-502 
lactate   dehydrogenase,   437-438,   501- 
502 

sulfite  oxidase,  451 
tyrosinase,  300-301 

Benzimidazole,  NAD  nucleosidase,  492 

Benzoate, 

acetate  accumulation  in  Proteus,  349 
acetoacetate  formation  from  butyrate, 
613 

amino  acid  deamination  in  kidney,  348 
D-amino  acid  oxidase,  340-348 
L-amino  acid  oxidase,  338,  348 
p-aminobenzoate  acetylation,  349 
ammonia  formation  in  kidney,  348 
carboxypeptidase,  306 
catechol  oxidase,  297-302 
chymotrypsin,  370,  372 
citrate  oxidation,   348 
crotonate  oxidation,  349 
dehydroshikimate  reductase,  606 
dopa  decarboxylase,  312 
fatty  acid  oxidation,  349 
glucose  dehydrogenase,  501 
glucose  metabolism,  349 
glutamate  dehydrogenase,  331 
D-glutamate  oxidase,  349 
a-ketoglutarate  oxidation,   348 
a-ketoisocaproate  decarboxylase,  349 
kynurenine:a-ketoglutarate     transami- 
nase, 608-609 

lactate    dehydrogenase,    501 
NADPH  dehydrogenase,  349 
oxidative  phosphorylation,  348 
phosphorylation  in  mitochondria,  349 
pyruvate  oxidation,  349 
respiration  (endogenous),  348 


SUBJECT  INDEX 


1141 


respiratory  quotient,  349 

shikimate  dehydrogenase,  349 

succinate  oxidation,  348 

tyrosinase,  300-301,  349 

yeast  growth,   349 
Benzoate  methyl  ester,  catechol  oxidase, 

298 
p-Benzoquinone,       Fusarium       conidial 

growth,   660 
Benzoyl-L-argininamide,  papain,  375 
Benzoyl-L-arginine,  papain,  375 
iV-Benzoyl-a-D-glucosamine,  structure  of, 

378 
Benzoyl-D-tryptophanamide,  chymotryp- 

sin,  371 
Benzoyl-L-tyrosinemethylamide,    chymo- 

trypsin,  371 
Benzylmalonate,  carboxypeptidase,  367 
a-Benzylmalonamide,  chymotrypsin,  370 
Benzyloxyamine,    dopamine/J-hydroxyl- 

ase,  320 
Benzylsulfate,   arylsulfatase,   443 
Betaine,     thetin.-homocysteine    transme- 

thylase,  356 
Betaine  aldehyde  dehydrogenase, 

o-iodosobenzoate,  707 

mercurials,  781 

protection  by  NAD,  781 
protection  bj'  substrate,  781 
Betaine:homocysteine  methyltransferase, 

dimethylglycine,  356 
Bicarbonate,  renal  transport  of, 

mercurials,  920-921 
Bioluminescence,  mercurials,  888-891 
Biotin, 

analogs  of, 

biotin  degradation,  589 
biotin  oxidase,  589 
E.  coli  growth,  588 
fermentation,  588-589 

biotin  oxidase,  589 

intestinal  transport  of,  analogs,  267 

structure  of,  588 
Biotin-diaminecarboxylate,     biotin     oxi- 
dase, 589 
Biotinol,  biotin  oxidase,  589 
Biotinol-diamine,  biotin  oxidase,  589 


Biotin  oxidase,  analogs,  589 
Biotinsulfone.  biotin  oxidase,  589 
Bis  (p-nitrophenyl)  disulfide,  determina- 
tion of  SH  groups,  640-641 
Blastocladiella  emersonii,  respiration  (en- 
dogenous), 

malonate,  169 
Blastocysts, 

glycolysis,  2-deoxyglucose,  393 

respiration  (endogenous),  malonate,  179 

respiration    (glucose),    2-deoxyglucose, 

393 
Blebbing  of  tumor  cells,  see  Sarcoma  37 
Blood, 

acetoacetate   accumulation,   malonate, 

138,  149 

cholesterol  level,  malonate,  150 

coagulation  of, 

o-iodosobenzoate,  725 
malonate,  219 

CO2  capacity  of,  malonate,  219 

glucose  level,  malonate,   149,  219 

glycolysis,  mercurials,  877 

K+  level,  malonate,  206 

lactate  level, 
malonate,  219 
thiamine  analogs,  520 

malonate  levels  in  vivo,  100-103 

mercurial  levels  in  vivo,  930,  958-960 

Na+  level,  malonate,  206 

pyruvate  level, 
malonate,   219 
pyrithiamine,  520,  527 

succinate  level,  malonate,  100,  102 
Blood  pressure, 

o-iodosobenzoate,   723 

malonate,  213 

pyrogaUol,  611 
Blood  vessels,  see  Vascular  smooth  muscle 
Blowfly,  see  Calliphora 
Bolaform  ions,  5,  44 
Bone,  growth  of  cultures, 

malonate,  199 
Bone  marrow,  nucleic  acid  biosynthesis, 

folate  analogs,  585 
Borate,  phosphatases,  439-440 
Borneol-a-glucuronide,     a-glucuronidase, 

426 


1142 


SUBJECT  INDEX 


Botrytis  allii,  growth  of, 

dehydroacetate,  632 
Botrytis  fabae,  histidine  transport, 

analogs,  267 
Brain, 

^-acetylaspartate  formation,  malonate, 

154 

acetylcholine  S5m thesis,  malonate,  165- 

166 

ADP  level,  2-deoxyglucose,  395 

amino  acid  decarboxylase,  a-methyl-m- 

tyrosine  in  vivo,  317 

amino  acid  metabolism,  malonate,  153 

amino  acid  synthesis  from  glucose,  2- 

deoxyglucose,  399 

amino  acid  transport,  analogs,  266-267 

y-aminobutyrate  level,  toxopyrimidine, 

578 

y-aminobutyrate  oxidation,  malonate, 

154 

6-aminonicotinamide-NAD     formation 

in,  505 

ATP  level,  2-deoxyglucose,  395 

C-l/C-6  ratio, 

2-deoxyglucose,  393-394 

malonate,  130 
catecholamine  levels, 

a-methyldopa,  317-318 

a-methyl-m-tyrosine,  316-318 
citrate  level,  sequential  inhibition  by 
malonate  and  fluoroacetate,   112 
creatine-P  level,  2-deoxygluco8e,  395 
cycle  intermediates  levels,  89 
dopamine   level,    a-methyl-w-tyrosine, 
316 

glucose  uptake,  2-deoxyglucose,  394 
glucose  utilization,  malonate,  126-127, 
134-135 
glutamate  decarboxylase  in  vivo, 

deoxypyridoxol,  569-570 

toxopyrimidine,  578 
glutamate  formation  from  glucose,  ma- 
lonate, 153 

glutamate  uptake,  malonate,  153 
glutamine  level,  2-deoxyglucose,  399 
glycolysis  (aerobic),  malonate,  127-128, 
134-135 
glycolysis  (anaerobic), 


dehydroacetate,  624 

D-glucosone,  385 
a-ketoglutarate    oxidation,    malonate, 
80-81,   84 
K+   uptake, 

2-deoxyglucose,  399 

malonate,  153 
lactate  oxidation, 

nicotinamide,  500 

nicotinate,  500 
malonate  decarboxylation  in,  232 
malonate  level  in  vivo,  102 
mercurial  levels,   958-959 
NAD  level,  6-aminonicotinamide,  505 
norepinephrine  level, 

a-methyl-m-tyrosine,  316 

pyrogallol,  611 
oxalacetate  oxidation,  malonate,  82 
oxidative  phosphorylation,  mercurials, 
873 

phospholipid    biosynthesis,    malonate, 
151 

pjrridoxal-P  level,  deoxypyridoxol,  567- 
569 

pyrithiamine  levels  in  vivo,  528 
pyruvate  oxidation,   malonate,   75-76, 
128 
pyruvate  utilization, 

malonate,   135 

oxygen,  658-659 
respiration  (endogenous), 

dehydroacetate,  623-624 

malonate,    175-177,    179,    181,    183- 

184 
respiration  (glucose), 

hydrogen  peroxide,  695 

kojic  acid,  350 

malonate,  124,  127,  133-135 

nicotinamide,  500 

nicotinate,  500 

oxygen,  658-659 

quinacrine,  560 

tartronate,  238 
respiration  (glutamate),  malonate,  152 
serotonin    level,    a-methyl-m-tyrosine, 
316 

succinate  accumulation,  malonate,  95, 
97 


SUBJECT  INDEX 


1143 


succinate  dehydrogenase, 
maleate,  36 
malonate,  31-32 
oxalacetate,  36 
succinate  levels  in  vivo,  malonate,  102 
succinate  oxidation, 
hydrogen  peroxide,  695 
malonate,  55 
thiamine-PP  level,  oxythiamine,  526- 
527 

Branched  chain  ketonuria  (maple  sugar 
urine  disease),  role  of  glutamate  de- 
carboxylase inhibition,  329 

Brevibaderium  flavum,   respiration   (glu- 
cose), 
mercurials,  880 

Bromate,  nitrite  oxidation,  450 

Bromelain, 
iodine,  685 

mercurials,  792,  810,  813 
pH  effects,  792 
rate  of  inhibition,  810,  813 

Bromide, 

^-ketoadipate   chlorinase,   453 

tyrosinase,  300 

uptake  by  barley  roots,  malonate,  116- 

117 

m-Bromobenzoate,  glutamate  dehydro- 
genase, 330 

Bromobenzoates,  D-amino  acid  oxidase, 
341 

5-Bromodeoxycytidine,  aspartate  carba- 
myltransferase,  469 

5-Bromofuroate,  glutamate  dehydrogen- 
ase, 330-331 

^-Bromopropionate,  structure  of,  41 

Bromopyruvate, 

glycerate  dehydrogenase,  430 
lactate  dehydrogenase,  437 

Bronchopneumonia  virus,  inactivation  by 
mercurials,  977 

Brucella  abortus,  respiration  (endogenous), 
malonate,  168 

Bunias  orientalis,  malonate  occurrence  in, 
225 

Bush  bean,  see  Phaseolus  vulgaris 

2,3-Butanedione,  pyruvate  decarboxy- 
lase, 431 


1 ,4-Butanediphosphonate, 

ionization  constants,  242 

succinate  dehydrogenase,   243 
a-ButylglucopjTanoside,    a-glucosidase, 

423 
2'-Butylthiamine,  thiamine  kinase,  523 
Butynamine  demethylase,  2,4-dichloro-6- 

phenylphenoxyethylamine,  592 
Butyrate, 

D-amino  acid  oxidase,  343 

conversion    to    acetoacetate,    analogs, 

613 

glutamate  decarboxylase,  328 

homoserine  kinase,  357 

kynurenine:a-ketoglutarate    transami- 
nase, 608 

lactate  dehydrogenase,   436 

leucine  decarboxylase,  352 

tyrosinase,  300 
Butyryl-CoA  dehydrogenase,  mercurials, 

781 

protection  by  FAD,  781 

protection  by  substrate,  781 


Cadaverine, 

kynureninase,  595 

oxidation  of,  analogs,  363 

structure  of,  361 
Cadaverine  oxidase,  see  Diamine  oxidase 
Cadmium,   complexes   with   di-  and   tri- 

carboxylates,   12 
Caffeate    (3,4-dihydroxycinnamate), 

acetylindoxyl  oxidase,  591 

dopa  decarboxylase,  314 

glutamate  decarboxylase,  314 

histidine  decarboxylase,  314 

indoleacetate  oxidase,  595 

peroxidase,  599 

succinate  dehydrogenase,  314 

tyrosinase,  314 

tyrosine  decarboxylase,  314 
Caffeine, 

D-araino  acid  oxidase,  545 

NAD  nucleosidase,  492 
Calcium, 

ATPase,  453 

intestinal  transport  of. 


1144 


SUBJECT  INDEX 


malonate,  208 
mercurials,  909,  913-914 
oxidative  phosphorylation,  453 
phosphorylase  b  kinase,  453 
uptake    by    mitochondria,    mercurials, 
909 
urinary  excretion,  mercurials,  921 

Caldariomyces  fumago,   respiration   (glu- 
cose), 
malonate,  125 

Calliphora  erythrocephala, 

a-ketoglutarate    oxidation,    malonate, 

80,  84 

oxidative   phosphorylation,    malonate, 

121 

Camellia  pollen,  malonate  metabolism  in, 
228 

Canavanine,  arginine  deiminase,  353 

Candida  albicans,  growth  of, 
dehydroacetate,  632 

Candida  utilis,  resistance  to  mercurials, 
983-984 

Cape  barley,  see  Hordeum  vulgare 

Caprate  (decanoate),  kynurenine: a-keto- 
glutarate transaminase,  595,  608 

Caproate   (hexanoate), 

glutamate  decarboxylase,  328 
kynurenine:a-ketoglutarate     transami- 
nase, 608 

Caprylate  (octanoate), 

kynurenine:  a-ketoglutarate     transami- 
nase, 608 
lactate  dehydrogenase,  436 

Carbamate  kinase,  see  Carbamyl-phos- 
phate  synthetase 

Carbamyl-ADP  phosphotransferase,  mer- 
curials, 834 

Carbamyl-phosphate  :  L-aspartate  carba- 
myltransferase,  see  Ornithine  carbamyl- 
transferase 

Carbamyl-phosphate  synthetase,  mercu- 
rial, 781,  804 

protection  by  acetylglutamate,  781 
relation  to  SH  groups,  804 

Carbobenzoxy-L-glutamate,   papain,   375 

Carbohydrate  transport,  see  also  specific 
sugars 
analogs,  262-264 


Carbon  dioxide,  fixation  of, 

mercurials,  892 
Carbon  dioxide  activating  enzyme,  mer- 
curials, 834 
Carbonic  anhydrase, 

ferricyanide,  675 

hydrogen  peroxide,  692 

iodine,  685 

mercurials,  834-835 

p-mercuribenzoate,  competitive  nature 

of  inhibition,   771-772 
Carbon  monoxide,  nitrogen  fixation,  291- 

292 
Carboxylate   group,    interaction   energy, 

276 
Carboxylesterase,  see  Aliesterase 
Carboxymethylcellulose,   chymotrypsin, 

457 
Carboxypeptidase, 

active  center  of,  365-366 

analogs,  365-367 

malonate,  60 

mercurials,  770,  781,818 
protection  by  Zn++,  781 
stimulation,   818 
Carboxypeptidase  A,  mercurials, 

failure  of  Zn++  to  reverse,  827 
Carcinoma,    see    also    Ascites    carcinoma 

cells.  Walker  carcinoma 

growth  of,   2-deoxyglucose,   400 

succinate  dehydrogenase,  malonate,  32 
Carcinostasis,  see  Tumors,  growth  of 
Carcinus  maenas,  oxidative  phosphoryla- 
tion, 

malonate,  121 
Caries,  from  dehydroacetate,  630-631 
Carnation  seeds,  growth  of, 

mercurials,  965 
Carotid  artery,  contractility, 

malonate,  212 
Carotid  body,  excitability, 

malonate,  211-212 
Carp  liver, 

fatty  acid  oxidation,  malonate,  137 

a-ketoglutarate  oxidation,  malonate,  80 

oxidative   phosphorylation,    malonate, 

121 

pyruvate  oxidation,  malonate,  75 


SUBJECT  INDEX 


1145 


succinate  accumulation,  malonate,  91 
Carrots, 

glucose  uptake,  2 -deoxy glucose,  394 

isocitrate  oxidation,  malonate,  79 

malate  oxidation,  malonate,  82 

pyruvate  oxidation,  malonate,  74 

respiratory  quotient,  malonate,   185 

succinate  accumulation,  malonate,  97 

succinate  oxidation,  malonate,  22 
Catalase, 

cystine,  662-663 

dehydroacetate,  623 

induction  of,  8-azaguanine,  478 

mercurials,   835 

rate  of  inhibition,  810 
relation  to  SH  groups,  803 

monoethylperoxide,  592,  694 
Catechol, 

acetylindoxyl  oxidase,  591 

D -amino  acid  oxidase,  344 

dehydroshikimate  reductase,  593,  605 

histidine  decarboxylase,   352 
Catecholamines,  see  also  specific  amines 

metabolism  of,  pyrogallol,  611-612 

release  of, 

o-iodosobenzoate,  725 
mercurials,  947 
a-methyldopa,  315-320 
a-methyl-TO-tyrosine,  315-316 
Catechol-0-methyl  transferase, 

effect  of  inhibition   on   catecholamine 

metabolism,  611-612 

pyrogallol,  592,  611-612 
Catechol  oxidase, 

analogs,  296-298 

benzoate, 

derivatives  of,  297-302 
pH  effects,   298-299 

cinnamate,   298 

dihydroxymaleate,   297 

quinacrine,  547,  549 
Cathepsin, 

iodine,  685,  688 

o-iodosobenzoate,   707 
Cathepsin  C,   analogs,  375 
Cauliflower  buds, 

citrate  oxidation,  malonate,  78,  87 

a-ketoglutarate  oxidation,  malonate,  79 


Cecropia  silkworn,  spermatid  meiosis  in, 

malonate,  199 
Cell  division,  see  also  specific  organisms  or 

tissues 

2-deoxyglucose,  400 

heparin,  462 

o-iodosobenzoate,  726-727 

malonate,  197-200 

mercurials,  963-970 
Cellobiase,  mercurials,  860 
Cellobiose, 

)S-glucosidase,  417 

a-mannosidase,  422 
Cellulose  polysulfatase,  ferricyanide,  675 
Cellulose,  biosynthesis  of, 

malonate,  132 
Central  nervous  system,  see  Brain 
Cephalosporin  C,  penicillinase,  599 
C-esterase,  mercurials,  816 
Chaetonium  globosum,  growth  of, 

dehydroacetate,  633 
Chaetopterus  eggs,  cleavage, 

malonate,  198 

mercurials,  965 
Chelidonate,  D-amino  acid  oxidase,   342 
Chick  embryo,  see  also  Embryogenesis 

development  and   differentiation, 
3-acetylpyridine,  494 
malonate,  199 

glucose  uptake  by  heart,  2-deoxyglu- 
cose, 394 

growth  of  heart  fibroblasts,  2-deoxyglu- 
cose, 400 

protein   biosynthesis   in   cultures,  ma- 
lonate, 156 

respiration   (endogenous)   of  condyles, 

malonate,   175 

respiration   (endogenous)  of  yolk  sac, 

malonate,  175 

succinate  oxidation  in   cartilage,   ma- 
lonate, 54 
Chicory, 

respiration  (endogenous)  malonate,  173, 

182-183 

respiratory  quotient,  malonate,  185 
Chitin  disulfate,  ^-fructofuranosidase,  465 
Chitin  sulfates,  hyaluronidase,  459 
Chlorate, 


1146 


SUBJECT  INDEX 


nitrite  oxidation,   450 
Nitrobacter  growth,  450 

Chlorella, 

hydrogen  production,  mercurials,  891 

photosynthesis,  malonate,  163 

primary  photogenic  agent,  mercurials, 

892 

succinate  oxidation,  malonate,  51 

Chlorella  pyrenoidosa,  malonate  metabo- 
lism in,  228 

Chlorella  vulgaris, 

respiration  (endogenous) 
malonate,  169 
malonic  diethyl  ester,  237 
mercurials,  881 
respiration  (glucose),  mercurials,  881 

Chloride, 

intestinal  transport  of,  mercurials,  910 
tyrosinase,  300-301 

Chlormerodrin  (Neohydrin),  see  also  Mer- 
curials, kidney 
structure  of,  917 

Chloroacetate, 

pantoate:/S-alanine  ligase,  597 
tyrosinase,  300 

Chloroacetyl-D-tyrosinamide,  chymotryp- 
sin,  371 

Chloroacetyl-L-tyrosinate,  chymotrypsin, 
371 

m-Chlorobenzoate,    glutamate    dehydro- 
genase, 330 

Chlorobenzoates, 

D-amino  acid  oxidase,  341 
catechol  oxidase,  298-299 

6-Chloro-2,8-dihydroxypurine, 
urate  metabolism  in  vivo,  285 
uricase,  284-285 

5-Chlorofuroate,    glutamate    dehydroge- 
nase, 330 

Chlorogenate, 

indoleacetate  oxidase,  595 
peroxidase,  599 

(5-Chlorolevulinate,    aminolevulinate    de- 
hydrase,  591 

2-Chloromethyl-5-hydroxy-l,4-pyrone,  d- 
amino  acid  oxidase,  342 

Chlorophyll,  mercurials,  891 

ChlorophyUase,  ferricyanide,  675 


a-Chloroproprionate ,    pantoate  :/9-  alanine 

ligase,  598 
/3-Chloropropionate,    pantoate :  ^  -  alanine 

ligase,  598 
6-Chloropurine, 

oxidation  to  6-chlorourate,  281,  285 

purine  metabolism  in  vivo,  281 

xanthine  oxidase,  281-282 
Chloropyru  vate , 

lactate  dehydrogenase,   437 

pyruvate  decarboxylase,  431 
4-Chlororesorcinol,  tyrosinase,  302,  304 
6-Chlororiboflavin  analog, 

flavokinase,  539 

Lactobacillus  growth,  537 

riboflavin  deficiency,  538 
7-Chlororiboflavin  analog, 

Lactobacillus  growth,   537 

riboflavin  deficiency,  538 
6-Chlorourate, 

formation    from    6-chloropurine,    281, 

285 

uricase,  284-285 

xanthine  oxidase,  282 
Chloroxythiamine,  structure  of,  517 
Cholestane,  cholesterol  esterase,  592 
Cholestan-3-one,  cholesterol  esterase,  592 
Cholesterol,  see  also  Sterols 

biosynthesis  of, 
malonate,    149-151 
a-phenylbutyrate,  614 
Cholesterol  esterase  (sterol  ester  hydro- 
lase), 

analogs,  592 

mercurials,  835 
Choline,  thetin:homocysteine  transmethy- 

lase,  356 
Choline  acetylase, 

iodine,  685 

o-iodosobenzoate,  707 

mercurials,  835 

spontaneous   reversal   of  inhibition, 
813-814 

pantothenate  analogs,  587 
Choline      dehydrogenase,     tetrathionate, 

698-699 
Choline  oxidase, 

analogs,  290-291 


SUBJECT  INDEX 


1147 


0-iodosobenzoate,  707,  717 
protection  by  choline,  717 

kojic  acid,   350 

malonate,   60 

oxygen  inactivation  of,  659 
Cholinesterase, 

dehydroacetate,  622 

ferricyanide,  675 

GSSG,  662 

iodine,  683,  685 

o-iodosobenzoate,  707 

mercurials,  772-773,  776-778,  788,  810, 

835-836,  937-938 

kinetic  analysis,  776-778 
possible  denaturation,  788 
rate  of  inhibition,  810 
role  in  muscle  action,  937-938 
type  of  inhibition,  772-773 

quinacrine,  549 

succinyl  peroxide,  694 

thiamine  analogs,  531-532 
Choline  sulfa tase, 

phosphate,  444 

sulfate,  444 

sulfite,  444 
Chondroitin  sulfate, 

fumarase,  465 

phosphatase  (acid),  465 

ribonuclease,  462 

synthesis  of,  malonate,  166 

ulcer  reduction,  458 
Chondroitin  sulfate  B,  hyaluronidase,  460 
Chorioallantoic     membrane,     respiration 

(endogenous), 

malonate,  175 
Chorioallantoic   membrane   (virus-infect- 
ed), 

glucose  uptake,  malonate,  126-127 

respiration     (endogenous),      malonate, 

126-127 
Choroid  plexus,  iodide  uptake, 

malonate,  209 
Chromatium,  NAD  iihotoreduction, 

mercurials,  891 
Chromatophores,  migration  in  culture, 

malonate,  203 
Chymotrypsin, 

acridines,  373 


active  center  of,  374 

analogs,  368-374 

hydrogen  peroxide,  694 

iodine,  686 

macroions,  457 

oxidation  of,  657 

periodate,  657 

proteins,  457 

quinolines,  373 

D-tryptophanides,  271 

D-tyrosinamides,  271 
Ciliary  body,  iodide  transport, 

malonate,  209 

nitrate,  267 
Ciliary  motility, 

malonate,  203 

mercurials,  981-982 
Cinnamide,  D-amino  acid  oxidase,  343 
Cinnamate, 

acetoacetate  formation  from  butyrate, 

613 

D-amino  acid  oxidase,  342,  346 

carboxypeptidase,  367 

catechol  oxidase,  298 

derivatives    of,    dopa    decarboxylase, 

311-314 

structure  of,  296 
Citraconate  (methylmaleate), 

fumarase,  279 

glutamate  decarboxylase,  328 

ionization  constants,  8 

structure   of,   279 
Citrate, 

accumulation  of, 

trans-sicomtsite,  273-274 
ferrocyanide,    677-678 
fluoromalonate,  239 
hydrogen  peroxide,  694 
malonate,   104-110,  223 
malonic  diethyl  ester,  236 
mercurials,  927 

formation  of,  malonate,  105,  108 

fumarase,  275 

glutamate  dehydrogenase,  331 

level   in  ascites   cells,   2-deoxyglucose, 

399 

malate  dehydrogenase,  596 

oxidation  of. 


1148 


SUBJECT  INDEX 


benzoate,  348 
malonate,  78-79,  86-87 
mercurials,  878 

phosphofructokinase,  385-386 

urinary   excretion   of,    malonate,    104, 

109-110 
Citrate  (isocitrate)  hydro-lyase,  see  Aco- 

nitase 
Citrate  synthetase   (condensing  enzyme, 

oxalacetate  transacetase), 

malonate,  63 

mercurials,  836 

palmityl-CoA,  614 
CitruUine, 

arginase,  337 

carbamyl  -  P :  ornithine   transcarbamyl- 

ase,  592 
Claviceps  purpurea  (ergot),  succinate  de- 
hydrogenase, 

fumarate  K,,  38 

malonate   K,,   33 
Cleavage,  see  Cell  division 
Clostridium,  nitrogen  fixation, 

hydrogen,  292 
Clostridium  histolyticum ,  growth  of, 

mercurials,  972 
Clostridium   kluyveri,   malonate   metabo- 
lism in,  232 
Clostridium  pasteurianum,  nitrogen  fixa- 
tion, 

nitric  oxide,  292 
Clostridium  saccharobutyricum,  growth  of, 

malonate,  195 
Clostridium  welchii,  growth  of, 

mercurials,  972 
Clover,  malonate  occurrence  in,  224,  226 
Cloxacillin,  penicillinase,  598 
Coagulase,    mercurials,    860 
Cobalt,   complexes   with  di-   and  tricar- 

boxylates,  12 
Cocciodioides  immitis,  spherulation  of, 

mercurials,  971,  973 
Cochliobolus  miyabeanus,  germination  of, 

mercurials,   973 
Cockroach,  see  also  Periplaneta 
Cockroach   muscle,   succinate   oxidation, 

malonate,  22 
Coenzyme  A, 


formation  of,  pathways,  586-587 
kidney  level,  mercurials,  927 
mercurial  complexes,  750 
yeast  level,  mercurials,  750,  885 
Coenzyme  A  analogs,  structures  of,  587 
Coenzyme  analogs,  482-590 

recombination  technic  for  demonstrat- 
ing inhibition,  250 
sites  of  action,   482-483 
Coffee  trees,  mercurial  fungicides  causing 

Zn-deficiency  disease,  966 
Coliphage, 

mercurial  inactivation  of,  977,  980 
proliferation  of, 
malonate,  194 
mercurials,  977,  980-981 
Colpidium,  succinate  accumulation, 

malonate,  91 
Colpidium  campylum, 

pyruvate  oxidation,  malonate,  74 
respiration  (endogenous),  malonate,  173 
Colpidium  colpoda, 

motility  of,  mercurials,  982 
toxicity  of  mercurials,  982 
Common  cold  virus,  infectivity  of, 

mercurials,  977 
Condensing  enzyme,  see  Citrate  synthe- 
tase 
Configurational  changes, 
enzymes, 

alcohol  dehydrogenase,  788-789 
penicillinase,  249,  615,  688 
phosphoribosyl-ATPpyrophosphory- 
lase,  351 

succinate  dehydrogenase,  46 
produced  by, 

mercurials,  787-790 
oxidants,  656 
SH  reagents,  648-650 
proteins,  657,  703-704,  761-762 
Contracture,   see  also   specific   inhibitors 
and  muscles 
heart, 

mercurials,  941-944 
porphyrindin,  669 
intestinal  muscle,  o-iodosobenzoate,  724 
muscle, 

o-iodosobenzoate,  723 


SUBJECT  INDEX 


1149 


mercurials,  896,  938 
tetrathionate,  699 

uterus,  o-iodosobenzoate,  724 
Copper, 

catalysis  of  SH  group  oxidation,  658 

complexes  with  di-  and  tricarboxylates, 

12 

/^-glucuronidase,  795 

toxicity  to  Nitocra,  962 

uptake  by  liver,  mercurials,  910,  913 
Coproporphyrinogen  oxidase,  mercurials, 

860 
Corn  (maize)  roots, 

respiration  (endogenous),  malonate,  172 
pH  effects,  191 

respiratory  quotient,  malonate,  185 
Cornea,  water  transport, 

mercurials,  911 
Coronary  floAv,  malonate,  213 
Corynebacterium,  succinate  oxidation, 

malonate,  52 
Corynebacterium  creatinovorans,  succinate 

accumulation, 

malonate,  92 
Corynebacterium  diphtheriae, 

growth  of,  dehydroacetate,  632 

succinate  dehydrogenase, 
fumarate  K,,  38 
malonate  Kj,  33 
Coumalate  (a-pyrrone-5-carboxylate),  D- 

amino  acid  oxidase,  342 
<rans-23-Coumarate,  phenylalanine  deami- 
nase, 355 
Coxsackie  virus,  infectivity  of, 

mercurials,  977 
Crabtree  effect, 

2-deoxyglucose,  396-398 

D-glucosone,  385 

oxamate,  435 
Crassostrea  virginica  (oyster)  mantle, 

respiration  (endogenous),  malonate,  174 

succinate  oxidation,  malonate,  54 
Creatinase,  mercurials,  836 
Creatine  kinase, 

adenosine,  446 

ADP,  447 

6-aminonicotinamide-NAD,  505 

iodine,  682,  685 


o-iodosobenzoate,  704,  707 

malonate,   60 

mercurials,  836 

nitrate,  446 

phosphate,  446 

sulfate,  446 

tripolyphosphate,  446 
Creatine  -  phosphate , 

muscle  homogenate  level,  o-iodosoben- 
zoate, 721 

tissue  levels,  2-deoxyglucose,  395 
Creatinine,  renal  transport  of, 

dehydroacetate,  625 
Crepis  capillaris  roots,  growth  of, 

mercurials,  966 
2?-Cresol,   tyrosine: a-ketoglutarate  trans- 
aminase, 306 
o-Cresotamide,    sulfanilamide    acetylase, 

601 
Crithidia  fasciculata, 

succinate  dehydrogenase,  malonate,  28 

succinate  oxidation,   malonate,   22-23, 

54 
Crocker  180  sarcoma, 

glycolysis,  ferricyanide,  677 

respiration  (endogenous),  malonate,  177 
Crotonase,  see  Enoyl-CoA  hydratase 
Crotonate  (/3-methylacrylate), 

D-amino  acid  oxidase,  343,  346 

fumarase,  279 

oxidation  of,  benzoate,  349 

structure  of,  279,  345 
Crotonyl-CoA,  acyl-CoA  dehydrogenase, 

591 
Crown  galls,  growth  of, 

malonate,  197 
Crown  gall  organism,  see  Agrobacterium 

tumefaciens 
Cryptococcus  terricolus, 

malonate  oxidation  in,  231 

respiration  (endogenous),  malonate,  231 
Cuckoopint,  see  Arum  maculatum 
Cupric  ions,  see  Copper 
Cyanate,  nitrite  oxidation,  450-451 
Cyanide, 

poisoning    by,    tetrathionate    antago- 
nism, 696 

renal  transport  of  PAH,  205 


1150 


SUBJECT  INDEX 


Cyanocobalamin, 
analogs  of, 

bacterial  growth,  589-590 
cyanocobalamin  biosynthesis,  589 
diol  dehydrase,  590 
methionine  biosynthesis,  590 
biosynthesis  of,  analogs,  589 

Cyclic  -  2',  3'-  adenosinemonophosphate, 
isocitrate  dehydrogenase,  509 

Cyclic  -  3 ',  5 '-  adenosinemonophosphate, 
phosphofructokinase,  474 

Cyclic-2',3'-guanosinemonopho8phate, 
deoxycytidylate  deaminase,  469 

Cyclobutane- 1 , 1  -dicarboxylate, 
intercharge  distance,  7 
succinate  dehydrogenase,  37,  40 

Cyclohexaneacetate,  chymotrypsin,  370 

y-CycIohexanebutyrate, 
chymotrypsin,  370 

kynurenine:a-ketoglutarate     transami- 
nase, 608-609 

Cyclohexanecarboxylate, 

kynurenine:a-ketoglutarate     transami- 
nase, 608-609 
tyrosinase,  300-302 

Cyclohexane- 1 ,2-dicarboxylate, 
intercharge  distance,  7 
kynurenine:a-ketoglutarate     transami- 
nase, 608 
succinate  dehydrogenase,  38,  40 

^-Cyclohexanepropionate,  chymotrypsin , 
369-370 

Cyclohexyl-DL-alanine,  phenylalanine  hy- 
droxylase, 354 

Cyclohydrolase,  folate  analogs,  585 

Cyclopentane-l,2-dicarboxylate, 
acetate  utilization,  241 
intercharge  distance,  7 
pyruvate  utilization,  241 
succinate  dehydrogenase,   37,   40,   241 

D-Cycloserine  (orientomycin,  Oxamycin), 
D-alanyl-D-alanine  synthetase,  360 
L-asparagine:a-ketoglutarate  transami- 
nase, 360 
bacterial  growth,  359-360 

antagonism  by  D-alanine,  359 
antagonism  by  alanylalanine,  360 
cell  wall  formation,  359 


central  nervous  system,  359 

GABA :  a  -  ketoglutarate   transaminase, 

359 

glutamate  decarboxylase,  359 

mycobacterial  growth,  359 

oxime  formation  with  pyridoxal-P,  359 

staphylococcal  resistance  to,   359 

structure  of,  359 

transaminases,  359 
L-Cycloserine, 

L-alanine:a-ketoglutarate  transaminase, 

360 

L-asparagine:a-ketoglutarate  transami- 
nase, 360 

bacterial  growth,  antagonism  by  L-ala- 

nine,  359 
Cystamine,  glucose  utilization  in  erythro- 
cytes, 663 
Cystamine  monosulfoxide,  3-phosphogly- 

ceraldehyde  dehydrogenase,  663 
D-Cysteine,  L-alanine  dehydrogenase,  354 
L-Cysteine, 

arginase,  337 

homoserine  kinase,  357 

serine  deaminase,  357 

tyrosine  decarboxylase,  307 
Cysteine  desulfurase, 

analogs,  357 

deoxypyridoxol  in  vivo,  570 

malonate,  60 
Cystine, 

enzyme  inhibitions,  661-664 

oxidation  of  protein  SH  groups,  661, 

663 
Cytidine, 

aspartate  carbamyltransferase,  467-468 

5 '-nucleosidase,  472 
Cytidinediphosphate   (CDP), 

aspartate  carbamyltransferase,  468 

NADH  oxidase,  511 
Cytidinemonophosphate  (CMP), 

adenylosuccinate  synthetase,  467 

aspartate  carbamyltransferase,  467-468 

deoxycytidylate  deaminase,  469 

pyrophosphatase,  475 

ribonuclease,  475 
Cytidinetriphosphate     (CTP),     aspartate 

carbamyltransferase,  468 


SUBJECT  INDEX 


1151 


Cytochrome  bj,  o-iodosobenzoate,  708 
Cytochrome  bj  reductase, 

iodine,  685 

mercurials,     coenzyme     displacement, 

787 
Cytochrome  c,  xanthyl-cytochrome  c,  592 
Cytochrome  c  oxidase, 

o-iodosobenzoate,    708 

malonate,  60 

mercurials,  837,  870-872 

quinacrine,  550 

thiols,  661,  663 
Cytochrome  c  reductase, 

mercurials,  relation  to  SH  groups,  804, 

809 

quinacrine,  547,  549 
Cytochrome  c-554  reductase,  quinacrine, 

549 
Cytosine,  D-amino  acid  oxidase,  545 

D 

Dahlia  leaves,  CO2  photochemical  fixation 

mercurials,  892 
dAMP,     see     Deoxyadenosinemonophos- 

phate 
Daptazole,  see  2,4-Diamino-5-phenylthia- 

zole 
Daraprim,  see  Pyrimethamine 
dCDP,  see  Deoxycytidinediphosphate 
dCMP,  see  Deoxycytidinemonophosphate 
dCTP,  see  Deoxycytidinetriphosphate 
DDD,see2,2'-Dihydroxy-6,6'-dinaphthyl- 

disulfide 
DDT  dehydrochlorinase,  ferricyanide,  675 
Deamino-AMP,     NADPHrcytochrome    c 

oxidoreductase,  511 
Deamino-ATP,  NAD  pyrophosphorylase, 

510 
2-Deaminofolate,  serine  biosynthesis,  585 
Deamino-NAD, 

NADH  oxidase,  511 

NADPH:cytochrome  c  oxidoreductase, 

511 
Deamino-NADP,    NADPH:glutathione 

oxidoreductase,  512 
Decane  - 1,10  -  dicarboxylate,   kynurenine: 

a-ketoglutarate  transaminase,  608 


Decanoate,  see  Caprate 
Dehydroacetate, 

antidotes  to,  628-629 

ATP  level  in  nuclei,  624 

bacterial  growth,  617-618,  631-633 

binding  to  plasma  proteins,  630 

blood  levels  of,  627-628 

cariogenic  action,  630-631 

central  nervous  system,  627 

chemical  properties,   618-620 

determination  in  tissues,  620 

diuresis,    625 

enzyme  inhibitions,  620-623 

fungal  growth,  632-633 

glucose  conversion  to  CO2,  624 

glycolysis  (anaerobic),  624 

growth  of  microorganisms,  pH  effects, 

633 

heart,  625 

hydrogenation  of,  619 

intestine,  624-625 

ionization  of,  619 

keto-enol   tautomerism,    618 

lethal  doses,  627-628 

metabolism  of,  629-630 

oxidative   phosphorylation,    623 

penicillin  levels  in  blood,  626 

permeability  to,   625 

purification   of,   620 

reaction  with  thiols,   621 

renal  transport  of  PAH,  205 

renal  transports,  625-626 

respiration    (endogenous),    623-624 

salivary  secretion  of,  630 

solubility    of,    619 

structure  of,  618-619 

succinate  oxidase,  620-622 

synthesis  of,  619-620 

tissue  distribution  of,  629-631 

toxicity  of,  627-629 

urinary  excretion  of,  629-631 

urinary  succinate  excretion,  628 

yeast  growth,  632 
Dehydroshikimate  reductase, 

active  center  of,  607 

analogs,   593,   604-606 
Dendraster  eggs, 

cleavage,  malonate,  198 


1152 


SUBJECT  INDEX 


development,  ferricyanide,  678 
Deoxyadenosinemonophosphate  (dAMP), 

adenylosuccinate  synthetase,  467 

aspartate  carbamyltransferase,  469 

deoxycytidylate  deaminase,   469 

phosphodiesterase,  473 
DeoxyAMP,    see    Deoxyadenosinemono- 
phosphate 
3-Deoxy-D-ara6o  -  heptonate -7 -phosphate, 

2-keto-3-deoxy-D-  arabo  -  heptonate-7-  P 

synthetase,  413 
DeoxyCDP,     see     Deoxycytidinediphos- 

phate 
DeoxyCMP,  see  Deoxycytidinemonophos- 

phate 
DeoxyCTP,     see     Deoxycytidinetriphos- 

phate 
Deoxycytidine,  aspartate  carbamyltrans- 
ferase, 468 
Deoxycytidinedisphosphate  (dCDP), 

hydrolysis  of,  deoxyCMP  and  deoxy- 

CTP,  446 

polynucleotide  phosphorylase,  474 
Deoxycytidinemonophosphate  (dCMP), 

aspartate  carbamyltransferase,  468 

deoxyCDP  hydrolysis,  446 

polynucleotide  phosphorylase,   474 
Deoxycytidinetriphosphate   (dCTP), 

aspartate  carbamyltransferase,  468 

deoxyCDP  hydrolysis,  446 

polynucleotide   phosphorylase,   474 
Deoxycytidylate  deaminase, 

analogs,  469 

mercurials,  837 

protection  by  deoxyCTP,  781 
Deoxycytidylate  kinase,  ADP,  469 
6-Deoxy-6-fluoroglucose, 

acetate  metabolism,  404 

fructose  fermentation  in  yeast,  404 

glucose  fermentation  in  yeast,  404 

glucose  oxidation  in  kidney,  393 

glucose  utilization,  403-405 

hexokinase,    404 

intestinal  transport  of,  404 

lactate  metabolism,  404 

lethal  doses,  404-405 

membrane  transport  of  hexoses,  404 

metabolism  of,  404 


oxidation  by  glucose  oxidase,  404 
2-Deoxygalactose,  respiration  (galactose), 

391-392 
4-Deoxygalactose,  respiration  (galactose), 

391-392 
2-Deoxygluconate,  metabolism  of,  389 
1-Deoxyglucose, 

intestinal  transport  of,  387 

a-methylglucoside  uptake,  394 
2-Deoxyglucose, 

absorption  of,  386-387 

acetate  metabolism  in  rabbit,  399 

acetate  oxidation,  397 

ADP  levels  in  tissues,  394-398 

amino  acid  sjTithesis  from  glucose,  399 

anaphylactoid  reaction,  401 

ATP  levels  in  tissues,  394-398 

C-l/C-6  ratio  in  brain,   383-394 

carcinostasis,  386,  400-401 

cell  division,  400 

central  nervous  system,  401 

citrate  level  in  ascites  cells,  399 

CO2  formation  from  glucose,  393 

Crabtree   effect,    396-397 

creatine-P  levels,  395 

distribution   in   tissues,   386-387 

DNA  level  in  carcinoma,  399 

epinephrine  release,  401 

Escherichia  coli  growth,  400 

ethanol  oxidation,  395-396 

fatty  acid  level  in  plasma,  399 

fatty  acid  oxidation,  397 

fibroblast   culture   growth,   400 

fructose  metabolism,  391,   398 

fructose   uptake,    394 

galactose  oxidation,  398 

/5-galactosidase  synthesis,  400 

glucokinase,  389-390 

glucose  membrane  transport,  390 

glucose  metabolism,  391-398 

glucose-6-phosphatase,  390 

glucose  uptake,  393-394 

glucose  uptake  in  vivo,  387-388,  401 

glutamine  level  in  brain,  399 

glycolysis,  391-394 

heart,  402-403 

hexokinase,    389-390 

hyperglycemia,  401 


SUBJECT  INDEX 


1153 


IMP  level  in  ascites  cells,  395 

intestinal  transport  of,  387 

K+  fluxes  in  atria,  403 

K+  uptake  by  brain,  399 

lethal   doses,   401 

lipogenesis  in  liver,  399 

mannose  uptake,   394 

metabolism  of,  386-389 

a-methylglucoside  uptake,  394 

palmitate  oxidation,   397 

pentose-P   pathway   stimulation,   393- 

394 

phosphorj-lation  of,  387-389 

epinephrine  effect,   387 

insulin  effect,  387 

K^'s  for  hexokinases,  388 
protein  biosynthesis,  399 
pyruvate  decarboxylation,  396 
pyruvate  oxidation,  392,  397 
pjTuvate  utilization  in  ascites  cells,  399 
resistance  of  HeLa  cells  to,  388 
respiration  (endogenous),  391-392,  396- 
397 

comparison  with  glucose,  396-397 
respiration  (glucose),  391-394,  397 

combined  with  iodoacetate,  397 
summary  of  mechanisms  of  action  on 
carbohydrate  utilization,  398 
toxicity,  401 

transport  into  cerebrospinal  fluid,  401 
urinary  excretion  of,  388 
utihzation  by  fungi,  387,  400 
virus  proliferation,  400 
3-Deoxyglucose, 

intestinal  transport  of,  387 
a-methylglucoside  uptake,  394 
6-Deoxygluco8e, 

glucose  oxidation,  403 
intestinal  transport  of,  403 
intestinal  transport  of  hexoses,  264,  403 
a-methylglucoside  uptake,  394 
2-Deoxyglucose-6-phosphate, 
formation  of,  387-389 
glucose  transport,  390 
glucose-6-P  dehydrogenase,  390-391 
glycogen  synthetase,  391 
oxidation  of,  388 
phosphoglucose  isomerase,  390 


UDPG:a-l,4-glucan-a-4-glucosyltrans- 

ferase,  391 
3-Deoxyglucose-6-phosphate, 

hexokinase,  380 

structure  of  378 
Deoxyguanosinemonophosphate  (dGMP), 

adenylosuccinate  synthetase,   467 

deoxycytidylate  deaminase,  469 
DeoxypjTidoxine,  see  DeoxypjTidoxol 
Deoxypyridoxol    (deoxypyridoxine,    des- 

oxypyridoxine ) , 

acrodynia,  566 

blood  cholesterol,   574 

blood   phospholipids,   574 

carcinostasis,  576 

cysteine  desulfurase,  570 

dermatitis,  577 

fatty  acid  biosynthesis,  574 

glutamate  decarboxylase,  569-570 

growth  of  microorganisms,  575-576 

intestinal  amino  acid  transport,  574 

intestinal   sugar   transport,    574 

leucocyte  count,  577 

malignant  carcinoid  syndrome,  574 

phosphorylation  of,  564 

pyridoxal  kinase,  565 

pyridoxal-P  level  in  tissues,  566 

p>Tidoxamine-P  oxidase,  566 

p\Tidoxine  levels  in  tissues,  566-568 

pyridoxol-P  oxidase,  566 

serine  biosynthesis,  570-571 

serotonin  metabolism,  574 

structure  of,  563 

toxicity,   562,566-567,   577-578 

Toxoplasma  infections,  576 

transaminases,  569-570 

tryptophan  metabolism,  572 

urinary  excretion  of  xanthurenate,  572 

vitamin  Bg  deficiency,  562,  577-578 
Deoxypyridoxol-phosphate, 

pyridoxamine-P  oxidase,  566 

pyridoxol  oxidation,  564 

pyridoxol-P  oxidase,   566 
5'-Deox>Tiboflavin,  riboflavin  biosynthe- 
sis, 539 
Deoxyribonuclease, 

ferricyanide,  675 

mercurials,  860 


1154 


SUBJECT  INDEX 


RNA,  462 

Deoxyribonucleates,  see  also  Nucleic  acids 
carcinoma  levels,  2-deoxyglucose,  399 
ribonuclease,  462 

Deoxyribose-phosphate  aldolase,  mercu- 
rials, 837 

2'-Deoxyribosyl-4-aminopyrimidone-2,5'- 
phosphate  deaminase,  ferri cyanide,  675 

Deoxythymidine,  aspartate  carbamyl- 
transferase,  469 

Deoxythymidine  kinase,  deoxyTTP,  470 

Deoxythymidinemonophosphate  (dTMP), 
deoxycytidylate  deaminase,  469 

Deoxythymidinetriphosphate  (dTTP),  de- 
oxythymidine kinase,  470 

Deoxyuridinemonophosphate  (dUMP),  de- 
oxycytidylate deaminase,   469 

Desert  locust,  see  also  Schristocera 

butyrate  oxidation  in  muscle,  malonate, 
137 

Desmethyldesthiobiotin , 
Escherichia  coli  growth,  588 
structure  of,  588 

Desoxypyridoxine,  see  DeoxypjTidoxol 

Desthiobiotin, 

biotin  oxidase,  589 
structure  of,  588 

Desulfovibrio  desulfuricans, 
growth  of,  mercurials,  972 
tetrathionate  reduction  in,  699 

a-l,6-Dextranglucosidase,  analogs,  417 

Dextrin- 1,6-glucosidase,  see  Amylo-1,6- 
glucosidase 

DFPase,  o-iodosobenzoate,  711 

dGMP,  see  Deoxyguanosinemonophos- 
phate 

Diacetylthiamine,  thiamine  kinase,  523 

Dialkylfluorophosphatase,  see  also  DFP- 
ase 
malonate,  role  of  Mn++  in  inhibition,  68 

Diamidines, 

diamine  oxidase,  363-365 
structures  of,   361 

Diamine  oxidase  (histaminase), 
analogs,   360-365 
hydrazine,  362 
malonate,  60 
semicarbazide,  362 


2,4-Diamino-6,7-dihydroxypteridine,  xan- 
thine oxidase,  289 
1 ,2-Diamino-4,5-dimethylbenzene,  cyano- 

cobalamin  biosynthesis,  590 
2,4-Diamino-7,8-dimethyl-10-ribityl-5,10- 

dihydrophenazine,  structure  of,  537 
2,4-Diamino-6-formylpteridine,    dihydro- 

folate  reductase,   583-584 
a,£-Diamino-^-hydroxypimelate,     diami- 

nopimelate  decarboxylase,  593 
2,4-Diamino-6-hydroxypteridine,dihydro- 

folate  reductase,  583-584 
2,6-Diamino-8-hydroxypurine,     xanthine 

oxidase,  282 
2,4-Diamino-6-methylpteridine,   dihydro- 

folate  reductase,   583-584 
2,4-Diamino-5-phenylthiazole  (amiphena- 

zole,  Daptazole),   thiamine  deficiency, 

531 
Diaminopimelate  decarbocylase, 

analogs,  593 

o-iodosobenzoate,  708 
2,3-Diaminopropionate,  aspartate:a-keto- 

glutarate  transaminase,  355 
2,6-Diaminopurine, 

dihydrofolate  reductase,  583-584 

purine  metabolism,  480 

xanthine  oxidase,  282 
Diaphragm, 

acetate  uptake,  mercurials,  912 

acetate  utilization,  propionate,  613 

amino  acid  uptake,  malonate,  155 

carbohydrate  uptake,  competition  be- 
tween sugars,  264 

C-l/C-6  ratio,  malonate,  130 

citrate  accumulation,  malonate,  104 

contractility,  mercurials,  896 

2-deoxyglucose  uptake, 
glucose,  387 
insulin,  387 

fructose  oxidation,  2-deoxyglucose,  398 

galactose     oxidation,     2-deoxyglucose, 

398 

glucose  uptake,   mercurials,  876,  893- 

894 

glucose  utilization, 

6-deoxy-6-fluoroglucose,  404 
2-deoxyglucose,  398 


SUBJECT  INDEX 


1155 


6-deoxyglucose,  403 
mercurials,  884 

glycogen  level,  mercurials,  884 

glycolysis,   2-deoxyglucose.  392 

malonate  decarboxylation  in,  232 

mercurial  levels  in  vivo,  930 

mercurial  penetration  into,   879 

mercuric  ion, 

complexing  material  from,  907 
uptake,  894-897 

respiration     (endogenous),      malonate, 

180-181,  187 

respiration  (glucose),   mercurials,   883, 

893-894,  898 

xylose  uptake,  mercurials,  911-912 
6-Diazo-5-oxo-L-norleucine  (DON ), 

formylglycinamide  ribonucleotide  ami- 

dotransferase,  333 

glutaminase,  356 

glutamine:fructose-6-P  transamidase, 

356 

inosinate  biosynthesis,  333 

phosphoribosyl-PP  amidotransferase, 

333 

structure  of,  333 
Dibenzamidines, 

diamine  oxidase,  364 

structures  of,  361 
Dibenzoylthiamine,  thiamine  kinase,  523 
3,5-Dibromobenzoate,    dopa    decarboxy- 
lase, 312-313 
3,5-Dibromot>Tosine,   tjTosinera-ketoglu- 

tarate  trasaminase,  306 
2,2'  -  Dicarboxy  -  4,4'  -  diiodoaminoazaben- 

zene,  determination  of  SH  groups,  641 
Dicarboxylate  ions, 

chelation  with  cations,  11-13 

intercharge  distances,  5-7,  44 

ionization  of,  7-11 

kynurenine:a-ketoglutarate     transami- 
nase, 608 

permeability  of  erythrocytes  to,  187-189 

pH  on  dianion  concentration,  191 
Dichloroacetate,    pantoate :  /J  -  alanine    li- 

gase,  597 
Dichloroarabitylflavin,  flavokinase,  539 
2,4-Dichlorocinnamate,  dopa  decarboxy- 
lase, 313 


2.3  -  DichloroisobutjTate,      pantothenate 
biosynthesis,  588 

2.4  -  Dichloro  -  6  -  phenylphenoxyethyla- 
nine,  butynamine  demethylase,  592 

a,a-Dichloropropionate,   pantoate :  ^  -  ala- 
nine ligase,  598 

Dichlororiboflavin  analog, 

L-amino  acid  oxidase,  540-541 
bacterial  growth,   537 
phosphorylation  of,  539 

6,7-Dichloro  -  9  -  (1 '-  d  -  sorbityl)  isoalloxa- 
zine,  carcinostasis,  538 

Dichromate,  enzyme  inhibition,  660 

Diethylethoxymethylenemalonate,   carci- 
nostasis, 202 

Diethyl-L-glutamate,   L-glutamate   dehy- 
drogenase, 331 

3,3-Diethylglutarate,   kynurenine:a-keto- 
glutarate  transaminase,  608 

6,7-Diethylribotlavin  analog, 
phosphorylation  of,  539 
structure  of,  536 
utilization  by  Lactobacillus,  539 

Diethylstilbestrol,    /J-hydroxysteroid    de- 
hydrogenase, 449 

Diethylthetin,  thetin:homocysteine  trans- 
methylase,  356 

Difluoromalonamide,  succinate  dehydro- 
genase, 239 

Difluoromalonate,    succinate   dehydroge- 
nase, 239 

Diguanidines, 

diamine  oxidase,  363-365 
structure  of,  361 

Dihydrofolate  reductase, 

amethopterin  in  vivo,  582-583 
analogs,  581-584 
mercurials,  816 

Dihydroorotase,  analogs,  470 

Dihydroorotatedehydrogenase,  analogs, 70 

Dihydroxyacetone,    phosphopentose   iso- 
merase,  411 

Dihydroxyacetone-phosphate, 
enolase,  409 
phosphopentose  isomerase,  411 

Dihydroxybenzoates, 

dehydroshikimate  reductase,  593,  605- 
606 


1156 


SUBJECT  INDEX 


dopa  decarboxylase,  312 

erythro  -  2,3  -  Dihydroxybutyrate  -  phos- 
phates, enolase,  410 

3,4-Dihydroxycinnamate,  see  Caffeate 

5  -  (3,4  -  Dihydroxycinnamoyi)  -  salicylate, 

dopa  decarboxylase,  312 

2,2'  -  Dihydroxy-6,6'  -  dinaphthyldisulfide 
(DDD),  determination  of  SH  groups, 
921 

3,4-Dihydroxyhydrocinnamate,  dopa  de- 
carboxylase, 313 

Dihydroxymaleate,  catechol  oxidase,  297 

2,4-Dihydroxy-6-methylpyrimidine,  dihy- 
droorotate  dehydrogenase,  470 

6,7-Dihydroxy-7-n-pentyl-8-(r-ribityl)lu- 
mazine,  riboflavin  synthetase,  543 

Dihydroxyphenylalanine,  see  Dopa 

Dihydroxyphenylalanine  decarboxylase, 
see  Dopa  decarboxylase 

2,5-Dihydroxyphenylpyruvate,  p-hydro- 
xyphenylpyruvate  oxidase,  306 

6,7-Dihydroxyriboflavin  analog,  ribofla- 
vin synthesis,   539 

Diimidotriphosphate,  oxidative  phospho- 
rylation, 448 

Diiodothyronines,  thyroxine  deiodinase, 
603 

3,5-Diidotyrosine, 

dopa  decarboxylase,  308 

tyrosine:  a-ketoglutarate  transaminase, 

306 

Diisopropylfluorophosphonatase,  see  DFP 
ase 

Diisothioureas, 

diamine  oxidase,  363-365 
structures  of,  361 

a,y-Diketovalerate,  lactate  dehydroge- 
nase, 437 

Dimercaptides,ofHg++ and  thiols,  746-751 

6  -  ( 2 , 6  -  Dimethoxybenzamido)    penicilli- 

nate, penicillinase,  599,  615,  688 
Dimethoxyphenylalanines,    dopa    decar- 
boxylase, 308 
Dimethoxyphenylethylamines,   dopa  de- 
carboxylase, 308 
Dimethylacetone,  glyoxylase,  594 
Dimethylacrylate,  d -amino  acid  oxidase, 
343 


iV^-Dimethyladenine, 

adenine  deaminase,  466 

inosine  phosphorylase,  471 
iV-(4  -  Dimethylamino  -  3,5  -  dinitrophenyl) 

maleimide,  determination  of  SH  groups, 

641 
5,6-Dimethylbenzimidazole, 

cyanocobalamin  synthesis  from,  589 

Lactobacillus  growth,  589-590 
Dimethylglutarates,   kynurenine :  a  -  keto- 

glutarate  transaminase,  608 
Dimethylglycine,  betaine  :  homocysteine 

transmethylase,  356 
Dimethylguanidine,  histidase,  353 
4,4-  Dimethyl  - 17  ^  -  hydroxyandrost  -  5  - 

eno(3,2-c)    pyrazole,    /?- hydroxy  steroid 

dehydrogenase,  449 
iV-Dimethylmalondiamide,    carcinostasis, 

202 
2,2-Dimethylsuccinate,    kynurenine:a-ke- 

toglutarate  transaminase,  608 
2,4  -  Dimethyl  -  cyclo  -  telluropentane  -  3,5- 

dione, 

bacterial   growth,   576 

structure  of,  563 
2,4-Dinitrophenol  (DNP), 

gastric  acid  secretion,  915-916 

mercurial  inhibition  of  ATPase,  869 

mitochondrial  swelling,  210 

porphyrin  synthesis,   162 

renal  transport  of  PAH,  205 
2,6-Dinitrophenol,  D-amino  acid  oxidase, 

348 
Dinitrophenol  reductase,  mercurials,  860 
^-2,4  -  Dinitrophenylpropionate,   chymo- 

trypsin,  369-370 
Diol  dehydrase, 

cyanocobalamin,  590 

hydroxocobalamin,  590 
Dioxindole,  acetylindoxyl  oxidase,  591 
Dipeptidase, 

analogs,  367-368 

o-iodosobenzoate,  708 

mercurials,  837 
o-Diphenol  oxidase,  see  Catechol  oxidase 
Diphenylglycolate,  glycolate  oxidase,  593 
Diphenylphosphate,   phosphatase  (acid), 

441 


SUBJECT  INDEX 


1157 


Diphosphite,  oxidative  phosphorylation, 

448 
2,3-Diphosphoglycerat€,  phosphoribomu- 

tase,  413 
Diplococcus  pneumoniae,  infection  by, 

malonate,  221-222 
(a,/3  -  Distearoyloxypropyl)dimethyl  -  {(i'- 

hydroxyethyl)ammonium  acetate,  leci- 

thinase  A,  595 
Disulfide   groups,    role    in    reactivity    of 

enzymes  with  SH  reagents,  644-645 
Disulfides, 

enzyme  inhibition,  661-663 

mixed,  see  Mixed  disulfides 

oxidation  of  SH  groups,  661,  663-664 

reduction  by  dithiothreitol,  640 
5,5'-Dithiobis  (2-nitrobenzoate),  determi- 
nation of  SH  groups,  641 
Dithioglycolate, 

ATPase,  663 

oxidation  of  protein  SH  groups,  661 
Dithiothreitol,  protein  disulfide  reduction, 

640 
6,8-Dithiourate,  uricase,  285-286 
Diuretics,  see  Mercurial  diuretics 
Division,  see  Cell  division 
DNA,  see  DeoxjTibonucleates 
DNAase,   see   Deoxyribonuclease 
DON,   see  6-diazo-o-oxo-L-norleucine 
Dopa, 

acetylindoxyl  oxidase,  591 

histidine  decarboxylase,  352 

kynureninase,  595 

metabolism  of,  pathways,  307 

phenylalanine   deaminase,   355 

tyrosine: a-ketoglutarate  transaminase, 

306 
Dopa  decarboxylase, 

analogs,  307-320 

deoxypyridoxol  in  vivo,  596-570 

folate  analogs,  586 
Dopamine, 

brain  levels,  a-methyl-w -tyrosine,  316 

dopa  decarboxylase,  308 

metabolism  of,  307 

tissue  levels  of,  a-methyldopa,  315-320 
Dopamine  /3-hydroxylase, 

analogs,  320 


a-methyldopa,  316 
Double  bond,  interaction  energy  due  to 

polarization,  276 
DPN,  see  NAD 

DT  diaphorase,  o-iodosobenzoate,  708 
dTMP,     see     Deoxythymidinemonophos- 

phate 
dTTP,  see  Deoxythymidinetriphosphate 
dUMP,  see  DeoxjTiridinemonophosphate 


Eagle's  KB  carcinoma,  growth  of, 
deoxypjTidoxol,  577 
mercurials,   968 

Earle  sarcoma  cells,  respiration  (endoge- 
nous), 
malonate,  177 

Eberthella  typhosa,  o-iodosobenzoate  kil- 
ling of,   727 

Echinococcus  granulosus,  respiration  (en- 
dogenous), 
malonate,  173 
mercurials,  882 

Echinus  esculentus  eggs,  respiration  (en- 
dogenous), 
malonate,  175 
mercurials,  882 

Echinus  miliaris,  development  of, 
mercurials,  964 

Echo  7  virus,  infectivity  of, 
mercurials,  976-977 

Ectromelia  virus,  inactivation  by  mercu- 
rials, 977 

Ehrlich  asctes   carcinoma   cells,   see  As- 
cites carcinoma  cells 

Elastase,  mercurials,  860 

Electrical  potentials,  see  Membrane  po- 
tentials 

Electrocardiogram,  see  Heart,  electrocar- 
diogram 

Embryogenesis,  see  also  Gastrulation  and 
specific  organisms 
deoxypyridoxol,  576 
ferricyanide,  678 
o-iodosobenzoate,  726-727 
malonate,    197-199 
mercurials,  963-965 


1158 


SUBJECT  INDEX 


porphyrindin,  670 
Embryos, 

glycolysis,  hydrogen  paroxide,  695 
respiration,  mercurials,  882 
Enantiomers,  as  analog  inhibitors,  268- 

271 
Encephalitis   virus,   see   Western   equine 

encephalitis  virus 
Encephalomyocarditis  virus,  inactivation 

by  mercurials,  977 
Endamoeha  histolytica,  growth  of, 

malonate,  196 
Endogenous  respiration,  see  Respiration 

(endogenous) 
Endomyces  vernalis,   pyrithiamine-resist- 

ant  strain,  529 
Enolase  (phosphopyruvate  hydratase), 
analogs,  409-410 

mercurials,  768,  789,  803,  810,  837 
aggregation,  789 
rate  of  inhibition,  810 
relation  to  SH  groups,  803 
mercuric  ion,  crystalline  complex  with, 
768 
Enoyl-CoA  hydratase  (crotonase), 
o-iodosobenzoate,  707-708 
mercurials,  837 
Enteroviruses,  adsorption  to  kidney  cells, 

mercurials,  981 
Enzymes, 

biosynthesis  of, 
D-asparagine,  269 
azatryptophan,  326 
fluorophenylalanine,  351 
5-methyltryptophan,  326 
tryptazan,  326 
induction  of, 

8-azaguanine,  478 
azatryptophan,  326 
malonate,  155-156 
mercurials,  888 
a-methyltryptophan,  325 
Epidermis,  mitosis  in, 
malonate,  199-200 
mercurials,  968 
Epidermophyton     floccosum,     respiration 
(endogenous), 
malonate,  169 


Epididymal  fat  pad,  glucose  oxidation, 
D-glucosamine,  382 

Epinephrine,  see  also  Catecholamines 
cardiac  stimulation, 
malonate,  217 
mercurials,  947 
2-deoxyglucose  phosphorylation,  387 
dopa   decarboxylase,   308 
formation  of,  pathways,  307 
release  of,  2-deoxyglucose,  401 
responses  to,  pjTogallol,  611 
tissue  levels  of,  a-methyldops,  315-320 
tyrosi  ne :  a  -  ketoglutarate  transaminase , 
305-306 
urinary  excretion,  pyrogallol,  612 

Epinine,     phenylalanine     /5-hydroxylase, 
600 

Equilin  dehydrogenase,  o-iodosobenzoate, 
708 

Ergot,  see  Claviceps 

Erythredema,  see  Acrodynia 

Erythrocytes, 

amino  acid  transport,  competition  by 
sugars,  267 

carbohydrate  uptake,  competition  be- 
tween sugars,   264 

glucose    uptake,    mercurials,    903-905, 
911 

glucose  utilization,  cystamine,  663 
GSH  level,  mercurials,  905-906 
heme  biosynthesis,  mercurials,  888 
iodine  hemolysis,  690 
K+  efflux,  mercurials,  903-905 
K+  influx,  mercurials,  908 
K+  trasport,  malonate,  209 
membrane  of,  mercurials,  906 
mercuric  ion  uptake,  897,  900-907 
Na+  transport,  malonate,  209 
nonelectrolyte  transport,  iodine,  690 
permeability, 

mercurials,   900-907 
to  dicarboxylates,  187-189 
porphyrin  biosynthesis, 
malonate,  159-163 
mercurials,  888 
protoporphyrin   biosynthesis, 
arsenite,  162 
2,4-dinitrophenol,  162 


SUBJECT  INDEX 


1159 


fluoroacetate,  162 

malonate,  162 
SH  groups  in,  897 
urate  transport,  hypoxanthine,  267 
Erytlu-ocytes  (Plasmodium -parasitized), 
ATP  level,  quinacrine,  560 
glucose  utilization,  quinacrine,  560 
respiration   (glucose), 

malonate,  124 

quinacrine,  560 
succinate  accumulation,  malonate,  91, 
93 
Erythrose-4-phosphate,     phosphoglucose 

isomerase,  407 
Escherichia  coli, 

adaptive  enzyme  synthesis,  malonate, 
155 

azatryptophan   incorporation   into  en- 
zymes of,  326 

cycle  intermediates  concentrations,  89 
2 -deoxy glucose  uptake,  387 
glutamate  dehydrogenation,  malonate, 
152 

glycolysis,  2-deoxyglucose,  387 
grov/th  of, 

azatryptophan,  326 

biotin  analogs,  588 

dehydroacetate,  632 

2-deoxyglucose,  400 

malonate,  195 

mercurials,  972 

methylindoles,  321,  323 

methylthio  analog  of  thiamine,  530 

4-methyItryptophan,  323 

D-phenylalanine,  268 
o-iodosobenzoate  killing  of,   727 
a-ketoglutarate  oxidation,  malonate,  79 
lactate  oxidation,   malonate,   78 
malate  oxidation,   malonate,   81 
malonate  metabolism  in,  228 
mercurial  uptake,  974-975 
oxidative   phosphorylation,    malonate, 
120 

protein  biosynthesis,  5-fluorouracil,  479 
pyruvate  oxidation, 

malonate,  74 

mercurials,  878 
resistance  to  mercurials,  983-984 


succinate  dehydrogenase, 
adipate,  35 
glutarate,  35 
malonate,  2,  21,  26,  187 
oxalate,  35 
tartronate,  36 

succinate  oxidation,  malonate,  52 

tryptazan  incorporation  into  enzymes 

of,  326 
Escherichia  coli  phage,  see  Coliphage 
Esterase    (microsomal),    stimulation    by 

mercurials,  816 
Esterases,  see  individual  enzymes 
Estradiol, 

cholesterol  esterase,  592 

/9-hydroxysteroid  dehydrogenase,  447, 

449 
Estradiol- 17/?-dehydrogenase,  mercurials, 

protection  by  estradiol,  781 

protection  by  NAD,  781 
Estra-l,3,5-trienes,  ^-hydroxysteroid  de- 
hydrogenase, 449 
Ethane,  nitrogen  fixation,  291 
1,2-Ethenediphosphonate, 

ionization   constants,   242 

succinate  dehydrogenase,   243 
1 ,2-Ethanedisulfonate, 

aspartase,  355 

inter  charge  distance,  7 

succinate  dehydrogenase,  242-243 
Ethanesulfonate,  sulfite  oxidase,  451 
Ethanol,  oxidation  of, 

2-deoxyglucose,   395-396 

mercurials,  898 
Ethanolamine,  choline  oxidase,  290 
Ethanolamine   oxidase,   quinacrine,   547, 

550 
Ethyl- 1 -acetyl  -  2  -  benzylcarbazate,  chy- 

motrypsin,  373 
2-Ethyl-3-amino-4-ethoxymethyl-5-ami- 

nomethylpyridine,     pyridoxal     kinase, 

564 
Ethylenediamine,  diamine  oxidase,  362 
Ethyl-D-glutamate, 

glutaminase,  333 

urinary  flow,  333 
iV-Ethyl-DL-leucine,    D-amino    acid    oxi- 
dase, 340 


r 


1160 


SUBJECT  INDEX 


Ethylmalonate, 

biosynthesis  of,  226 
carcinostasis,  201 
lethal  dose,  201 
occurrence  of,  225 

6-Ethyl-8-mercaptooctanoate,  acyl  trans- 
fer, 590 

Ethylmercuri-p-toluene  sulfonanilide,  see 
Granosan  M 

3  -  Ethyl-4-methyIthiazole,      thiaminase, 
524 

Ethyloxalacetate,  malic  enzyme,  597 

iV^-Ethyl-DL-phenylalanine,  D-amino  acid 
oxidase,  340 

<S-(iV-Ethylsuccinimido)-GSH,glyoxylase, 
593 

2'-Ethylthiamine,  thiamine  kinase,  523 

a  -  Ethylthioglucopyranoside,     a  -  gkicosi- 
dase,  423 

Euglena  gracilis, 

acetate  oxidation,  malonate,  77 
growth  of,  pyrithiamine,  529 
malonate  metaboHsm  in,  228 
succinate  dehydrogenase,  malonate,  28 
succinate  oxidation,  malonate,  51,  53, 
56 

Exopenicillinase,  iodine,  688 


FAD  (flavin  adenine  dinucleotide), 
analogs  of,  see  Riboflavin,  analogs  of 
L-lactate  oxidase,  543 
lysolecithin  oxidase,  543 
metabohsm  of,  pathways,  535 
NADH   :   ferricyanide  oxidoreductase, 
510 
succinate  oxidase,  543 

FAD  pyrophosphorylase, 
isoriboflavin,  542 
riboflavin,  542 

Fagus,  see  Beech 

False  feedback  inhibition,  321 

Fasciola    hepatica,    respiration    (endoge- 
nous), 
malonate,  173 

Fats,  see  Lipids 

Fatty  acids, 


biosynthesis  of, 

acyl-CoA  analogs,  613-614 

deoxypyridoxol,  574 

malonate,  146-149 

mercurials,  887 

a-phenylbutyrate,  614 

propionate,  613 
formation  from  malonate,  234 
formation  from  propionate,  malonate, 
146 

kynurenine:a-ketoghitarate     transami- 
nase, 608-609 

pH  effects,  609 
oxidation  of, 

benzoate,  349 

2-deoxyglucose,  397 

malonate,  135-137,  141-143 

mercurials,  887 
plasma  levels,  2-deoxyglucose,  399 
Fatty  acid  synthetase,  mercurials, 
protection   by   acetyl-CoA,    781 
Fatty  acid  thiokinase,  mercurials,  887 
Feedback  inhibition, 
analogs,  351 

anthranilate  synthesis,  321 
false,  321 

glutamate  dehydrogenase,  514 
pyrimidine  metabolism,  478-481 
Fermentation,    see    also    substances    fer- 
mented and  Yeast,  fermentation 
biotin  analogs,  588-589 
D-glucosone,  384-385 
iodine,  689 
mercurials,  875 
tripolyphosphate,  383 
Ferric  ions,  see  also  Iron 

glutamate  dehydrogenase,   863 
Ferricyanide,  670-678 
Aspergillus  growth,  677 
chemical  properties,  670-671 
citrate  accumulation,   677-678 
enzyme  inhibitions,  672-676 
glucose  oxidation,  677-678 
glycolysis,  673,  677 
itaconate    metabolism,    678 
mitochondrial  swelling,  678 
NADH  oxidation  by,  673 
oxidation  of, 


SUBJECT  INDEX 


1161 


amino  acids,  672 
NADH,  673 

protein  SH  groups,  670-672 
SH  groups,  670-671 

oxidation-reduction  potential,  670-671 

pentose-P  pathway,  677 

phospholipid  biosynthesis,  678 

porphyrin  biosynthesis,  678 

purification  of,  671 

respiration,  678 

sea  urchin  egg  development,  678 

yeast  growth,  678 
Ferrocyanide, 

isocitrate  dehydrogenase,  677-678 

tricarboxylate  cycle,  677-678 
Fertilization, 

o-iodosobenzoate,  726-727 

malonate,  198 

mercurials,  963-964 
Ferulate,  peroxidase,  599 
Fibroblasts,  growth  of, 

mercurials,  968-969 
Fibrosarcoma,  growth  of, 

deoxypyridoxol,  576 
Ficin,  mercurials, 

dimeric  complex  with  Hg++,  770 

relation  to  SH  groups,  804 
Flagellar  motility, 

o-iodosobenzoate,    727 

malonate,  203 
Flavin  adenine  dinucleotide,  see  FAD 
Flavin  mononucleotide,  see  FIVIN 
Flavins, 

analogs  of,  see  Riboflavin,  analogs  of 

tissue  levels,  galactoflavin,  539-540 
Flavobacterium, 

protocatechuate  transport,  p-aminosa- 

Ucylate,  267 

renal  transport  of  PAH,  p-aminosalicy- 

late,  613 
Flavokinase  (riboflavin  kinase), 

analogs,  539 

mercurials,  854 

permanganate,  660 
Flavotin, 

structure  of,  537 

succinate  oxidase,  543 
Flexner-Jobling  tumor, 


malonate  levels  in  vivo,  100-102 

succinate  accumulation,  malonate,  100- 

103 
Fluoride, 

/?-ketoadipate  chlorinase,  453 

phosphorylase,  406 

renal  transport  of  PAH,  205 

tjTosinase,  300 
Fluoride  dimer,  phosphatase  (acid),  441- 

442 
Fluorine,  use  in  forming  analogs,  258-259 
Fluoroacetate, 

lethal  doses,  239 

porphyrin  biosynthesis,  162 

renal  transport  of  PAH,  205 

sequential    inhibition    with    malonate, 

112 
Fluoroacetyl-L-tyrosinate,  chymotrypsin, 

371 
Fluoroamino     acids,     amino     acid     and 

protein  metabolism,  351 
Fluorobenzoates,   D -amino  acid  oxidase, 

341 
5-Fluorocytidine,     aspartate     carbamyl- 

transferase,  467 
5-Fluorodeoxycytidine,   aspartate  carba- 

myltransferase,  469 
5-Fluorodeoxyuridine,   thymidylate  syn- 
thetase, 476 
5-Fluorodeoxyuridinemonophosphate 

(FdUMP),  thymidylate  synthetase,  47(  , 

479 
/3-Fluoro-DL-malate, 

fumarase,  279 

malate  dehydrogenase,  279 
Fluoromalonate, 

citrate  accumulation,  239 

decarboxylation  of,  239 

lethal  doses,  239 

succinate  dehydrogenase,  239 
Fluoromalonic  diethyl  ester,  lethal  doses, 

239 
5-Fluoroorotate, 

aspartate  carbamyltransferase,  468 

conversion  to  5-fluoro-UMP,  478 

dhydroorotase,  470 

UTP   biosynthesis,   478-479 
Fluorooxalacetate, 


1162 


SUBJECT    INDEX 


aspartate :  a  -  ketoglutarate      transami- 
nase, 334 

malate  dehydrogenase,  596 

transamination  to  fluoroaspartate,  334 
p-  Fluoropheny  lalanine , 

incorporation  into  proteins,  351 

maltase  biosynthesis  in  yeast,  351 

L-phenylalanine:sRNA    Ugase    (AMP), 

354 

protein  biosynthesis,  351 

replacement  of  phenylalanine  in  pro- 
teins, 351 
;5-FluorophenylaIanine,  tyrosine  transport 

by  brain,  266 
Fluorophosphate,     see     Monofluorophos- 

phate 
Fluoropyrimidines,    pyridimidine    meta- 

bohsm,  478-481 
5  -  Fluor  otryptophan , 

tryptophan  pyrrolase,  325 

L-tryptophan:sRNA  ligase  (AMP),  326 
6-Fluorotryptophan, 

anthranilate  metabohsm,  321 

tryptophan  pyrrolase,  325 

L-tryptophan:sRNA  ligase  (AMP),  326 
3-Fluorotyrosine, 

tyrosinase,  304-305 

tyrosinera-ketoglutarate  transaminase, 

305-306 

L-tyrosine:sRNA  ligase  (AMP),  307 
5-Fluorouracil, 

abnormal  enzymes  produced  by,  479- 

480 

protein  biosynthesis,  479 

thymidylate  synthetase,  476 
5-Fluorouridine,  thymidylate  synthetase, 

476 
5-Fluorouridinemonophosphate  (FUMP), 

thymidylate  synthetase,  476 
FMN  (flavin  mononucleotide), 

D-amino  acid  oxidase,  540-541 

glutamate  racemase,  542 

L-lactate  oxidase,  543 

NADH:ferricyanideoxidoreductase,510 

riboflavin  transglucosidase,  543 

succinate  oxidase,  543 
Folate, 

analogs,  of,  579-586 


acetylations,  586 
ATP  level  in  tissues,  585 
enzyme  inhibitions,  586 
folate  deficiency,  581 
folate  reduction,  581-584 
nucleic  acid  biosynthesis,  584-585 
protein  biosynthesis,  585 
structures  of,  580 
metabolic  functions  of,  579 
metabolism  of,  pathways,  579 
photolytic    oxidation    products   as  in- 
hibitors of  xanthine  oxidase,  285-287 
structure   of,    580 

Folate  reductase,  see  Dihydrofolate  reduc- 
tase 

Folinate  (citrovorum  factor) 

formation  from  folate,  analogs,  582 
structure  of,  580 

Foot-and-mouth   virus,   proliferation   of, 
malonate,   194 

Formaldehyde,   pyruvate  decarboxylase, 
432,  600 

Formaldehyde  polymers,   hyaluronidase, 
459-461 

Formate, 

glutamate  decarboxylase,  328 
kynurenine:a-ketoglutarate     transami- 
nase, 608 

lactate  dehydrogenase,  436 
tyrosinase,  300 

Formate  dehydrogenase,  hypophosphite, 
593 

Formate  hydrogenlyase, 
hypophosphite,  593 
o-iodosobenzoate,    708 

Formate  transacetylase,  ferricyanide,  675 

Formylglycinamide  phosphoriboside  syn- 
thetase, analogs,  333 

9-Formylmethylriboflavin,    flavokinase, 
539 

6-Formylpteridine,  see  Pterin-6-aldehyde 

Form yltetrahydro folate  synthetase,  ana- 
logs, 585 

A'^-Formyl-L-tyrosine,  tyrosinase,  304-305 

N  -  Formyl  -  L-  tyrosinemethylamide,   chy- 
motrypsin,  371 

Fowl  plague  virus,  infectivity  of, 
mercurials,  977,  980 


SUBJECT    INDEX 


1163 


Frog  eggs,  protein  biosynthesis, 

mercurials,  887 
^-Fructofuranosidase   (invertase), 
analogs,  421 
dichromate,  660 
ferricyanide,  675 
hydrogen  peroxide,  691-692 
iodine,   683,   685,   688 
macroions,   465 
mercurials,    837-838 

increase  of  inhibition  by  thiols,  827- 

828 

pH  effects,  791-793 

protection  by  sucrose,  781 

rate  of  inhibition,  811-812 

spontaneous  reversal,  813 

type  of  inhibition,  772 
methylglucoside  anomers,   271 
oxidation  of,  657 
periodate,   660 
permanganate,  660 
succinyl   peroxide,   694 
/3-Fructofuranosyl-2-deoxyglucose,  forma- 
tion from  2-deoxyglucose,  389 
Fructokinase,  see  also  Hexokinase 
analogs,  376 
galactose,  376 
glucose,  376 
o-iodosobenzoate,  708 
mannose,  376 
mercurials,  837 
Fructose, 
aldolase,  407 
/^-amylase,  421 
/3-fructofuranosidase,  421 
a-galactosidase,  418 
glucosamine  phosphorylation  382 
glucose  uptake  by  lymph  node,  263 
a-glucosidase,  416 

intestinal  transport  of,  deoxypyridoxol, 
574 

a-mannosidase,  422 
oxidation  of,  2-deoxyglucose,  398 
phosphopentose  isomerase,  411 
uptake    by   lymph   node,   2-deoxyglu- 
cose, 394 
Fructose-l,6-diphosphatase, 

induction  of,   8-azaguanine,  478 


malonate,  60 

mercurials,  837 

nucleotides,  470 
Fructose-l,6-diphosphate, 

glucose  dehydrogenase,  410 

hexokinase,   379 

2-keto-3-deoxy-D-ara6o-heptonate-7-P 

synthetase,  413 
Fructose-phosphate,  aldolase,  407 
Fructose-6-phosphate, 

aldolase,  407 

glucose  dehydrogenase,  410 

hexokinase,  379 

phosphodeoxyribomutase,  413 

phosphopentose  isomerase,  411 
Fructose-phosphates,  structures  of,  378 
Fucono-l,5-lactone,   /J-galactosidase,  429 
Fucose,  a-galactosidase,  417 
Fucus  ceranoides, 

iodide  uptake,  mercurials,  910,  912 

respiration    (endogenous),    mercurials, 

881,  912 
Fumarase  (fumarate  hydratase,  L-malate 

hydro-lyase), 

trans-Rconitate,  273 

active  center  of,  274,  278 

analogs,  274-279 

binding  energies  of,  275-277 
pH  effects,  277-279 

a-hydroxy-/3-sulfopropionate,  243 

o-iodosobenzoate,  708,  717 
protection  by  substrate,  717 

macroions,  465 

malonate,   60,   65 

mercurials,  778,  781,  838 

protection  by  phosphate,  778 
protection  by  substrates,  781 
Fumarate, 

D-amino  acid  oxidase,  343 

antagonism    of    malonate    inhibitions 

112-117,  135 

aspartase,  355 

glutamate  decarboxylase,  328 

glutamate  dehydrogenase,  330-332 

intercharge  distance,  6 

ionization  constants,  8 

D-lactate  dehydrogenase,  437 

malic  enzyme,  596 


1164 


SUBJECT    INDEX 


oxalacetate  decarboxylase,  597 
oxidation  of,  malonate,  81 
phosphofructokinase,   385 
reduction  of,  malonate,  48-49 
succinate  dehydrogenase,  34-35,  38 
Fumarate  reductase,  malonate,  49 
Fundulus  heteroclitus  eggs,  cleavage  of, 

mercurials,  963 
Fungi, 

growth  of, 

dehydroacetate,  632-633 
2-deoxyglucose,  400 
ferrocyanide,  677 
malonate,  195-196 
oxythiamine,  520,  529 
pyrithiamine,  516,  528-529 
thiamine  analogs,  516,  520,  528-529 
tungstate,  614 
respiration    (glucose),    malonate,    133- 
134 
Furan-2-acrylate,  D-amino  acid  oxidase, 

342 
Furan-2-carboxylate,   D-amino  acid  oxi- 
dase, 242,  346 
2-Furoate, 

D-araino  acid  oxidase,  344 
glutamate  dehydrogenase,  331 
structure  of,  330 
Furyglycine,   glycine   uptake   by   ascites 

cells,  265 
Fusarium,  conidial  growth, 
p-benzoquinone,  660 
permanganate,  660 
Fusarium  decemcellulare,  mercurial  accu- 
mulation in  condia,  969 
Fusarium  graminearum ,  growth  of, 
dehydroacetate,  632 


GABA,  see  y-Aminobutyrate 
Galactarate    (mucate),    /3- glucuronidase, 

424,  427 
Galactoflavin, 

carcinostasis,  538 

flavin  levels  in  liver,  539-540 

flavokinase,  539 

glutamate  oxidation  in  vivo,  544 


/5-hydroxybutyrate  oxidation   in  vivo, 

544 

oxidative  phosphorylation  in  liver,  544 

riboflavin  deficiency,  538 

serotonin  metabolism,  544 

structure  of,  536 

succinate  oxidase,   543 
Galactono-l,5-lactone,  ^-galactosidase, 

429 
L-Galactono-y-lactone  dehydrogenase, 

mercurials,  protection  by  substrate,  781 

quinacrine,  547,  550 

riboflavin,  540,  542 
Galactose, 

^-amylase,  421 

fructokinase,  376-377 

fructose  uptake  by  ascites  cells,  263 

a-galactosidase,  417-418 

/?-galactosidase,  418 

glucose  uptake  by  lymph  node,  263 

a-glucoidase,  416-417 

intestinal  transport  of, 
analogs,  263 
6-deoxyglucose,  403 

mutarotase,  413-414 

oxidation  of,  see  also  Respiration  (ga- 
lactose) 

2-deoxyglucose,  398 

renal  transport  of,  glucose,  262 

uptake  of,  2-deoxyglucose,  394 
Galactose    dehydrogenase,    ferricyanide, 

675 
Galactose  oxidase,  quinacrine,  550 
Galactose- 1 -phosphate,   UDPglucose   py- 

rophosphorylase,  603 
Galactose-6-phosphate, 

hexokinase,  379-380 

phosphodeoxyribomutase,  413 
a  -  Galactosidase , 

analogs,  417-418 

5-fluorouracil,  inactivation  in  vivo  479- 

480 

inactive  form  induced  by,  479-480 

mercurials,  838 
/3-Galactosidase, 

analogs,  418-419,  429 

hydrogen  peroxide,  691-692 

iodine,  683,  685 


SUBJECT    INDEX 


1165 


mercurials,   838 

spontaneous  reversal,  814 
synthesis  of, 

azatryptophan,  326 
2-deoxyglucose,  400 
jralacturonate, 
a-glucuronidase,  426 
^-glucuronidase,  424,  427 
polygalacturonase,  421 
Vallate,  dehydroshikimate  reductase,  593, 

605 
Ganglia, 
acetylcholine  response,  mercurials,  949 
transmission   through, 
malonate,   211-212 
mercurials,  949 
Gardner  lymphosarcoma,  amino  acid  up- 
take, 

malonate,  155 
Gases,  see  also  specific  gases 
hydrogenase,  293-294 
nitrogen  fixation,  291-296 
Gasterosteus  aculeatus,   mercurials, 
respiration   (endogenous),   882 
toxicity,  963 
Gastric  acid  secretion, 
antimycin  A,  915-916 
2,4-dinitrophenol,  915-916 
malonate,  187,  208 
mercurials,  914-915 
Gastric  mucosa, 

respiration    (  endogenous  ),    malonate, 
175,   179 

respiration  (glucose),  mercurials,  883 
Gastrula,  respiration  (endogenous), 

mercurials,  882 
Gastrulation  see  also  Embryogenesis 
malonate,  198-199, 
mercurials,  964 
GDP,  see  Guanosinediphosphate 
Germination, 
Aspergillua  spores,  malonate,  195 
Bacillus  cereus,  D-alanine,  270 
Bacillus  subtilis, 
malonate,   195 
mercurials,  972 
quinacrine,  546 
fungi,  effect  on  malonate  inhibition  of 


respiration,    133-134 

Neurospora  ascospores,  malonate,   195 

Puccinia  uredospores,   malonate,   195- 

196 
Glucarate, 

/?-glucuronidase,  424,  427 

glucuronide  synthesis  in  liver,  428 

phosphatase  (acid),  442-443 
Glucaro-l,4-3,6-dilactone,     structure    of, 

425 
Glucaro- 1 ,4-lactone, 

/^-glucuronidase,  423-428 
in  vivo  inhibition,  428 

glucuronide  formation  in  liver,  428 

structure  of,  425 
Glucaro-3,6-lactone, 

/^-glucuronidase,  424,  427 

structure  of,  425 
/S-Glucofuranuronides,  structures  of,  425 
Glucokinase,  see  also  Hexokinase 

iV-acetyl-D-glucosamine,  390 

2-deoxyglucose,  389-390 

D-glucosamine,  390 

mannose,  376 
Gluconate,  /3-glucosidase,  417 
Gluconate  ethyl  ester,  ^-glucosidase,  417 
D-Gluconate  oxidase,  riboflavin,  542 
Gluconate-6-phosphate   dehydrase,    mer- 
curials, 885 
Gluconate  -  6  -  phosphate  dehydrogenase, 

mercurials,  838 
Gluconic-y-lactone,  /?-glucosidase,  417 
Gluconokinase,  mercurials,  781,  839 

protection  by  ATP,  781 

protection  by  gluconate,  781 
Glucono- 1 ,4-lactone, 

cellulytic  rumen  enzymes,  429 

a-glucuronidase,  426 

/3-glucosidase,  429 

isoamylase  (debranching),  429 

structure  of,  425 

thioglycosidase,  429 
Glucono-l,5-lactone, 

a-glucuronidase,  426 

/3-gluco8idase,  429 
/3-Glucopyranuronides,  structures  of,  425 
Glucosaccharo-l,4-lactone,    see    Glucaro- 

1,4-Iactone 


1166 


SUBJECT    INDEX 


D-Glucosamine, 

ATP  level  in  ascites  cells,  383 

glucokinase,    390 

glucose  oxidation  in  fat  pad,  382 

glucose  uptake  by  Scenedesmus,  383 

glycogen  formation  in  liver,  382 

hexokinases,  381-383,  390 

phosphoglucomutase,  382 

phosphopentose  isomerase,  411 

phosphorylase,  382 

phosphorylation   of, 

iV-acetylglucosamine,  382 
hexoses,  382 

pyruvate  oxidation,  383 

structure  of,  381 

UDPglucose-glycogen    glucosyltransfe- 

rase,  382 

UDPglucose  pycophosphorylase,  382 
D-glucosamine-6-phosphate, 

glucose-6-P  dehydrogenase,  411 

phosphoglucose  isomerase,  407 
D-Glucosamine-6-phosphate  deaminase, 

mercurials,  816 
D-Glucosamine  phosphokinase,  iV-acetyl- 

glucosamine,  593 
Glucose, 

aldolase,  407 

a-amylase,  420 

j3-amylase,  421 

arabinose  uptake  by  heart,  263 

blood  levels  of,  malonate,  149,  219 

2-deoxyglucose  uptake,  387,  389 

a-l,6-dextranglucosidase,  417 

distribution  of  C"  from  labeled,  ma- 
lonate, 130-132 

fermentation  by  yeast, 
D-glucosone,  384-385 
tripolyphosphate,  383 

/3-fructofuranosidase,  421 

fructokinase,  376-377 

galactose  transport  by  kidney,  262 

a-galactosidase,  417 

/5-galactosidase,  418 

glucosamine  phosphorylation  382 

glucose-6-phosphatase,  412 

a-glucosidase,  416-417,  423 

/3-glucosidase,  417 

intestinal  transport  of, 


deoxypyridoxol,  574 

malonate,  207 

mercurials,  916 
a-mannosidase,  422 
metabolism  of, 

malonate,  122-135 

malonic  diethyl  ester,  236-237 
oxidation  of,  see  also  Respiration  (glu- 
cose) 

6-deoxy-6-fluoroglucose,  393 

2-deoxyglucose,  393,  398 

6-deoxyglucose,  403 

D -glucosamine,  382 
phosphopentose  isomerase,  411 
phosphorylase,  405 
renal  transport  of, 

dehydroacetate,  625 

malonate,  205 

mercurials,  920 
uptake  of, 

analogs,  263 

6-deoxy-6-fluoroglucose,  404 

2-deoxyglucose,  390,  394 

mannoheptulose,  376 

mercurials,  893-894,  910-912 
utilization  of, 

cystamine,  663 

dehydroacetate,  624 

ferrocyanide,  678 

mercurials,  884,  903-905 

oxamate,  435 

quinacrine,  560 
Glucose  dehydrogenase, 
analogs,  410,  500-502 
benzoate,  501 
malonate,  61 
mercurials,  838 
nicotinamide  analogs,  500-502 
nucleotides,  501 
pyridoxal,  501-502 
quinacrine,  550 
Glucose  - 1 ,6  -  diphosphates,    hexokinase, 

379 
Glucose  oxidase,  pterin-6-aldehyde,  288 
Glucose-6-phosphatase, 
analogs,  412 
2-deoxyglucose,  390 
induction  of,  8-azaguanine,  478 


SUBJECT    INDEX 


1167 


mercurials  in  vivo,  927 
Glucose- 1  -phosphate, 

glucose  dehydrogenase,  410 

hexokinase,  379-380 

level  in  parasitized  erythrocytes,  qui- 

nacrine,  560 

phosphatase,  439 

phosphorylase,  405 
Glucose-6-phosphate, 

fructokinase,  381 

glucose  dehydrogenase,  410 

hexokinase,  377,  379,  381 

phosphoarabinose  isomerase,  411 

phosphodeoxyribomutase,  413 

phosphoglucomutase,  413 

phosphopentose  isomerase,  411 

phosphorylase,  405 
Glucose-6-phosphate  dehydrogenase, 

3-acetylpyridine-NAD,  497 

analogs,  411 

l,5-anhydro-D-glucitol-6-P,  379 

2-deoxyglucose-6-P,  390-391 

o-iodosobenzoate,  608 

mercurials,  839 

protection  by  NADP,  781 

nicotinamide,  503 

nucleotides,  508 

quinacrine,  550 
Glucose-phosphate   isomerase,   see   Phos- 

phoglucose  isomerase 
Glucose-phosphates,  structures  of,  378 
Glucose  respiration,  see  Respiration  (glu- 
cose) 
a-Glucosidase,  see  also  Maltase 

analogs,  416-417,  423 

mercurials,  772 
/3-Glucosidase, 

analogs,  417 

mercurials,  839 

protection  by  substrates,  782 
D-Glucosone, 

Crabtree  effect,  385,  397 

formation  of,  385 

glucose    fermentation    in    yeast,    384- 

385 

glycolysis  (anaerobic),  385 

hexokinase,  384-385 

hyperglycemia,  384 


lethal  doses,  384 

structure  of,  383-384 

toxicity,  384 
L-Glucosone,     glucose     fermentation     in 

yeast,  384 
Glucuronate, 

a-glucuronidase,  426 

/^-glucuronidase,  424,  427 

hexokinase,  380 

phosphatase  (acid),  442 

structure  of,  425 
Glucuronate  - 1  -  phosphate,    a  -  glucuroni- 
dase, 426 
Glucuronate-6-phosphate,  hexokinase, 

380 
D-Glucurone, 

a-glucuronidase,  426 

/^-glucuronidase,  424,  427 

glucuronide  synthesis  in  liver,  428 

structure  of,   425 
a  -  Glu  curonida  se , 

analogs,  426 

D-glucurone,  426 

a-and  /J-glucuronides,  426 
/S-Glucuronidase, 

analogs,   423-428 

copper,  795 

inhibition   in   urinary   bladder   cancer, 

428 

malonate,  61 

menthyl-a-glucuronide,  272 

mercurials,  839 

pH  effects,  791,  793,  795-797 
type  of  inhibition,  772 

silver,  795 
Glucuronidases, 

heparin,  465 

hyaluronate,  465 
Glutamate, 

cysteine  desulfurase,  357 

glutaminase,  332 

metabolism  of, 
analogs,  327-336 
malonate,  152 
pathways,  327 

oxidation  of, 

galactoflavin  in  vivo,  544 
mercurials,  878 


1168 


SUBJECT    INDEX 


D-Glutamate, 

L-glutamate  decaroxylase,  269 
L-glutamate  dehydrogenase,  330,  332 
L-glutamine  synthetase,  269,  336 
y-glutamyltransferase,  336 
L-Ghitamate, 

L-amino  acid  oxidase,  340 
D-glutamate  oxidase,  336 
phosphatase  (acid),  441-442 
pyridoxamine  -  oxalacetate      transami- 
nase, 600 
L-Glutamate  :  ammonia  Hgase  (ADP),  see 

Glutamine  synthetase 
L-Glutamate  decarboxylase, 
acetate,  328 
analogs,  327-329 
cafFeate,  314 
D- cycloserine,  359 
deoxypyridoxol  in  vivo,  569 
D-glutamate,  269 
mercurials,  772,  782,  811-812 

protection  by  pyridoxal-P,  782 

rate  constant  for  inhibition,  811-812 

temperature  effects,  811 

type  of  inhibition,  772 
permanganate,  660 
toxopyrimidine,  578 
L-Glutamate  dehydrogenase, 
analogs,  329-332 
ferric  ions,  863 
o-iodosobenzoate,  708 
p-iodosobenzoate,  702 
malonate,  61 

mercurials,   789,    793,   816,   819,   826, 
840-841,  863 

pH  effects,  793 

reversal  with  GSH,  826 

splitting  into  subunits,  789,  819 

stimulation,  816,  819 
nicotinamide,  863 
nucleotide  binding  sites,  514 
nucleotides,  508,  514 
oxygen  inactivation  of,  659 
silver,  863 
zinc,  863 
D-Glutamate  oxidase, 
benzoate,  349 
L-glutamate,  336 


o-iodosobenzoate,  709 

mercurials,  840 
Glutamate  racemase, 

FMN,  542,  544 

riboflavin,  542,  544 
Glutamate  semialdehyde  reductase, 

o-iodosobenzoate,  709,  717 
protection  by  substrate,  717 

mercurials,  841 

nucleotides,  507 
Glutamate  transaminases,  see  also  Trans- 
aminases 

analogs,  334 
Glutaminase    (L-glutamine    amidohydro- 

lase), 

active  center  of,  332 

analogs,  332-333,  356 

mercurials,  841 

quinacrine,  550 
Glutaminase  I,  quinacrine,  550 
Glutaminase  II,  quinacrine,  551 
L-Glutamine, 

brain  level,  2-deoxyglucose,  399 

L-glutamate  dehydrogenase,  331 
L-Glutamine  amidohydrolase,  see  Gluta- 
minase 
Glutamine :  fructose  -  6  -  phosphate    trans- 

amidase,  DON,  356 
Glutamine:pyruvate     transaminase,     see 

Transaminases 
Glutamine   synthetase   (L-glutamate:am- 

monia  ligase), 

ADP,  471 

analogs,  336 

D-glutamate,  269 

GSSG,  662 

mercurials,  841 

methionine  sulfoximine,  335 
y-Glutamylalanylglycine,  glyoxylase,  594 
y-Glutamyl-/3-sulfoalanylglycine,  glyoxyl- 
ase, 594 
y-Glutamyltransferase, 

analogs,  336 

malonate,  61 
Glutarate, 

aspartate :  a  -  ketoglutarate      transami- 
nase, 334 

carcinostasis,  201 


SUBJECT    INDEX 


1169 


cysteine  desulfurase,  357 

fumarase,  275 

glutamate  decarboxylase,  328 

glutamate  dehydrogenase,  330-332 

hydroxamic  acid  formation  from,  233 

intercharge   distance,   5-6 

ionic  length,  188 

ionic  volume,  188 

ionization    constants,    8 

kynureninera-ketoglutarate     transami- 
nase,   607-609 

lethal  doses,  201 

permeability  of  erythrocytes  to,  188 

pyridoxamine:    oxalacetate    transami- 
nase, 600 

succinate  dehydrogenase,   35 

urinary  citrate,  106,  109 

urinary  a-ketoglutarate,  110-111 
Glutathione, 

erythrocyte  levels,  mercurials,  905-906 

glycolysis  (aerobic),   637 

role  in  enzyme  and  metabolic  activity, 

636-637 

tissue  levels,  tetrathionate,  696,  700 
Glutathione  (oxidized), 

enzyme  inhibitions,  662-663 

succinate  dehydrogenase,  temperature 

effects,   663-664 
Glutathione   oxidase,   thioglycolate,   593 
Glutathione   reductase,   mercurials,   773, 

841 
Gly  ceraldehy  de , 

glucose  phosphorylation,  377 

glycolysis,    377 
Glyceraldehyde-3-pho8phate,  see  3-  Phos- 

phoglyceraldehyde 
Glyceraldehyde-3-phosphate  dehydrogen- 
ase, see  3-Phosphoglyceraldehyde  de- 
hydrogenase 
Glycerate,  phosphatase  (acid),  441 
Glycerate  dehydrogenase, 

bromopyruvate,  430 

mercurials,  817,  841 

phenylpyruvate,  430 

pyruvate,  430 
Glycerate-2,3-diphosphatase, 

mercurials,  stimulation,  817-818,  820 

phosphoglycerates,  413 


Glycerate  kinase, 

o-iodosobenzoate,  709 

mercurials,  841 
GIycerate-2-phosphatase,  mercurials,  817 
Glycerate-3-phosphatase,  mercurials,  817 
Glycerol, 

metabolism  of,  malonate,   164 

permeability  to,  mercurials,  906-907 
Glycerol-2-phosphate,  enolase,  409 
Glycerol-3-phosphate  dehydrogenase, 

hydrogen  peroxide,   693 

mercurials,  relation  to  SH  groups,  804 
Glycerol-3-phosphate  oxidase,   hydrogen 

peroxide,  693 
/5-Glycerophosphatase, 

dichromate,  660 

iodine,  685 

mercurials,  860 

inhibition  in  vivo,  926-927 

permanganate,  660 
/?- Glycerophosphate,  phosphatases,  439 
a-Glycerophosphate  dehydrogenase, 

malonate,  61 

mercurials,  842 

porphyrexide,  668 
Glycine, 

L-alanine   dehydrogenase,   354 

L-amino  acid  oxidase,  340 

arginase,  337 

dipeptidase,  368 

intestinal  transport  of,  analogs,  265 

phosphatase  (acid),  441 
Glycine  iV-acylase,  hippurates,  355 
Glycine:a-ketoglutarate  transaminase,  see 

Transaminases 
Glycine   methyltransferase,   o-iodosoben- 
zoate, 709 
Glycogen, 

formation  of, 

2-deoxyglucose-6-P,  391 
D-glucosamine,  382 

level  in  diaphragm,  mercurials,  884 
Glycolaldehyde-2-phosphate,  formation  of 

threose-2,4-diphosphate  from,  408 
Glycolate  (hydroxyacetate), 

glyoxylate  transacetatse,  594 

tartronate  semialdehyde  reductase,  602 


1170 


SUBJECT    INDEX 


Glycolate  oxidase, 

analogs,  438 

diphenylglycolate,  593 

o-iodosobenzoate,  709 

malonate,  61 

mercurials,  860 
Glycolysis,   see  also  Glycolysis   (aerobic) 

and  Glycolysis  (anaerobic) 

o-iodosobenzoate,  704 

ionic  regulation  of,  453 

malonate,  125-130,  134-135 
effect  of  fumarate,   126 

mercurials,  874-877,  884 

role  of  GSH  in,  637 

tetrathionate,  699 

threose-2,4-diphosphate,  409 
Glycolysis  (aerobic), 

2-deoxyglucose,  391-394 

ferricyanide,  677 

hydrogen  peroxide,  695 

oxamate,  434 

quinacrine,  560 

tartronate,  238 
Glycolysis  (anaerobic), 

dehydroacetate,  624 

2-deoxyglucose,  391-392 

ferricyanide,  673,  677 

hydrogen  peroxide,  695 

iodine,  689 

o-iodosobenzoate,  721 

macroions,  465 

oxalate,  414 

oxamate,  434 

phosphate,  414 

ribonucleonate,  414 

sulfate,  414 

tartronate,  238 
Glycosidase,   analogs,  415-429 
a,/3-Glycosylphosphatase,  mercurials,  842 
Glycylglycine,  D-amino  acid  oxidase,  340 
Glycylglycine  dipeptidase,  mercurials,  842 
Glycylleucine,  D-amino  acid  oxidase,  340 
Glycylphenylalanine,  cathepsin  C,  375 
Glycyltyrosine,  carboxypeptidase,  367 
Glyoxal,  tartronate  semialdehyde  reduc- 
tase, 602 
Glyoxalate,  see  Glyoxylate 
Glyoxylase,  analogs,  593-594 


Glyoxylate, 

cycle    inhibition    by    formation   of   y- 

hydroxy-a-ketoglutarate,  615-616 

metabolism  of,  malonate,  165 

pyruvate  decarboxylase,  431 

tartronate  semialdehyde  reductase,  602 
Glyoxylate  cycle,  role  in  malonate  inhibi- 
tion of  tricarboxylate  cycle,  71-72 
Glyoxylate  reductase, 

analogs,  438 

malonate,  61 

mercurials,  842 

protection  by  glyoxylate,  782 
protection  by  NADH,  782 
Glyoxylate  transacetase,  analogs,  594 
GMP,  see  Guanosinemonophosphate 
Goldfish  gills,  Na+  transport, 

mercurials,  909 
Granosan  M,  toxicity,  954 
Growth,  see  also  specific  organisms  and 

tissues 

2-deoxyglucose,  400-401 

malonate,  196-202 

mercurials,  963-970 

thiamine  analogs,  527 
GSH,  see  Glutathione 
GSSG,  see  Glutathione  (oxidized) 
GTP,  see  Guanosinetriphosphate 
GTPase,  see  Guanosinetriphosphatase 
Guaiacol,     dehydroshikiniate    reductase, 

593,  605 
Guanase  (guanine  deaminase), 

pterin-6-aldehyde,  288 

xanthopterin,  288 
Guanidine,  histidase,  353 
Guanidines, 

diamine  oxidase,  360-365 

structures  of,  361 
Guanidinium  ion,  resonance  structures  of, 

361 
Guanidinovalerate,  carboxypeptidase,  367 
Guanine, 

adenylosuccinate  synthetase,  467 

ATPase,  445 
Guanine  deaminase,  see  Guanase 
Guanosine, 

adenylosuccinate  synthetase,  467 

5 '-nucleotidase,   472 


SUBJECT    INDEX 


1171 


Guanosinediphosphate  (GDP), 
adenylosuccinate  synthetase,  467 
glutamate  dehydrogenase,   508 
IMP  dehydrogenase,  471 
isocitrate  dehydrogenase,   509 
NADH  oxidase,  511 
polynucleotide  phosphorylase,  474 

Guanosinemonophosphate    (GMP), 
adenylosuccinate  synthetase,  467 
deoxycytidylate  deaminase,  469 
IMP  dehydrogenase,  471 
phosphatase,  439 
ribonuclease,  475 

Guanosinemonophosphate  reductase,  nu- 
cleotides, 471 

Guanosinemonophosphate  sjTithetase,  see 
Xanthosine-5 '-phosphate  aminase 

Guanosine  phosphorylase,  mercurials,  842 

Guanosinetriphosphatase  (GTPase), 
ADP,  446 
IDP,  446 

Guanosinetriphosphate  (GTP), 

asprtate  carbamyltransferase,  468 
glutamate  dehydrogenase,  514 
IMP  dehydrogenase,  471 
pyrophosphatase,  475 

Gulonate  dehydrogenase,  mercurials,  842 

Gulose,  fructokinase,  376 

Gymnodinium  nelsoni,  respiration  (endo- 
genous), 
malonate,  169 

Gynaecotyla    adunca,    respiration    (endo- 
genous), 
malonate,  173,  183 

H 

Hadacidin,  adenylosuccinate  synthetase, 

467 
Heart, 

acetate  accumulation,  malonate,  140 

acetate  oxidation,  mercurials,  878,  88r 

acetate  utilization,  propionate,  613 

acetylcholine  response, 
malonate,  217 
mercurials,  946-947 

acetylpyridines,  494,  499-500 

aminomalonate  decarboxylation  in,  239 


arabinose  uptake,  glucose,  263 
catecholamine  levels  in, 

a-methyldopa,  317-328 

a-methyl-m-tyrosine,  317-318 
C-l/C-6  ratio,  malonate,  130-131 
citrate  accumulation, 

malonate,   104 

mercurials,  927 
citrate  levels  in  vivo,  sequential  inhibi- 
tion by  malonate  and  fluoroacetate,  112 
citrate  oxidation,  malonate,  79 
conduction, 

dehydroacetate,  625 

o-iodosobenzoate,  724 

mercurials,  942-945 
contractility, 

2-deoxyglucose,  402-403 

o-iodosobenzoate,   723 

malonate,  128,  213-215,  217 

mercurials,  941-946 

porphyrindin,  669 
contracture, 

mercurials,  941-944 

porphyrindin,  669 
dehydroacetate,  positive  inotropic  ac- 
tion of,  625 
electrocardiogram, 

malonate,  215 

mercurials,  945 
epinephrine  response, 

malonate,  217 

mercurials,  947 
fumarate  oxidation,  malonate,  81 
glucose  metabolism,  malonate,   131, 
216 

glucose  uptake,  3-methylglucose,  263 
glycolysis  (anaerobic),  malonate,  128 
a-ketoglutarate  accumulation,   malon- 
ate, 111 
a-ketoglutarate  oxidation, 

malonate,  80,  84 

mercurials,  879 
K+  fluxes,  2-deoxyglucose,  403 
K+/Na+  ratio,  mercurials,  946 
K+  uptake  in  mitochondria,  mercurials, 
909 
malate  oxidation, 

malonate,  82 


1172 


SUBJECT    INDEX 


mercurials,  879 
malonate, 

compared  with  ouabain,  216 

decarboxylation  of,  232-233 

formation  in,  226 

levels  in  vivo,  102 

metabolism  of,  216,  228 

pH  effects,  191 
malonic  ethyl  ester,  236 
malonyl-CoA  formation  in,  233-234 
membrane  potentials, 

dehydroacetate,  625 

2-deoxyglucose,  403 

malonate,  214 

mercurials,  896,  945-946 
mercurials  in  vivo,  930,  940-941,  943- 
945,  959 

methylmalonate  in,  224 
nicotinamide,  500 

ouabain  stimulation  of,  2-deoxyglucose, 
402 

oxalacetate  oxidation,  malonate,  82 
oxidative  phosphorylation, 

malonate,  121-122 

mercurials,  873-874 
permeability  to  Ca++,  mercurials,  947 
pyruvate  oxidation, 

acetylene-dicarboxylate,   241 

malonate,  75-77,  216 

mercurials,  878-879 

propane-tricarboxylate,  241 
rate  of  beating, 

dehydroacetate,  625 

o-iodosobenzoate,  723 

malonate,  128,  213-215 

mercurials,  941-944 

porphyrindin,  669 

thiamine  analogs,  527 
refractory  period,  malonate,  214 
respiration,  kojic  acid,  350 
respiration     (endogenous),     malonate, 
176-179,  181 
respiration  (glucose), 

l,5-anhydro-D-glucitol-6-P,  379 

mercurials,  883,  948 
ribose  metabolism,  malonate,  132 
succinate  accumulation,  malonate,  91, 
94-95,  128 


succinate   dehydrogenase, 
alkylmalonates,  37 
dicarboxylate  ions,  35-39 
malonate,  22-24,  31-32,  47-48 
succinate  levels  in  vivo,  malonate,  100- 
102 

succinate  oxidation,  malonate,  55 
thiamine-PP   level,   thiamine   analogs, 
525-527 

tissue   culture   cells   from,    mercurials, 
968 

transamination      in,     deoxypyridoxol, 
569-570 

tricarboxylate   cycle,   mercurials,   877- 
879 
vagal  effects  on,  mercurials,  946-947 

HeLa  cells, 

2-deoxyglucose-6-P  oxidation,  388 
glucose  uptake,  oxamate,  434 
glycolysis  (aerobic),  oxamate,  434 
growth  of,  oxamate,  434 
mitosis  in,  mercurials,  969 
Na+  pump,  oxamate,  434 
resistance  to  2-deoxyglucose,  388 
RNA  biosynthesis,  969 

Helianthus  annus,   malonate   occurrence 
in,  225 

Helianthus    tuberosus    (Jerusalem     arti- 
choke), respiration  (endogenous), 
malonate,   172 

Helium,  nitrogen  fixation,  291 

Helix  pomatia  hepatopancreas, 

cycle  intermediates  oxidations,  malon- 
ate, 79-81 

respiration,  malonate,  114-115,  174 
succinate  oxidation,  malonate,  54 

Heme,  biosynthesis  of, 
mercurials,  888 

Hemicentrotus  pulcherrimus  eggs,  devel- 
opment of, 
ferricyanide,  678 
o-iodosobenzoate,  727 

Hemoglobin, 

biosynthesis    of,    a-amino-^-chlorobu- 
tyrate,  351 
ferricyanide,   670-671 
mercurials,  649,   755-757,   760 

Hemolysis,  mercurials,  900-907 


SUBJECT    INDEX 


1173 


Hemophilus     parainfluenza  e,     glutamate 

oxidation, 

malonate,   152 
Hensenula  ellipsoidospora,  growth  of, 

malonate,  195 
Heparin, 

amylases,  464 

cell  division,  462 

/3-fructofuranosidase,   465 

fumarase,  465 

glucoronidases,  465 

hyaluronidase,  459-460 

lipoprotein  lipase,  463 

lysozyme,  459 

ribonuclease,  461 

trypsin,  456 
Hepatomas, 

malonate  levels  in  vivo,  102 

succinate  dehydrogenase,  malonate,  30, 

50 

succinate  levels  in  vivo,  malonate,  102 
Heptanoate,    kynurenine: a-ketoglutarate 

transaminase,  608 
Herpes  virus, 

inactivation  by  mercurials,  978 

proliferation  of,  mercurials,  981 
Hevea  brasiliensis,  malonate, 

formation  in,  226 

occurrence  in,  225 
Hexanoate,  see  Caproate 
Hexokinase,  see  also  Fructokinase,  Glu- 

cokinase,  and  other  kinases 

iV-acetyl-D-glucosamine,  390 

analogs,  376-386 

competition   between  substrates,   376- 

377 

6-deoxy-6-fluoroglucose,  404 

2-deoxyglucose,   389-390 

D-glucosamine,  391-383,  390 

D-glucosone,   383-385 

hexose-phosphates,  377-381 

mercurials,  782,  788,  806-807,  824,  842- 

843 

denaturation,  788 
protection  by  glucose,  782 
protection   by  Zn++,   782 
reversal  ^\•ith  cysteine,  824 
titration  of  SH  groups,  806-807 


nucleotides,  383 

quinacrine,  551 

tripolyphosphate,  383 
Hexylamine,  leucine  aminopeptidase,  368 
4-Hexylresorcinol,  tyrosinase,  304 
Hill  reaction,  mercurials,  891 
Hippurate,  glycine  iV-acylase,  355 
Histaminase,  see  Diamine  oxidase 
Histamine, 

metabolism  of,  aminoguanidine,  363 

release  of, 

chymotrypsin  inhibitors,  374 
hydrocinnamate,  374 
indole,  374 

3-indolepropionate,  374 
o-iodosobenzoate,  725 
malonate,  212-213 
mercurials,  949 
skatole,  374 
Histidase, 

analogs,  353 

D-histidine,  269 
D-Histidinal,  L-histidine  formation,  269 
Histidine, 

diamine  oxidase,  365 

phosphoribosyl  -  ATP     pyrophosphory- 

lase,  351 
D-Histidine, 

L-amino  acid  oxidase,  340 

histidase,  269 
L-Histidine, 

arginase,  337 

dipeptidase,  367 

homoserine  kinase,  357 

uptake  by  Botrytis,  analogs,  267 
Histidine  decarboxylase, 

analogs,  352-353 

caflfeate,  314 

permanganate,  660 
Histidine  hydrazide,  histidine  decarboxy- 
lase, 353 
D-Histidinol,   L-histidine   formation,   269 
Homobiotin, 

biotin  oxidase,  589 

structure  of,  588 
Homocysteine, 

serine  deaminase,  357 

tyrosine  decarboxylase,  307 


1174 


SUBJECT    INDEX 


Homogentisate,     p  -  hydroxyphenylpyru- 

vate  oxidase,  306 
Homogentisate  oxidase, 

o-iodosobenzoate,  709,  717 

mercurials,  relation  to  Fe++,  771-772, 

787 
Homogentisicase, 

ferricyanide,  673,  675 

o-iodosobenzoate,    protection   by   sub- 
strate, 717 

mercurials,  782 

protection  by  Fe++,  782 
protection  by  homogentisate,  782 
Homolog,  definition  of,  246 
Homooxybiotin,  yeast  fermentation,  588- 

589 
D-Homoserine,  L-threonine  synthetase,  357 
DL  -  Homoserine,      thetin  :  homocysteine 

transmethylase,  357 
L-Homoserine,  aspartokinase,  356 
Homoserine  deaminase, 

o-iodosobenzoate,  709 

mercurials,  843 
Homoserine  kinase,  analogs,  357 
Hordeum  vulgar e  (cape  barley),   see  also 

Barley  roots 

succinate  dehydrogenase,  malonate,  27 
Hormosira  hanlcsii, 

K+  uptake,  mercurials,  908 

respiration    (endogenous),    mercurials, 

881 
Horseshoe  crab,  see  Limulus 
Hyaluronate, 

glucoronidases,  465 

lysozyme,  459 

ribonuclease,  462 
Hyaluronidase,  macroions,  459-461 
Hydrazine,  diamine  oxidase,  362 
Hydrocinnamate,    see  also    /3-Phenylpro- 

pionate 

acetoacetate  formation  from  butyrate, 

613 

D-amino  acid  oxidase,  342-343 

ammonia  formation  in  kidney,  348 

carboxypeptidase,  365-366 

chymotrypsin,  369-370,  372 

dopa  decarboxylase,  311 

histamine  release,  374 


Hydrocinnamide,  chymotrypsin,  372 
Hydrogen, 

formation  in  Chlorella,  mercurials,  891 

nitrogen  fixation,  292-293 

physical  properties,  295 
Hydrogenase, 

analogs,  293-294 

ferricyanide,  676 

iodine,  685 

o-iodosobenzoate,  709 

malonate,  61 

quinacrine,  551 
Hydrogen  bonding, 

reactivity  of  enzyme  SH  groups,  644- 

646 

SH  groups,  640 
Hydrogenlyase, 

hydrogen  peroxide,  693 

permanganate,  660 
Hydrogenomonas  facilis,  COg  fixation, 

mercurials,  892 
Hydrogen  peroxide,  see  also  Peroxides 

GSH  oxidation  by,  694 

potentiation  of  azide  toxicity  696 

potentiation  of  toxicity  by  iodoacetate, 

696 
Hydroquinone, 

D-amido  acid  oxidase,  344 

catechol  oxidase,  298 

tyrosinase,  304 
Hydroquinone     monobenzyl     ether,     see 

Monobenzone 
Hydroxocobalamin,  diol  dehydrase,  590 
Hydroxyacetate,  see  Glycolate 
D-a-Hydroxy  acid  dehydrogenase, 

analogs,  436-437 

malonate,  61,  65 

oxalate,  435-437 
L-a-Hydroxy  acid  oxidase, 

o-iodosobenzoate,  709 

mercurials,  843 

quinacrine,  551 
L-3-Hydroxyacyl-CoA     hydro-lyase,     see 

Enoyl-CoA  hydratase 
3  -  Hydroxy  -  N  -  allylnormophinan,     mor- 
phine iV-demethylase,  597 
2-Hydroxy-6-aminopurine,  xanthine  oxi- 
dase, 282 


SUBJECT    INDEX 


1175 


p-Hydroxyamphetamine,     phenylalanine 
/3-hydroxylase,  599 

17/S  -  Hydroxyandrosta  - 1 ,4  -  diene  -  3  -  one, 
Zl*-3-ketosteroid  reductase,  450 

3-Hydroxyanthranilate  oxidase, 
anthranilate,  594 
o-iodosobenzoate,  709 
mercurials,   843 

protection  by  Fe++,  782 
protection  by  substrate,  782 
type  of  inhibition,  772 

Hydroxyaspartate,  aspartate :  a  -  ketoglu- 
tarate  transaminase,  355 

p-Hydroxybenzaldehyde,  succinate  semi- 
aldehyde  dehydrogenase,  601 

p-Hydroxybenzamide,  sulfanilamide  ace- 
tylase,  601 

rw-Hydrobenzoate,  dehydroshikimate  re- 
ductase, 606 

o-Hydroxybenzoate,  see  Salicylate 

p-Hydroxybenzoate, 

dehydroshikimate  reductase,  593,  605 
p-hydroxyphenylpyruvate  oxidase,  306 
tyrosine:a-ketoglutarate  transaminase, 
305-306 

Hydroxybenzoates, 

D-amino  acid  oxidase,  341 
catechol  oxidase,  297-299 
glutamate  dehydrogenase,  331 

p-Hydroxybenzyloxyamine,  dopamine  /?- 
hydroxylase,  320 

2-Hydroxy-5-bromobenzamide,     sulfanil- 
amide acetylase,  601 

/5-Hydroxybutyrate, 

conversion  to  acetoacetate,  6-aminoni- 
cotinamide  in  vivo,  505 
oxidation  of, 

galactoflavin  in  vivo,  544 

iodine,  689 

mercurials,  883 

thiamine  analogs  in  vivo,  520-521 

/3-Hydroxybutyrate  dehydrogenase, 
analogs,  594 
malonate,  61 
mercurials,  843 

3-Hydroxycinnamate, 
dopa  decarboxylase,  314 
glutamate  decarboxylase,  314 


histidine  decarboxylase,  314 
succinate  dehydrogenase,  314 
tyrosine  decarboxylase,  314 

a-Hydroxyethanesulfonate,  glycolate  oxi- 
dase, 438 

9-(2'-Hydroxyethyl)riboflavin  analog, 
flavokinase,   539 
xanthine  oxidase  in  vivo,  544 

/9-Hydroxyglutamate, 
glutaminase,  333 
glutamine  synthetase,  336 

Hydroxy   group,    interaction   energy   of, 
276 

p-Hydroxyhippurate,  glycine  iV-acylase, 
355 

^-Hydroxyisobutyrate    dehydrogenase, 
mercurials, 

protection  by  NAD,  782 
protection  by  substrate,  782 

a-Hydroxyisocaproamide,  leucine  amino- 
peptidase,  368 

a-Hydroxyisocaproate,  glutamate  decar- 
boxylase, 329 

a-Hydroxyisovalerate,   glutamate  decar- 
boxylase, 329 

y-Hydroxy-a-ketoglutarate, 
aconitase,  615-616 

formation  from  glyoxylate  and  oxala- 
cetate,  615 

isocitrate  dehydrogenase,  615-616 
a-ketoglutarate  oxidation,   616 
respiration  (endogenous),   616 
succinate  oxidation,  616 
tricarboxylate  cycle,  615-616 

}'-Hydroxy  -  a  -  ketoglutarate    synthetase, 
mercurials,  843 

Hydroxylamine     reductase,     mercurials, 
787 

f5-Hydroxylysine,   glutamine  synthetase, 
336 

Hydroxymalonate,  see  Tartronate 

Hydroxymethanesulfonate,  glycolate  oxi- 
dase, 438 

2-Hydroxymethylene-17a-methylandro- 
stan-17/S-ol-3-one,  ^-hydroxysteroid  de- 
hydrogenase, 449 

a-Hydroxy-/3-oxalosuccinate,  see    Oxalo- 
malate 


1176 


SUBJECT    INDEX 


Hydroxyphenylacetates, 

glutamate  decarboxylase,  329 
tyrosine:a-ketoglutarate  transaminase, 
305-306 

Hydrox  yphenylalanines,  tyrosine  -  a  -  keto- 
glutarate  transaminase,  305-306 

p-Hydroxyphenyllactate,  p-hydroxyphe- 
nylpyruvate  oxidase,  306 

m-Hydroxyphenylpyruvate,  p-hydroxy- 
phenylpyruvate  oxidation,  272,  306, 
595 

p-Hydroxyphenylpyruvate, 
glutamate  decarboxylase,  329 
oxidation    of,    m- hydroxy phenylpyru- 
vate,  272,  306,  595 
peroxidase,  599 
pyruvate  decarboxylase,  431 

Hydroxyphenylpyruvates,  dopa  decar- 
boxylase, 312 

p-Hydroxyphenylpyruvate  oxidase, 
analogs,  305-306 
m-hydroxypenylpyruvate,  306,  595 

Hydroxyproline, 

A  i-pyrroline-5-carboxylate   dehydroge- 
nase, 336,  355 

A  i-pyrroline-5-carboxylate     reductase, 
355 

/3-Hydroxypropionate-phosphate,cnolase, 
410 

/5-Hydroxypropionitrile,  structure  of,  41 

Hydroxypurines,  keto-enol  equilibria,  280 

3-Hydroxypyridine,  glucose  dehydroge- 
nase, 501 

a-Hydroxy-2-pyridine  methanesulfonate, 
glycolate  levels  in  tobacco,  439 
glycolate   oxidase,   438 
photosynthesis,  439 
structure  of,  438 

Hydroxypyruvate, 

lactate  dehydrogenase,  437 
pyruvate   decarboxylase,   432 

Hydroxypyruvate  reductase,  mercurials, 
protection  by  hydroxypyruvate,  782 
protection  by  NADH,  782 

a-Hydroxysteroid  dehydrogenase,  estra- 
trienes,  449 

)5-Hydroxysteroid  dehydrogenase,  ana- 
logs, 447,  449,  450 


a-Hydroxy-/S-sulfopropionate,    fumarase, 
243,  275,  277 

5-Hydroxytryptamine,  see  Serotonin 

5  -  Hydrox  y  try  ptophan , 

histidine  decarboxylase,  353 
tryptophan  pyrrolase,  324-325 
L-tryptophan:sRNA  ligase  (AMP),  326 

7-Hydroxytryptophan,   tryptophan   pyr- 
rolase, 324 

5-Hydroxytryptophan  decarboxylase, 
o-iodosobenzoate,  709,  718 
a-methyldopa,  309-310 

Hymenolepsis  diminuta, 

malonate  metabolism  in,  228 
respiration  (endogenous),  malonate,  173 
succinate  dehydrogenase,  malonate,  28 

Hypochlorite,  amino  acid  oxidation  by, 
658 

Hypophosphate, 

oxidative  phosphorylation,  448 
succinate  dehydrogenase,   243 

Hypophosphite, 

formate  dehydrogenase,  593 
formate  hydrogenlyase,  593 

Hypoxanthine, 

D -amino  acid  oxidase,  545 

urate  uptake  by  erythrocytes,  267 


Idose,   fructokinase,  376 

IDP,  see  Inosinediphosphate 

Imidazole, 

diamine  oxidase,  365 
histidase,  353 

histidine  decarboxylase,  352 
A  i-pyrroline-5-carboxylate   dehydroge- 
nase, 355 

A  i-pyrroline-5-carboxylate     reductase, 
355 

Imidazoleacetate,  histidase,  353 

Imidazole-iV^-methyltransferase,   mercu- 
rials, 843 

Imidazolonepropionate  hydrolase,  hydro- 
gen peroxide,  693 

Imidodipeptidase,  see  Prolidase 

IMP,  see  Inosinemonophosphate 

Indene,  tryptophanase,  324 


SUBJECT    INDEX 


1177 


Indigo f era  endecaphylla,  3-nitropropionate 

the  toxic  principle  in,  244 
Indole, 

acetylindoxyl  oxidase,  591 

chymotrypsin,  371,  374 

histamine  release,  374 

histidine  decarboxylase,  352 

tryptophanase,  323-324 

tryptophan  pyrrolase,  325 

L-tryptophan:sRNA  ligase  (AMP),  326 
3-Indoleacrylate,    tryptophan    pyrrolase, 

325 
3-Indoleacetate, 

D-amino  acid  oxidase,  342 

carboxypeptidase,  366 

chymotrypsin,  371 

histamine  release,  374 

tryptophanase,  323 

tryptophan  pyrrolase,  325 

uptake  by  Avena  coleoptile,  mercurials, 

911 
Indoleacetate  oxidase,  analogs,  595 
3-Indolealdehyde,  acetylindoxyl  oxidase, 

591 
3-Indolebutyrate, 

chymotrypsin,  370-371 

tryptophan  pyrrolase,  325 
2-Indolecarboxylate, 

D-amino  acid  oxidase,  344,  346 

glutamate  dehydrogenase,  331 
3-Indolecarboxylate,  glutamate  dehydro- 
genase, 331 
/9-3-Indoleethylamine, 

structure  of,  324 

tryptophanase,  323 
3-Indolepropionamide,  chymotrypsin,  371 
3-Indolepropionate, 

carboxypeptidase,  367 

chymotrypsin,  369-371,  373 

histamine  release,  374 

tryptophanase,  323 

tryptophan  pyrrolase,  325 
Indolylpyruvate   keto-enol   tautomerase, 

iodine,  686 
Indoxyl-sulfate,    acetylindoxyl    oxidase, 

591 
Infantile  acrodynia,  see  Acrodynia 
Infections,   bacterial, 


malonate,  221-224 
Inflammation, 

o-iodosobenzoate,  724 

mercurials,  950 
Influenza  virus, 

infectivity  of,  mercurials,  978,  980-981 

proliferation  of, 

2-deoxyglucose,  400 
malonate,  126-127,  193 
Inhibition  index,  253-254 
Inhibitor  constants,  definition  of,  252 
Inosine, 

5 '-nucleotidase,  472 

pyridoxal  kinase,  475 
Inosinediphosphate  (IDP), 

ATPase,  445 

isocitrate  dehydrogenase,  509 
Inosine  hydrolase,  analogs,  471 
Inosinemonophosphate  (IMP) 

adenylosuccinate  lyase,  466 

D-amino  acid  oxidase,  545 

fructose- 1 ,6-diphosphatase,   470 

GMP  reductase,  471 

level  in  ascites   cells,   2-deoxyglucose, 

395 

5'-nucleotidase,  471 

pyridoxal  kinase,  475 

thiamine  kinase,  475 
Inosinemonophosphate   dehydrogenase, 

analogs,  471,  481 
Inosine  phosphorylase,  analogs,  471 
Inosinetriphosphatase   (ITPase), 

ADP,  446 

IDP,  446 

mercurials,  relation  to  SH  groups,  804 

quinacrine,  551 
Inosinetriphosphate  (ITP) 

ATPase,  445 

isocitrate  dehydrogenase,  509 

NADH  oxidase,  511 

pyridoxal  kinase,  475,  477 
Inositol  dehydrogenase,  o-iodosobenzoate, 

709 
Insulin,  2-deoxyglucose  uptake,  387 
Interference  inhibition,  42 
Intestine, 

acetylcholine  response,  hydrogen  pero- 
xide, 696 


1178 


SUBJECT    INDEX 


amino  acid  transport, 

competition    between    amino    acids, 

264-265 

deoxypyridoxol,  574 
biotin  transport,  analogs,  267 
carbohydrate    transport,     competition 
between  sugars,  264 
Ca++  transport, 

malonate,  208 

mercurials,  909,  913-914 
CI"  transport,  mercurials,  910 
contracture,  o-iodosobenzoate,  724 
2-deoxyglucose  transport,  387 
electrical  potential  across,  mercurials, 
916-917 
galactose  transport, 

analogs,  263 

6-deoxyglucose,  403 

mannose,  263 
glucose  transport, 

6-deoxyglucose,  264 

deoxypyridoxol,  574 

mercurials,  916 
glycine  transport,  analogs,  265 
I~  transport,  mercurials,  910 
mercurial  levels  in,  930,  959 
methionine  transport,  deoxypyridoxol, 
574-575 
motility  of, 

dehydroacetate,  624-625 

hydrogen  peroxide,  696 

o-iodosobenzoate,  724 

malonate,  212 

mercurials,  948-949 
Na+  transport,  mercurials,  916,  985 
short-circuit    current,    2-deoxyglucose, 
387 

triiodothyroacetate   transport,   mercu- 
rials, 911,  913-914 
uracil  transport,  mercurials,  911 
water  transport,  mercurials,  916,  985 
Invertase,  see  /S-Fructofuranosidase 
lodate, 

oxidation  of  nitrite,  450 
oxidation  of  SH  groups  by,  656 
Iodide, 

active  transports  of,  malonate,  209 
intestinal  transport  of,  mercurials,  910 


^-ketoadipate  chlorinase,  453 

tyrosinase,  300 

uptake  by  ciliary  body, 
malonate,  209 
nitrite,  267 

uptake  by  Fucus,  mercurials,  910,  912 

uptake  by  thyroid, 
malonate,  209 
mercurials,  910 
Iodine, 

active  transports,  690 

bacterial  growth,  690 

chemical   properties,   679-680 

enzyme  inhibitions,  682-688 

glycolysis,  689 

mitochondrial  swelling,   689 

NADH  oxidation  by,  689 

oxidation  of  protein  SH  groups,  680- 

681 

oxidative  phosphorylation,  688-689 

reaction  with  SH  groups,  680 

renal  SH  groups,  689 

skin  electrical  potential,  690 

solubility,   679 

sulfenyl  iodide  formation,  681 
lodoacetamide, 

group  attached  to  enzymes,  649 

urease,  643 
lodoacetate, 

group  attached  to  enzymes,  649 

heart,  antagonism  of  malonate  effect, 

216 

kidney,  transport  of  PAH,  205 

toxicity,  potentiation  by  hydrogen  per- 
oxide, 696 
7H-Iodobenzoate,    glutamate    dehydroge- 
nase, 330 
o-Iodobenzoate, 

bacterial  growth,  727 

formation  from  o-iodosobenzoate,  702 

toxicity,  725 
lodobenzoates,  D-amino  acid  oxidase,  341 
w;-Iodosobenzoate,    702-703 
o-Iodosobenzoate, 

ascorbate  oxidation  by,  703 

bacterial  growth,  727-728 

catecholamine   release,    725 

central  nervous  system,  724 


SUBJECT    INDEX 


1179 


chemical  properties,   701-703 

enzyme  inhibitions,  704-721 
kinetics,  715-716 
pH  effects,  716-717 
protection,  717-718 
reversal,  718 
variation  with  substrate,  718-721 

glycolysis,  721 

heart,  723-725 

histamine  release,  725 

inflammatory  activity,  724 

intestine,  724 

K^  in  mitochondria,  722 

lethal  doses,  725 

leucocytosis,   725 

muscle,  723 

NADH  oxidation  by,  703 

nerve,  724 

neuroblastic  damage,  724 

oxidation  of  SH  groups,  657,  702-703 

preparation  of,  702 

purification  of,  702 

reaction  with  protein  SH  groups,  703- 

704 

sea  urchin  egg  development,  726-727 

solubility,  716 

titration  of  enzyme  SH  groups,   714- 

715 

toxicity,  724-726 

urease,  643 

uterus,  724 

viruses,  728 
p-Iodosobenzoate, 

ascorbate  oxidation  by,  703 

glutamate  dehydrogenase,  702 
3'-IodothjTonine,    thjToxine    deiodinase, 

603 
lodotyrosine,  see  L-MonoiodotjTosine 
o-Iodoxybenzoate, 

bacterial  growth,  727 

chemical  properties,  702 
Ion  antagonisms,  452-453 
Ionization  constants, 

dehydroacetate,  619 

dicarboxylic  acids,  8 

hydrogen   paroxide,   690 

hypoiodous  acid,  679 

o-iodobenzoic  acid,  702 


o-iodosobenzoic  acid,   702 

mercuric  complexes,  736 

SH  groups,  637-638 
Ipomea  batatas  (sweet  potato),  oxidative 

phosphorylation, 

malonate,  120 
Iproniazid,  structure  of,  488 
Iron,  see  also  Ferric  ions 

catalysis  of  SH  oxidation,  658 

complexes  with  di-  and  tricarboxylates, 

12 

incorporation  into  protoporphjTin,  mer- 
curials, 888 

intestinal  transport  of,  malonate,  208 
Iron  oxidase, 

mercurials,  843 

quinacrine,  551 
Isatin,  acetylindoxyl  oxidase,  591 
Isoamylase,   mercurials,  843 
Isoamylase    (debranching),    glucono-1,4- 

lactone,  429 
IsoamylbutjTate,  lipase,  595 
Isobologram,   malonate  and  ouabain  on 

heart,  216 
Isobutylamine,  leucine  decarboxylase,  352 
Isobutyrate,     glutamate     decarboxylase, 

328 
Isocaproamide,    leucine   aminopeptidase, 

368 
Isocaproate, 

glutamate  decarboxylase.  328 

leucine  aminopeptidase,  368 
Isocitrate, 

oxalosuccinate  decarboxylase,  597 

oxidation  of,  malonate,  79,  86 

phosphofructokinase,   385-386 
Isocitrate  dehydrogenase, 

ferrocyanide,  677-678 

y-hydroxy-a-ketoglutarate,  615-616 

o-iodosobenzoate,  709 

mercurials,   843-844 

protection  by  isocitrate,  782 
protection  by  Mn++,  782 

nucleotides,  509,  513 

propane-tricarboxylate,  240 
Isocitrate  lyase, 

o-iodosobenzoate,  709 

malonate,  61 


1180 


SUBJECT    INDEX 


Isohypophosphate,  oxidative  phosphory- 
lation, 448 
D-Isoleucine,     L-alanine    dehydrogenase, 

354 
L-Isoleucine, 

arginase,  337 

dipeptidase,  367 

homoserine  kinase,  357 
Isomaltose, 

a-dextran-l,6-glucosida8e,  417 

a-glucosidase,  416,  423 
Isomeric  analogs,  268-274 

anomeric,   271-272 

definition  of,  257 

enantiomorphic,  268-271 

geometric,  272-274 

positional,  272 
Isomorphic   groups,   257 
Isoniazid, 

cadaverine  oxidation,  363 

glucose  dehydrogenase,  501 

NAD  nucleosidase,  488,  490-491,  493 

structure  of,   488 
Isoniazid-NAD,  NAD  nucleosidase,  489 
Isonicotinamide,  NAD  nucleosidase,  488, 

493 
Isonicotinate, 

glucose  dehydrogenase,  501 

NAD  nucleosidase,  491 
Isonicotinyl  hydrazide,  see  Isoniazid 
Isopenteny  1  -  pyrophosphate    isomerase , 

mercurials,  887 
Isophthalate, 

D-amino  acid  oxidase,  341,  344 

aspartate :  a  -  ketoglutarate      transami- 
nase, 334 

glutamate  dehydrogenase,  330-331 

intercharge  distance,  6 

ionization  constants,  8 

kynurenine:a-ketoglutarate     transami- 
nase, 607-607 
Isoporphobilinogen,  porphobilinogen  de- 
aminase, 600 
Isoriboflavin, 

L-amino  acid  oxidase,  540-542 

FAD   pyrophosphorylase,   542 

flavokinase,  539 

Lactobacillus  growth,  537 


riboflavin  deficiency,  538 

structure  of,  536 
Isostere,  definition  of,  246 
Isosteric  groups,  257 
Isovalerate, 

D-amino  acid  oxidase,  343 

glutamate  decarboxylase,  328 

leucine  decarboxylase,  352 
Isoxanthopterin , 

structure  of,  287 

xanthine  oxidase,  289 
Isoxanthopterin-6-carboxylate,    xanthine 

oxidase,  289 
Itaconate, 

biosjTithesis  of,  ferrocyanide,  678 

fumarase,  279 

glutamate  decarboxylase,  328 

ionization  constants,  8 

metabolism  of,  ferrocyanide,  678 

structure  of,  279 

succinate  dehydrogenase,  38,  601 
ITP,  see  Inosinetriphosphate 
ITPase,  see  Inosinetriphosphatase 

J 

Jensen  sarcoma, 

acetoacetate   accumulation,   malonate, 

138 

malonate  levels  in  vivo,  102 

respiration  (endogenous),  malonate,  179 

succinate  levels  in  vivo,  malonate,  102 
Jerusalem  artichoke,  see  Helianthus  tube- 

rosus 

K 

/3-Ketoadipate, 

aminolevulinate  dehydrase,  591 
kynurenine:a-ketoglutarate    transami- 
nase, 595 

pyridoxamine :  oxalacetate      transami- 
nase, 600 

/3-Ketoadipate  chlorinase,  halide  ions,  453 

2  -  Keto  -  3  -  deoxy  -  D  -  arabo  -  heptonate  -  7  - 
phosphate  synthetase,  analogs,  413 

a-Ketoglutarate, 

accumulation  of,  malonate,  110-111 
glutamate  dehydrogenase,  330 
D-a-hydroxy  acid  dehydrogenase,  437 


SUBJECT    INDEX 


1181 


lactate  dehydrogenase,  437 
oxidation  of, 
benzoate,  348 
malonate,  72,  79-81,  83-86 
mercurials,  879 
propane-tricarboxylate,  240 
«?eso-tartrate,  432 
thiamine  analogs  in  vivo,  520-521 
phosphofructokinase,  385 
pyridoxamine :  oxalacetate      transami- 
nase, 600 
pyruvate  decarboxylase,  431 

a-Ketoglutarate   oxidase, 

6-aminocotinamide  in  vivo,   505 
y-hydroxy-a-ketoglutarate,  616 
malonate,  61-62,  83-86 
mercurials,  844 

inhibition  in  vivo,  926 
oxygen  inactivation  of,  659 

a-Ketoglutarate,    reductase,    mercurials, 
844 

^-Ketoglutarate,  kynureninera-ketogluta- 
rate  transaminase,  608 

2-Keto-L-gulonate,   /3-glucuronidase,   424 

a-Ketoisocaproamide,  leucine  aminopepti- 
dase,  368 

a-Ketoisocaproate,  glutamate  decarboxy- 
lase, 329 

a-Ketoisocaproate  decarboxylase, 
benzoate,  349 
mercurials,  844 

a-Ketoisovalerate,  glutamate  decarboxy- 
lase, 329 

Ketomalonate, 

pyridoxamine :  oxalacetate      transami- 
nase, 600 
pyruvate  decarboxylase,  430-431 

zli-3-Ketosteroid  reductase,  estrone,  449 

zl*-3-Ketosteroid  reductase,  analogs,  449- 
450 

Kidney, 

acetate  oxidation,  malonate,  78,  232 

acetate  utilization,  propionate,  613 

acetoacetate     metabolism,     malonate, 

144 

ADP  levels  in,  2-deoxyglucose,  395 

D-alanine  oxidation,  benzoate,  341 

amino  acid  deamination,  benzoate,  348 


amino  acid  decarboxylase  in  vivo,  a- 
methyl-wi-tyrosine,  317 
amino  acid  transport,  competition  be- 
tween amino  acids,  264 
2J-amonohippurate  transport, 

azide,  205 

cyanide,  205 

dehydroacetate,  205 

2,4-dinitrophenol,  205 

fluoride,  205      ' 

fluoroacetate,  205 

iodoacetate,  205 

malonate,  205 

mercurials,  205 

phlorizin,  205 
ammonia  formation, 

benzamide,  348 

benzoate,  348 

phenylacetate,  348 

/3-phenylpropionate,  348 
ATP  levels  in,  2-deoxyglucose,  395 
Ca++  uptake  in  mitochondria,  mercu- 
rials, 909 

C-l/C-6  ratio,  malonate,  130 
cholesterol  biosjTithesis,  malonate,  150 
citrate  accumulation  in  vivo,  malonate, 
104-105 
citrate  levels, 

mercurials,  927 

sequential    inhibition    by    malonate 

and  fluoroacetate,  112 
coenzyme  A  level  in  vivo,  mercurials, 
927 

dehydroacetate,  626 
fatty   acid   oxidation,    malonate,    114, 
137,  141-143 

fluoromalonate  decarboxylation,  239 
galactose  transport,  glucose,  262 
glucose  oxidation, 

6-deoxy-6-fluoroglucose,  393,  404 

2-deoxyglucose,  393,  403 
glucose-6-phosphatase  in  vivo,  mercu- 
rials, 927 

glucose  uptake,  malonate,  127 
glutamate  metabolism,  malonate,  152 
jff-glycerophosphatase  in  vivo,   mercu- 
rials, 926-927 
glycolysis,  2-deoxygIuco8e,  392-393 


1182 


SUBJECT    INDEX 


glycolysis  (aerobic),  ferricyanide,  677 
GSH  levels,  tetrathionate,  700 
histological  changes, 

malonate,  219-220 

tetrathionate,  700 
a-ketoglutarate  oxidase  in  vivo,  mer- 
curials, 926 

a-ketoglutarate    oxidation,    malonate, 
81,  85 

K+  transport,  malonate,  205-206 
malonate, 

decarboxylation  of,  232 

levels  in  vivo,  100,  102 

oxidation  of,  233 

toxicity  of,  2,  219-220 
mercurials,  see  also  Mercurials,  kidney 

levels  in  vivo,  959-961 

resistance  to,  985 
mitochondria    of,    cycle    intermediates 
concentrations  in,  88-89 
NAD(P)  diaphorase  in  vivo,  mercurials, 
926 

Na+  transport,  malonate,  205-206 
oxalacetate  oxidation,  malonate,  82 
oxidative  phosphorylation,  mercurials, 
873,  927 

phosphatases  in  vivo,  mercurials,  926- 
927 

protein    disulfide    reductase    in    vivo, 
mercurials,  926 

pyrithiamine  levels  in  vivo,  528 
pyruvate  oxidation, 

malonate,  75 

meso-tartrate,  432 
respiration, 

<ra»5-aconitate,  273 

kojic  acid,  350 

malonate,  219 
respiration  (acetate), 

dehydroacetate,  626 

malonate,  206 
respiration  (endogenous), 

benzoate,  348 

dehydroacetate,   623-624 

malonate,  175-179,  181,  183 

mercurials,  833 
respiration  (glucose),   mercurials,  883- 
884,  927-928 


respiratory  quotient,  malonate,   185 
SH  groups  in,  iodine,  689 
sorbitol  dehydrogenase  in  vivo,  mercu- 
rials, 926 

succinate  dehydrogenase, 
dicarboxylate  ions,  35-38 
in  vivo,  mercurials,  925-926 
malonate,  29,  31-32 
succinate  levels  in  vivo,  malonate,  100, 
102, 

succinate  oxidation,  malonate,  56 
thiamine-PP  levels,  pyrithiamine,  525 
thiols  in,  mercurials,  931-923 
transaminases  in  vivo,  deoxypyridoxol, 
570 

tubular    transports,    see    specific    sub- 
stances 

Kidney  bean  leaves,  malonate  occurence 
in,  224,  226 

Klebsiella  pneumoniae,  infection  by, 
malonate,  221 

Kojic  acid, 

D-amino  acid  oxidase,  342,  349-350 
bacterial   growth,   349 
central   nervous   system,   349 
choline  oxidase,  350 
Leptospira   growth,    349 
lethal  doses,  349-350 
occurrence  of,  349 
phagocytosis,  350 
L-proline  oxidase,  350 
structure  of,  345,  617 
succinate  oxidase,  350 
tyramine  oxidation,  350 
urate  oxidation,  350 

Kynureninase, 
analogs,  595 
nicotinylalanine,  610 

Kynurenine,     kynurenine     formamidase, 
595 

Kynurenine  formamidase, 
analogs,  595 
mercurials,  860 

Kynurenine     hydroxylase,     nicotinylala- 
nine, 610 

Kynurenine :  a  -  ketoglutarate      transami- 
nase, see  Transaminases 


SUBJECT    INDEX 


1183 


Lactate, 

blood  levels  of, 
malonate,  219 
thiamine  analogs,  520 

enolase,  409 

formation  of,  see  Glycolysis 

metabolism  of,  analogs  432-438 

oxidation  of, 

6-aminonicotinamide,   504 
malonate,  78 

tyrosinase,  300 
D-Lactate,  phosphatase  (acid),  441 
L-Lactate, 

D-lactate   dehydrogenase,  437 

phosphatase  (acid),  441-443 
D-Lactate:  cytochrome  c  oxidoreductase, 

analogs,  437 

malonate,  62 

oxalate,  435 
L-Lactate:  cytochrome   c   oxidoreductase, 

analogs,  437 
D-Lactate  dehydrogenase,  analogs,  437 
L-Lactate  dehydrogenase, 

acetate,  436 

active  center  of,  434 

6-aminonicotinamide  in  vivo,  505 
„/ianalogs,  432-437 

benzoate,  501 

hydrogen  peroxide,  693 

iodine,  682-683,  686 

o-iodosobenzoate,  710,  717-718 

potentiation    of   inhibition   by   sub- 
strate, 717-718 

isoenzymes  of,  oxalate,  436 

malonate,   62,   65 

mercurials,  768,  772-773,  786-787,  802, 

804,  808-812,  814,  825,  844-845 
coenzyme  displacement,  786-787 
complex  with  mercuric  ion,  768 
protection  by  NADH,  782 
rate  constant  for  inhibition,  811 
relation  to  SH  groups,  802,  804,  808- 
809 

reversal  with  cysteine,  825 
spontaneous  reversal,  814 
type  of  inhibition,  772-773 

nicotinamide   analogs,   500-502 


nicotinylhydrazide-NAD,  497 

oxalate,  435-436 

oxamate,  433 

porphyrexide,  668 

quinacrine,  547,  551 
pH  effects,   557-558 

tartronate,  237-238,  436 

thionicotinamide-NAD,  497 
D-Lactate  oxidase, 

quinacrine,  552 

riboflavin,  542 
L-Lactate  oxidase, 

FAD,  543 

FMN,  543 

o-iodosobenzoate,  710 

malonate,  62 

quinacrine,  552 

riboflavin,  542-543 
Lactate  oxidative  decarboxylase,  quina- 
crine, 552 
Lactobacillus,  hydrogen  peroxide  forma- 
tion in,  695 
Lactobacillus  acidophilus,  growth  of, 

dehydroacetate,  632 
Lactobacillus  arabinosus, 

glutamate  utilization,  analogs,  327 

growth  of, 

ion  analogs,  452 
5-phosphoribonate,  411 
Lactobacillus  brevis, 

growth   of,   dehydroacetate,   632 

pentose-P  utilization,  mercurials,  885- 

886 
Lactobacillus  casei, 

folate  metabolism,  analogs,  582 

growth  of, 

3-acetylpyridine,  494 
dehydroacetate,  632 
mercurials,  972 
quinacrine,  546 
riboflavin  analogs,  537-538 

mevalonate  incorporation,   mercurials, 

886 
Lactobacillus  fermenti,  growth  of, 

dehydroacetate,    632 

thiamine  analogs,  530 
Lactobacillus  helveticus,  growth  of, 

deoxypyridoxol,  575 


1184 


SUBJECT    INDEX 


co-methylpyridoxol,  575 
Lactobacillus  lactis,  cyanocobalamin  ana- 
logs, 589-590 
Lactobacillus  plantarum,  growth  of, 

dehydroacetate,  632 
Lactonase-I,  mercurials,  845 
Lactose, 

/3-galactosidase,  418 

a-glucosidase,  416 
Lecithinase  A,  analogs,  595 
Leguminosae,    malonate    occurrence    in 

various  species,  225-226 
Leptospira  icterohaem orrhagiae , 

growth  of,  kojic  acid,  349 

respiration  (endogenous),  malonate,  168 
Lethal  doses, 

3-acetylpyridine,  499 

adipate,  201 

6-aminonicotinamide,  504 

dehydroacetate,  627-628 

6-deoxy-6-fluoroglucose,  404-405 

2-deoxyglucose,  401 

ethylmalonate,  201 

fluoromalonate,  239 

fluoromalonic  diethyl  ester,  239 

D-glucosone,  384 

glutarate,  201 

hydrogen  peroxide,  696 

o-iodosobenzoate,  725 

kojic  acid,  349-350 

malonamide,  201 

malonate,  201,  218 

mercurials,  955-957 

oxythiamine,  530 

pyrithiamine,  530 

tetrathionate,  700 
Lethal   synthesis,    analog   incorporation, 

247,  258 
DL-Leucinamide,   D-amino   acid   oxidase, 

340 
L-leucinamide,  dipeptidase,  367 
D-Leucine,    L-amino    acid   oxidase,    268- 

340 
L-Leucine, 

alanine:  a-ketoglutarate     transaminase, 

334 

D-amino  acid  oxidase,  340 

arginase,  337 


dipeptidase,  367-368 
leucine  aminopeptidase,  368 
phosphatase  (acid),  441 

Leucine  aminopeptidase,  analogs,  368 

Leucine  decarboxylase, 
analogs,  352 

mercurials,  783,  810,  814,  845 
protection  by  leucine,  783 
protection  by  pyridoxal-P,  783 
rate  of  inhibition,  810 
spontaneous  reversal,  814 

L-Leucinol,  leucine  aminopeptidase,  368 

Leucocytes, 

glycolysis  (aerobic),  oxamate,  434 
migration  of,  mercurials,  968 
phagocytosis,  malonate,  203 
pyruvate  oxidation,   malonate,   76 

Leucocytosis,  o-iodosobenzoate,  725 

Leuconostoc  mesenteroides,  growth  of, 
5-phosphoribonate,  411 

Leucophaea  maderae,  succinate  dehydro- 
genase, 
malonate,  29 

Leucopterin, 

structure  of,  287 
xanthine  oxidase,  289 

DL-Leucylglycine,  D-amino  acid  oxidase, 
340 

DL-Leucylglycylglycine,  D-amino  acid  oxi- 
dase, 340 

Leucyl-L-tyrosine,  carboxypeptidase,  367 

Leukemia, 

deoxypyridoxol  use,  576 
malonate  use,  202 

Leukemic  cells, 

folate  metabolism,  analogs,  582 
glycolysis,  2-deoxylucose,  392 

Lilium , 

microspore  meiosis,  mercurials,  966-967 
pollen  of,  malonate  metabolism  in,  228 

Limulus  polyphemus  gill  cartilage,  succi- 
nate dehydrogenase, 
malonate,  19,  29 

Lipase, 

analogs,  595-596 
dehydroacetate,  623 
ferricyanide,  676 
iodine,  686 


SUBJECT    INDEX 


1185 


o-iodosobenzoate,  710,  718-719 

variation    of   inhibition    with    subs- 
trates, 718-721 
malonate,  62 
mercurials,  719,  860 
porphyrexide,  668 
Lipids,  see  also  Fatty  acids.  Phospholipids, 
Sterols 

biosynthesis  of,  mercurials,  886-887 
metabolism  of,  see  also  Lipogenesis 
malonate,  135-151 
propionate,   613-614 
Lipoamide    dehydrogenase,    see    NADH: 

lipoamide  oxidoreductase 
Lipoate, 

analogs  of,  590 

acetyllipoate   biosynthesis,   590 
acyl  transfer,  590 
phosphotransacetylase,  590 
pyruvate  decarboxylation,  590 
pyruvate  oxidation,  590 
mercurials,  complexes  with,  750-751 
mercurial  toxicity,  antidote  for,  751 
pyruvate  oxidase,  antagonism  of  mer- 
curial inhibition,  751 
Lipogenesis,  2-deoxyglucose,  399 
Lipoprotein  lipase,  macroions,  463 
Lithium,  phosphotransacetylase,  452 
Liver, 

acetate  oxidation,  malonate,  78 
acetate  utihzation,  propionate,  613 
acetoacetate   accumulation,   malonate, 
138-144 

acetoacetate     metabolism,     malonate, 
115,  139-140,  144 

ADP  levels  in,  2-deoxyglucose,  395 
D-alanine  oxidation,  benzoate,  341 
amino  acid  levels  in,  mercurials,  954 
amino  acid  oxidation,  kojic  acid,  350 
aminomalonate  decarboxylation  in,  239 
arginine  conversion  from  citrulline,  ma- 
lonate, 157-158 

aspartate  oxidation,  malonate,  152 
ATP  levels, 

aminopterin,  585 
2-deoxyglucose,  395 
malonate,  157-158 


Ca++  uptake  by  mitochondria,  mercu- 
rials, 909-910 

cholesterol  metabolism,  malonate,  150 
citrate  accumulation  in  vivo,  malonate, 
104 

citrate  oxidation,  trans-aconitate,  273 
Cu++  uptake,  mercurials,  910,  913 
cycle  intermediates  concentrations  in, 
89 

damage  by  mercurials,  954 
enzyme  induction  in,  8-azaguanine,  478 
fatty  acid  biosynthesis, 

malonate,  148-149 

mercurials,  887 

propionate,  613 
fatty  acid  dehydrogenation  (anaerobic), 

malonate,  136-137 
fatty  acid  oxidation, 

malonate,  114,  137,  141-143 
folate  metabolism,  analogs,  582 
glucose  metabolism,   6-deoxy-6-fluoro- 
glucose,  404 

glucuronide  biosynthesis,  analogs,  428 
glutamate      accumulation,     malonate, 
153-154 
glutamate  oxidation, 

galactoflavin,  544 

malonate,  152 

mercurials,  878 
glycogen  formation,  D-glucosamine,  382 
glycogen  levels  in,  oxythiamine,  520 
glycolysis, 

2-deoxyglucose,  383 

malonate,  126 
glycolysis  (aerobic),  malonate,   128 
GSH  levels  in,  tetrathionate,  700 
/5-hydroxybutyrate  oxidation, 

galactoflavin,  544 

mercurials,  883 
isocitrate  oxidation,  malonate,  79,  86 
a-ketoglutarate  oxidation,  malonate,  80 
K+   uptake   by   mitochondria,   mercu- 
rials, 909,  914 

lipogenesis,  2-deoxyglucose,  399 
malate  oxidation,  malonate,  82 
malonate, 

decarboxylation  in,  233 

levels  in  vivo,  100,  102 


1186 


SUBJECT    INDEX 


metabolism  in,  228,  232-234 
mercurial  levels,  930,  959-960 
methylmalonate  formation  in,  226 
methylmalonyl-CoA  and  succinyl-CoA 
interconversion  in,  235 
Mg++   uptake   by   mitochondria,   mer- 
curials, 909 
NAD  levels  in, 

6-aminonicotinamide,  505 

aminopterin,  585 
Na+  transport  by  mitochondria,  mer- 
curials, 909 

oxalacetate  oxidation,  malonate,  82 
oxidative  phosphorylation, 

malonate,  119,  121 

mercurials,  873 
phenylalanine  incorporation  into  pro- 
teins,  mercurials,   887 
phospholipid    biosynthesis,    malonate, 
151 

propionate  metabolism,  malonate,  145 
protein  biosynthesis, 

p-fluorophenylalanine,  351 

malonate,   156, 

mercurials,  887 
pyridoxine   levels  in,   deoxypyridoxol, 
567 

pyrithiamine  levels  in  vivo,  528 
pyruvate  oxidation, 

malonate,  75 

mercurials,  878 

pyrithiamine,  520 

quinacrine,  560 
respiration, 

kojic  acid,  350 

quinacrine,  559-560 

tartronate,  237 
respiration  (endogenous), 

<rans-aconitate,   273 

6-aminonicotinamide,  504 

dehydroacetate,  623-634,  629 

malonate,  176,   183 

nicotinamide,  503 
respiratory  quotient,  malonate,  184 
squalene  biosynthesis,  mercurials,  886 
succinate  accumulation,  malonate,  91, 
145 


succinate  dehydrogenase, 
malonate,  29-31 
phthalate,  37 
succinate  levels  in  vim,  malonate,  100, 
102 

succinate  oxidase,  malonate,  22 
thiamine-PP  levels  in,   thiamine  ana- 
logs, 525-527 

transaminase  in  vivo,  deoxypyridoxol, 
570 
urea  formation, 

deoxypyridoxol,  573 
malonate,   116 
Loblolly  pine,  see  Pinus  taeda 
Locust,  see  also  Schistocera 
fat  body  of, 

fatty  acid  formation  from  malonate, 
234 

malonate    inhibition    of   fatty    acid 
biosynthesis,   147 
malonate  oxidation  in,  233 
sarcosomes,  a-ketoglutarate  accumula- 
tion by  malonate.  111 
Locusta    migratoria    muscle,    respiration 
(endogenous), 
malonate,  175 
Lombricine  kinase, 
o-iodosobenzoate,  710 
mercurials,  845 
Lucerne  (green  alfalfa),  malonate  occur- 
rence in,  225 
Luciferase, 

ferricyanide,  676 
mercurials,  845 
Lumichrome,  flavokinase,  539 
Lumiflavin, 

flavokinase,  539 

riboflavin   transglucosidase,   543 
structure  of,  536 
Luminescence,  see  Bioluminescence 
Lung, 

malonate  decarboxylation  in,  232 

malonate  levels  in  vivo,  102 

mercurial  levels  in  vivo,  930,  959 

pentose-P  pathway,  malonate,  130 

respiration      (endogenous),     malonate, 

176-177 

succinate  levels  in  vivo,  malonate,  102 


SUBJECT    INDEX 


1187 


transaminases  in  vivo,  deoxyp>Tidoxol, 

570 
Lupine,  see  also  Lupinus 
Lupine    mitochondria, 

a  -  ketoglutarate    oxidation,    malonate, 

80 

oxidative   phosphorylation,    malonate, 

119 

succinate  oxidation,  malonate,  23 
Lupinus  albus, 

oxidative   phosphorylation,    malonate, 

119-120 

succinate  dehydrogenase,  malonate,  33 
Lymph  node, 

glucose  uptake,  analogs,  263 

glycolysis,  2-deoxyglucose,  392 

protein  biosynthesis,  malonate,  156 
Lymphoma,   protein   biosjmthesis,    ame- 

thopterin,  585 
Lymphosarcoma,  growth  of, 

6-aminonicotinamide,  505 

deoxypyridoxol,   576 

riboflavin  analogs,  538 
Lysine, 

arginase,  335,  337-338 

histidase,  353 
D-Lysine,  D-amino  acid  oxidase,  340 
L-Lysine, 

aspartokinase,  356 

blood  urea,  338 

homoserine  kinase,  357 

intestinal  transport  of,  analogs,  265 

leucine  decarboxylase,  352 

urea  formation  in  vivo,  338 
L-Lysine  decarboxylase,  mercurials,  845 
Lysolecithin  oxidase,  FAD,  543 
Lysozyme, 

macroions,  459 

mercuric  ion  complex,  770 
Lytechinus  eggs,  cleavage, 

malonate,   198 
Lyxoflavin, 

phosphorylation  of,  539 

structure  of,  536 

utilization  by  Lactobacillus,  539 
Lyxoflavin    dinucleotide,    D-amino    acid 

oxidase,  544 
Lyxose,  /S-galactosidase,  418 


M 

Macroions, 

amylases,  464 

chymotrypsin,  457 

;8-fructofuranosidase,  465 

hyaluronidase,  459-461 

inhibitions  by, 

factors  determining,  455-456 
kinetics  of,  454-456 
pH  effects,  454-457,  461,  464 
specificity  of,  454 

interaction  with  enzymes,  454-455 

lipoprotein  lipase,  463 

lysozyme,  459 

pepsin,  457-458 

phosphatase  (acid),  464 

polynucleotide  phosphorylase,  463 

ribonuclease,  461-463 

trypsin,  456-457 
Macrophages,  respiration  (glucose), 

oxamate,  435 
Magnesium, 

mitochondrial  uptake,  mercurials,  909 

urinary  excretion,  mercurials,  921 
Malate, 

fumarase,  275-277 

glutamate  dehydrogenase,  332 

ionization  constants,  8 

lactate  dehydrogenase,  437 

oxalacetate  decarboxylase,  597 

oxidation  of, 
malonate,  81-82 
mercurials,  879 

phosphofructokinase,  385 

renal  excretion  of,  malonate,  207 

succinate  dehydrogenase,  36 
D- Malate, 

phosphatase  (acid),  441 

D-tartrate  dehydrase,  601 

weso-tartrate  dehydrase,  601 
L-Malate, 

D-a-hydroxy  acid  dehydrogenase,  437 

phosphatase   (acid),   441 

D-tartrate  dehydrase,  601 

tartronate  semialdehyde  reductase,  602 
Malate  decarboxylase,  malonate,  62 
Malate  dehydrogenase, 

analogs,  596 


1188 


SUBJECT    INDEX 


)S-fluoromalate,  279 

a-hydroxysulfonate,  438 

iodine,  686 

D-malate,  268 

malonate,  62 

mercurials,  845-846 

relation  to  SH  groups,  802-804,  807 

808 

type  of  inhibition,  772 

nicotinamide,  503 

nucleotides,  509,  513 

porphyrexide,  668 

quinacrine,  552 

tartronate,  237 
Malate  dehydrogenase  (decarboxylating) 

(malic  enzyme), 

analogs,  596-597 

o-iodosobenzoate,  704-710 

malonate,  63,  65 

mercurials,  846 

protection  by  Mn++,  779,  783 
protection  by  substrates,  779,  783 

tartronate,  238 
L-Malate  hydro-lyase,  see  Fumarase 
Malate  synthetase,  see  Glyoxylate  trans- 
ace  tase 
Malate:vitamin  K  oxidoreductase, 

mercurials,  846 

quinacrine,  552 
Maleate, 

alanine;  a-ketoglutarate     transaminase, 

334 

D-amino  acid  oxidase,  343 

aspartate :  a  -  ketoglutarate      transami- 
nase, 334 

chelation  with  cations,  12 

fumarase,  275-277 

glutamate  decarboxylase,  328 

intercharge  distance,  6 

ionic  length,  188 

ionic  volume,  188 

ionization  constants,  8 

permeability  of  erythrocytes  to,  188 

phosphatase  (acid),  442 

succinate  dehydrogenase,  34-36 

tyrosinase,  301 

urinary  citrate,  109 

urinary  a-ketoglutarate,   110 


Maleate  dehydrogenase,  malonate,  63 

Malignant  carcinoid  syndrome,  deoxypy- 
ridoxol,  574 

Malonamide, 
lethal  dose,  201 
tumoristasis,  201 

Malonate, 

acetate  oxidation,  77-78,  232 
acetoacetate  metabolism,  138-144,  149, 
234 

acetylaspartate  formation,   154 
acetylcholine  synthesis,  165-166 
cis-aconitate  oxidation,  79 
adaptive  enzyme  synthesis,  155 
amino  acid  accumulation  by,  103 
amino  acid  metabolism,  151-154 
y-aminobutyrate  oxidation,   154 
aspartate :  a  -  ketoglutarate      transami- 
nase, 334 

ATP  level  in  liver  homogenate,  157-158 
bacterial  growth,  195 
bacterial  infections,  221-224 
biosynthesis  of,  2,  226-227 
blood  coagulation  219 
blood  glucose,  164,  219 
blood  lactate,  219 
blood  pressure,  213 
blood  pyruvate,   219 
bond  characteristics,  4 
Br~  uptake  by  barley  roots,  116-117 
carcinostasis,  200-202 
cell  division,  117,  197-200 
cellulose  biosynthesis,  132 
chelation  of  cations,  11-14,  66-69 

cardiac  effects  of,  214-216 
chemotaxis  of  leucocytes,  203 
cholesterol  metabolism,  149-150 
chondroitin  sulfate  synthesis,  166 
ciliary  activity,  203 
citrate  accumulation  by,  104-110 
citrate  levels  in  tissues,  223 
citrate  oxidation,  78-79,  86-87 
coliphage,  194 
coronary  flow,  213 
decarboxylation  (nonenzymic)  of,  3 
decarboxylation  in  tissues,  232-234 
determination  of,  13 
diuretic  action,  221 


SUBJECT    INDEX 


1189 


electrocardiogram,  215 
embryogenesis,    197-199 
fatty  acid  oxidation,  114,  135-137 
fatty  acid  biosynthesis,   146-149 
flagellar  activity,  203 
fumarase,  275 

fumarate  antagonism  of  inhibitions  by, 
112-117 

fumarate  oxidation,  81 
fungal  growth,  195-196 
gastric  acid  secretion,  187,  208 
glucose  metabolism,  122-135 
glutamate  decarboxylase,  328 
glycerol  metabolism,  164 
glycolysis,   125-130 
glyoxylate  metabolism,  165 
heart,  213-217 
histamine  release,  212-213 
hydrogen  bonding  in,  4,  9 
hydroxamic  acid  formation,  233 
D-a-hydroxy  acid  dehydrogenase,  437 
incorporation  into  lipids,  231 
intercharge  distance,  5-6 
intestinal  transport  of,  207-208 
ionic  length  and  volume,   188 
ionic  species,  pH  effects,  9-10 
ionization  of,  5,  8-10 
iron  incorporation  into  heme,  163 
isocitrate  accumulation,  106 
isocitrate  oxidation,  79,  86 
ketogenic  activity,  138-144 
a-ketoglute^rate   oxidation,    72,    79-81, 
83-86 

kynurenine:  a-ketoglutarate     transami- 
nase, 608 

lactate  formation,  126-128 
lactate  oxidation,  78 
lethal  doses,  201,  218 
lipid  metabolism,  135-151 
malate  oxidation,  81-82 
metabolism  of,  1-2,  105,  107,  138,  216, 
224-235 

methylmalonate  accumulation  by,  145 
mitochondrial  swelling,  210-211 
mitosis,   197-200 
muscle,  212 
nephrotoxicity,  219-221 


nerve  conduction  and  potentials,  211- 

212 

occurrence  of,  2,  224-226 

oxalacetate  oxidation,  66-68,  82 

oxidative  phosphorylation,  118-122 

partition  ratios,  189 

pentose-P  pathway,  130-132 

permeability  to,  51,  56,  167,  186-192 

phagocytosis,  203,  223 

pH  effects  on  inhibitions  by,  181-182, 

189-192,  195-197 

phosphatase  (acid),  442 

phospholipid  metabolism,   151 

photosynthesis,  163-164 

plant  growth,   196-197 

plasma  K+  and  Na+,  206 

plasma  pH,  206 

porphyrin  biosynthesis,  158-163 

propionate  metabolism,   144-146 

protein  biosynthesis,  155-157 

purification  of,  3 

pyruvate  oxidation,  73-77 

reduction  of  Ca++  and  Mg++  concen- 
trations, 13-14 

renal  excretion  of  malate,  207 

renal  transports,   203-207 

respiration  (endogenous),  166-186 
effect  of  functional  activity,  183-184 
effect  of  ions  and  buffer,  183 
effect  of  tissue  age,  182-183 
factors  determining  inhibition,  166- 
167,  180 

kinetics,  180-182 
significance  of  inhibition,  186 

respiration  (glucose),  123-125,  127,  133- 

135 

respiratory  quotient,  184-185 

sequential  inhibition  with  fluoroacetate, 

112 

specificity  of  inhibition,  58-59,  72,  117- 

118 

stability  of,  3 

structure  of,  4 

succinate  accumulation  by,  90-104 

succinate   decarboxylation   to    propio- 
nate, 165 

succinate  dehydrogenase,  15-50 
activation  of,  45-46 


1190 


SUBJECT    INDEX 


binding   energy,   42 

competitive  nature  of  inhibition,  21- 

25 

effect  of  ATP,  48 

effect  of  Ca++,  46-48 

effect  of  electron  acceptor,  18-20 

effect  of  Mg++,  48 

effect  of  osmolarity,  46-47 

effect  of  temperature,  46 

evidence  for  site  of  inhibition,  20-21 

inhibition  constants,  25,  33-34 

nature  of  binding,  40-45 

variation  between  species  and  tissues, 

49-50 
succinate  oxidation  in  cellular  prepa- 
rations, 50-58 

tissue  levels  in  vivo,  100-102 
toxicity,  217-221 
tricarboxylate  cycle,  69-88 

effect  of  alternate  pathways,  71-72 

effect  of  oxalacetate  supply,  69-71 

effect  of  succinate  accumulation,  70- 

71 

variation  with  time,  70 
uptake  by  leaves,   107 
urea  formation,  157-158 
urinary  acetone,   138 
urinary  citrate,  99,   104-110 
urinary  excretion  of,  101 
urinary  flow,  206 

urinary  a-ketoglutarate,  99,  110-111 
urinary   pH,    206 
urinary  succinate,  98-99,  101 
virus  proliferation,  192-194 
yeast  growth,  195 
Malonate  decarboxylase,  229 
Malonate    semialdehyde,    occurrence    of, 

225 
Malondialdehyde, 
structure  of,  41 

succinate  dehydrogenase,  40-41 
Malondiamide, 
structure  of,  41 
succinate  dehydrogenase,  41 
Malonic  diethyl  ester, 

acetate  oxidation,  236-237 
glucose  oxidation,  236-237 
mycobacterial  infections,  224 


succinate  dehydrogenase,  236 
Malonic  esters,  ionization  of, 

constants,   1 1 

rates,  11 
Malonic  ethyl  esters. 

Bacillus  cereus  sporulation,  236 

carcinostasis,  236 

effects  on  metabolism,  236-237 

hydrolysis  of,  236 

ionization  constants,  236 

keto-enol  tautomerism,  4 

mycobacterial  infections,  236 

permeability  to,   11 

prolongation    of   hexobarbital    action, 

236 
Malonyl-CoA-COj  exchange  enzyme,  mer- 
curials, 847 
Malonyl  semialdehyde  pantetheine,  reac- 
tion with  jj-mercuribenzoate,  751 
Maltase,  see  also  Amylomaltase  and  Iso- 

amylase 

o-iodosobenzoate,  710 

mercurials,  860 

methylglucoside   anomers,   271 

synthesis  of, 

p-fluoro phenylalanine,  351 
tryptazan,  326 
Maltose, 

a-amylase,  420 

/3-amylase,  421 

a-dextran-l,6-glucosdase,  417 

)S-galactosidase,  418 

^-glucosidase,  417,  423 

a-mannosidase,  422 
Maltose  4-glucosyltransferase,   see  Amy- 
lomaltase 
Malus  malus, 

citrate   oxidation,   malonate,   79 

succinate  oxidation,  malonate,  53 
Malus  sylvestris,  see  also  Apple 

oxidative    phosphorylation,    malonate, 

120 
Mammary  carcinoma,  growth  of, 

deoxypyridoxol,  576-577 

malonate  and  derivatives,  202 
Mammary  gland, 

citrate  utilization,  ^rans-aconitate,  273 

fatty  acid  biosynthesis, 


SUBJECT    INDEX 


1191 


<ran5-aconitate,  273 
mercurials,  887 

fatty  acid  formation  from  acetate  or 

propionate,   malonate,    146-147 

glutamate  oxidation,  malonate,  152 

pyruvate  oxidation,  malonate,  75 

respiration  (acetate),  malonate,  146 

respiration  (endogenous),  malonate,  176 
Mandelate,  lactate  dehydrogenase,  437 
Mannarate,  /3-glucuronidase,  424 
Mannaro -1,4-3,6- dilactone,    fi - glucoroni- 

dase,  424 
Mannitol,  a-mannosidase,  422 
Mannoheptulose, 

fructokinase,  376 

glucokinase,  376 

glucose  uptake  into  tissues,  376 
Manono-l,4-lactone,   a-mannosidase,  429 
Mannose, 

a-amylase,  420 

fructokinase,  376 

galactose  transport,  263 

glucokinase,   376 

glucose   uptake   by   lymph   node,   263 

a-glucosidase,  416-417 

a-mannosidase,  422 

toxicity  to  bees,  414 

uptake  by  lymph  node,  394 
Mannose-6-phosphate, 

glucose-6-P  dehydrogenase,  411 

hexokinase,  379-380 

phosphopentose   isomerase,   411 
a-Mannosidase, 

analogs,  422,  429 

mercurials,  812 
Mannuronate, 

a-glucuronidase,  426 

y3-glucoronidase,  424 
Mannurone,  jS-glucuronidase,  424 
Maple  sugar  urine  disease,  see  Branched 

chain  ketonuria 
Marigold  stem  cultures,  growth  of, 

malonate,  197 
Marinogamm arus  m arinus , 

respiration    (endogenous),    mercurials, 

882 

toxicity  of  mercurials,  861 
Melanin,  formation  of. 


analogs,  303-305 
Melezitose,  a-glucosidase,  416 
Melibiose,   galactosidases,   418 
Membrane  potentials,  see  also  specific  tis- 
sues 
heart, 

acetylpyridines,  499-500 
dehydroacetate,  625 
malonate,  214-215 
mercurials,  896,  945-946 
intestine,  mercurials,  916-917 
kidney,  mercurials,  923,  936 
muscle,  mercurials,  937-938 
nerve, 

malonate,  211-212 
mercurials,  949-950 
oxythiamine,  531 
pyrithiamine,  531 
skin, 

iodine,  690 
mercurials,  950 
porphyrindin,  669-670 
Membrane  transport,  see  also  Active  trans- 
port. Permeability,  and  the  specific  or- 
ganisms and  tissues 

analog  inhibition  of,  261-268 
Menadione  reductase,  see  NADH:mena- 

dione  oxidoreductase 
Menthyl-a-glucuronide, 
a-glucuronidase,  426 
^-glucoronidase,  272 
Mepacrine,  see  Quinacrine 
Meralluride,  see  also   Mercurials,  kidney 

structure  of,  917 
Merbromin    (Mercurochrome),    structure 

of,  970 
Mercaptalbumin,  mercurials,  757-761 
Mercaptans,  see  Thiols 
Mercaptides, 

formation  of,  642 
of  mercurials  and  thiols,  746-751 
Mercaptoacetate,  lactate  dehydrogenase, 

437 
a-Mercaptobutyrate,   lactate   dehydroge- 
nase, 437 
Mercapto  groups,  see  Sulf hydryl  groups 
Mercaptomerin,   see   Mercurin  and   Mer- 
curials, kidney 


1192 


SUBJECT    INDEX 


a-Mercaptopropionate,   lactate   dehydro- 
genase, 437 

6-Mercaptopurine, 

adenosine  deaminase,  466 
inosinic  acid  metabolism,  281 
metabolic  products  of,  481 
purine   metabolism,   480 
xanthine  oxidase,  282 

6-Mercaptopurine  nucleotide, 
adenylosuccinate  lyase,  466,  481 
adenylosuccinate  synthetase,  467 
polynucleotide  phosphorylase,  481 

Mercaptosuccinate,     lactate'   dehydroge- 
nase, 437 

Mercuhydrin,  see  Meralluride 

Mercurialism,  see  Mercurials,  toxicity 

Mercurials, 

accumulation  in  kidneys  in  vivo,  923-924 
acetate  uptake  by  diaphragm,  912 
acrodynia,  953-954 
active  transports,  907-917,  936-937 
actomyosin,  938-940 
affinities  for  ligands,  732-741,  744 
ameboid  movement,  982 
amino  acid  complexes  with,  pH  effects, 
760-761 

amino  acid  levels  in  liver,  954 
auxin  transport,  967 
bacterial  growth,  970-976 
bacterial  uptake  of,  974-975 
bioluminescence,   888-891 
catecholamine  release  by,  947 
Ca++  uptake  by  mitochondria,  909-910 
cell  membranes,  892-907 
central  nervous  system,  951-952 
chemical  properties,  730-745 
citrate  accumulation  in  vivo,  927 
coenzyme  A  complexes  with,  750 
COa  fixation,  892 

colored,  histochemical  SH  determina- 
tions,  766-768 

comparison  of  inorganic  and  organic, 
743-745 

conditioned  reflexes,  985 
crystalhne  mercuri-enzymes,  768-770 
Cu++  uptake  by  liver,  910,  913 
determination  of  protein  SH   groups, 
762-768 


dissociation  into  Hg++,  930-935 
diuretics,  see  also  Mercurials,  kidney 

release  of  Hg++,  930-935 

structures  of,  917 

type  structure  of,  930 
electrocardiogram,  945 
electron  transport,  870-872 
embryogenesis,  963-965 
enzyme  inhibitions,  768-869 

coenzyme  displacement,  784-787 

comparison  of  mercurials,  860-861 

configurational  changes,  787-790 

effect  of  buffers,  797-798 

effect  of  ionic  strength,  797 

effect  of  pH,  790-797 

kinetics,  771-778,  908-815 

meaning  of  K„  861-862 

methods   of  expressing    inhibitions, 

861-862 

mutual    depletion    situations,    861- 

862 

protection,  778-784 

rates  of,  809-815 

reversals,    821-828 

type  of,  771-778 
ethanol  oxidation,  898 
fatty  acid  biosynthesis,  887 
fatty  acid  oxidation,  887 
functionality,  743 
ganglionic  transmission,  949 
gastric  acid  secretion,  914-916 
glucose  uptake,  893-894,  903-905,  910- 
911 

glycolysis,  874-877 
growth  stimulation,  967-968,  971 
GSH  level  in  erythrocytes,  905-906 
heart,  940-948 
hemoglobin,    755-757,    760 
hemolysis,  900-907 
histamine  release  by,   949 
histological  changes,  954-955 
intestine,  909-917,  948-949 

membrane  potentials,  916-917 

motility,  948-949 

transports,    909-911,    913-914,    916- 

917 
invertebrates,  961-963 
K+  efflux, 


SUBJECT    INDEX 


1193 


erythrocytes,  903-905 
yeast,  898-900 

kidney,  917-937 

coenzyme  A  levels,  927 

damage  to,  924-925 

enzyme  activities  in  vivo,  925-927 

glomerular  filtration,  918 

mechanisms  involved,  930-937 

membrane  potentials,  923,  936 

relation  of  structure  to  action,  930- 

935 

resistance  to  toxic  effects,  985 

scheme  of  reactions  in,  934 

SH  groups  in,  921-923 

sites  of  action,  920-925 

summary  of  actions,   918-920 

tubular  transports,  205,  918-921,  928, 

936 

K+  uptake,  908-909,  914 

lethal  doses,  955-957 

lipid  biosynthesis,  886-887 

lipid  solubility  of,  743-744 

lipoate  complexes  of,  750-751 

mercaptalbumin,  757-761 

Mg++  uptake  by  mitochondria,  909 

Mg++  urinary  excretion,  921 

mitosis,  963-970 

muscle,  937-940 

Na+  transport,   909 

nephrotoxicity,    924-925 

nerves,  949-950 

neuroblastic  damage,  964-965 

nuclear  damage,  965 

organic,   chemical   properties   of,   742- 

745 

ovabulmin,   753-755,    760 

oxidative  phosphorylation,  872-874 
in  vivo,  927 

parthogenesis  by,  963-964 

pentose-P  pathway,   885-886 

permeability  of  erythrocytes,  900-907 

phosphate  uptake,  910,  912-913 

phospholipid  biosynthesis,  887 

photophosphorylation,  892 

photosynthesis,  891-892 

plant  growth,  965-968 

porphyrin  biosynthesis,  888 

proteins, 


biosynthesis  of,  887-888 

configurational  changes,  761-762 

pH  effects,  760-761 

reactions  with,  751-768 
protozoa,  981-982 
reaction  with  disulfides,  750 
reaction  with  NAD,  774 
reaction  with  SH  groups,  764-751 

pH  effects,   749-750 
reaction  with  thiamine,  774 
resistance  to,  983-985 
respiration,   879-886 

stimulation  of,  879,  881-882,  884 
respiration  (glucose),  893-894,  898 
retina,  952-953 
skin,  950 

spindle  formation,  969 
sterol  biosynthesis,  886-887 
sucrose  uptake,  911 
thioester  splitting  by,  751 
tissue  distributions  of,  930,  955,  958- 
961 

titration  of  enzyme  SH  groups,   798- 
809 

titration  of  protein  SH  groups,  762-766 
toxicity,  940-941,  944,  950-957 
transfer  ENA  biosynthesis,  820 
tricarboxylate  cycle,  877-879 
tumor  uptake  of  in  vivo,  969-970 
uracil  transport,  911 
urinary  Ca++  excretion,  921 
urinary  excretion  of,  928-930,  952,  955, 
958,  960 
viruses,   976-981 
water  transport  by  cornea,  911 
xylose  uptake  by  diaphragm,  911-912 
p-Mercuribenzoate  (p-MB),  see  also  Mer- 
curials 

aldolase,  643,  649 
group  attached  to  enzymes,  649 
hemoglobin  SH  groups,  649 
lipase,  variation  of  inhibition  with  sub- 
strate, 719 

nucleic  acid  complexes  of,  744 
3  -  phosphoglyceraldehyde     dehydroge- 
nase, 650 

phosphorylase  a,  648-649 
preparation  of  solutions  of,  743 


1194 


SUBJECT    INDEX 


purification,  745 

reaction  with  NAD,  774 

release  of  Hg++,  744-745 

solubility,  743 

stability,  745 

structure  of,  742 

synthesis  of,  745 

titration  of  SH  groups,  763-766 

urease,  643 

Mercuric  chloride,  see  also  Mercurials  and 
Mercuric  ion 
solubility,  730-731 
solubility  product,  731 
structure  of,  730-731 

Mercuric  ion,  see  also  Mercurials  for  gen- 
eral topics 

amine  complexes  of,  738-740 
amino  acid  complexes  of,  737-741,  760- 
761 

bacterial  growth,  antagonism  by  thia- 
mine, 774 

bond  characteristics,  731,  744-745 
concentrations    in    solution,    732-736, 
740-741 

cysteine  complexes  of,  739,  748-749 
EDTA  complexes  of,  738 
electronegativity,  731 
GSH  complex  of,  748-749 
halide  complexes  of,  730-738 
hydration  of,  736 
hydroxyl  ion  complexes  of,  736 
methionine  complexes  of,  739 
nucleic  acid  complexes  of,  741 
nucleoside  complexes  of,  739 
nucleotide  complexes  of,  741 
oxidation-reduction  potentials,  731 
phosphoribosyl  -  ATP    pyrophosphoryl  - 
ase,  on  inhibition  by  histidine,  351 
purine  complexes  of,  739 
pyrimidine  complexes  of,  739 
pyrophosphate  complex  of,  738 
thioglycolate  complex  of,  748-749 
uptake  by  Achromohacter ,  897 
uptake  by  diagragm,  894-897 
uptake  by  erythrocytes,  897,  900-907 
uptake  by  yeast,  898-900 

4-Mercuri-4'-dimethylaminoazabenzene, 
determination  of  tissue  thiols,  767-768 


Mercurimalonamide,  possible  inhibitor  of 
succinate  dehydrogenase,  240 

Mercurin  (mercaptomerin,  Thiomerin),  see 
also  Mercurials,  kidney, 
structure  of,  917 

Mercuripapain,  769-770 

p-Mercuriphenylazo-/3-naphthol,  determi- 
nation of  tissue  thiols,  766-767 

4-(p-Mercuriphenylazo)-l-naphthylamine- 
7  -  sulfonate,  determination  of  tissue 
thiols,  768 

p-Mercuriphenylsulfonate,   see  also   Mer- 
curials and  7;-Mercuribenzoate, 
structure   of,    742 

Mercurochrome,  see  Merbromin 

Mercurophylline,  see  Mercurials,  kidney 

Merphenyl,  see  Phenylmercuric  ion 

Mersalyl,  see  also  Mercurials,  kidney, 
cyclic  mercaptides,  746-747 
renal  transport  of  PAH,  205 
structure  of,  917 

Merthiolate,  see  Thimerosal 

Mesaconate  (methylfumarate), 
fumarase,  275-277,  279 
glutamate   decarboxylase,    328 
ionization  constants,  8 
structure  of,  279 

Meso  porphyrin,  tryptophan  pyrrolase, 
603 

Mesoxalate,  malic  enzyme,  597 

Metanephrine,  urinary  excretion  of, 
pyrogallol,  612 

Metaphen,  see  ISiitromersol 

Metaphosphate,  phosphatases,  439 

Methanediphosphonate,  succinate  dehy- 
drogenase, 243 

Methanedisulfonate  (methionate), 
intercharge  distance,   7 
succinate  dehydrogenase,  243 

Methanesulfonate,  sulfite  oxidase,  451 

Methicillin,  penicillinase,  598 

Methionate,  see  Methanedisulfonate 

L-Methionine, 

intestinal   transport   of  D-methionine, 

265 

synthesis  from  serine,  cyanocobalamin 

analogs,  590 

Methionine  sulfoximine, 


SUBJECT    INDEX 


1195 


glutamine  synthetase,  335 

methionine  incorporation  into  proteins, 

335 
Methotrexate,  see  Amethopterin 
Methoxyacetate,  sarcosine  oxidase,  601 
Methoxybenzoates, 

D-amino  acid  oxidase,  341,  344 

catechol  oxidase,  298-299 
p-Methoxybenzoyl-D-tryptophanamide, 

chymotrypsin,  371 
3-Methoxydopamine, 

dopamine   ^-hydroxylase,   320 

phenylalanine  /^-hydroxylase,  600 
4-Methoxypyridoxal, 

central  nervous  system,  573 

embryogenesis,  576 

GABA  in  brain,  573-574 
Methylacetate,   lipase,   596 
/3-Methylacrylate,  see  Crotonate 
2-Methyladenine, 

adenine  deaminase,  466 

inosine  phosphorylase,  471 
iV-Methyladenine, 

adenine  deaminase,  466 

inosine  phosphorylase,  471 
a-Methylaspartate,     aspartate:a-ketoglu- 

tarate  transaminase,  334 
p-Methylbenzoate,  catechol  oxidase,  298- 

299 
Methylbenzoates,   D-amino  acid  oxidase, 

341 
jS-Methylcysteine,  glyoxylase,  594 
a-Methyldopa, 

blood  pressure,  315 

bronchoconstriction    from    S-hydroxy- 

tryptophan,  315 

cardiac  stimulation  by  dopa,  314-315 

catecholamine  levels  in  tissues,  315-320 

dopa  decarboxylase,  308-320 
in    vivo    inhibition,    314-316 
kinetics,  309-310 
mechanism  of  inhibition,  309-310 

formation  of  abnormal  amine  from,  318 

5-hydroxytryptophan   decarboxylase, 

309-310 

in  vivo  inhibition,  314 

miosis,  315 

pressor  response  to  dopa,  314 


reaction  with  pyridoxal-P,  309 

tissue  levels  of,  318-320 

tyrosine  decarboxylase,  309 

urinary  excretion  of  amines,  315 
A^-Methyldopa,  dopa  decarboxylase,  308 
a-Methyldopamine,   phenylalanine   /3-hy- 

droxylase,  600 
7-Methyldulcitylflavin,  flavokinase,  539 
Methylenesuccinate,   see   Itaconate 
3-Methyl-3-ethylglutarate,  kynurenine:a- 

ketoglutarate  transaminase,  608 
7-Methylfolate, 

blood  pressure,  586 

dopa  decarboxylase,  586 

structure  of,  580 

tyrosine  decarboxylase,  586 
/?-Methylfructofuranoside,  sucrose  trans- 

fructosylase,  421 
Methylfumarate,  see  Mesaconate 
Methylgalactosides,   a-galactosidase,   418 
3-Methylglucarate,   /S-glucuronidase,   424 
3-Methylglucaro-l,4-lactone,  /3-glucuroni- 

dase,  424 
a-Methylglucopyranoside,  sucrose  trans- 

fructosylase,  421 
3-0-Methylglucose, 

galactose  transport,  263 

glucose  uptake,  263 
a  -  Me  thy  Iglu  coside , 

a-amylase,  420 

^-fructofuranosidase,  271 

galactose  transport,  263 

a-glucosidase,  416 

maltase,  271 

a-mannosidase,  422 

phosphorylase,   271-272,   405 

transport  into  E.  coli,  2-deoxyglucose, 

394 
/3-Methylglucoside,  a-glucosidase,  416 
Methylglucosides,   maltose   transglucosy- 

lase,  415 
Methylglucuronates,  /^-glucuronidase,  424 
Methylglucoronides,  /^-glucuronidase,  427 
a  -Methylglutamate , 

aspartate :  a  -  ketoglutarate      transami- 
nase, 334 

glutamate  decarboxylase,  327 

glutamine  synthetase,  336 


1196 


SUBJECT    INDEX 


y-glutamyltransferase,  336 
/9-Methylglutamate, 

glutaminase,  333 

urinary  flow,  333 
a-Methylglutarate, 

kynurenine:a-ketoglutarate     transami- 
nase, 595,  608 
/5-Methylglutarate, 

glutamate  dehydrogenase,  331 

kynurenine:a-ketoglutarate     transami- 
nase, 608 
(S-Methylglutathione,  glyoxylase,  594 
Methylguanidine, 

diamine  oxidase,  362 

histidase,  353 
2-Methylhistine,  phosphoribosyl-ATP  py- 

rophosphorylase,  351 
4  -  Methyl  -  5  -  (/5  -  hydroxyethyl)  thiazole- 

NAD,  inactivation  of  dehydrogenases, 

514 
4  -  Methyl  -  5  - (/?-hydroxyethyl)thiazole-di- 

phosphate,  pyruvate  decarboxylase,  516 
2-Methyl-3-hydroxy-5-hydroxymethylpy- 

rimidine,  pyridoxal  kinase,  565 
a-Methyl-3-hydroxyphenylalanine,  see  a- 

Methyl-m-tyrosine 
a-Methyl-5-hydroxytryptophan,  histidine 

decarboxylase,  353 
Methylindoles, 

tryptophanase,  321 

tryptophan  synthetase,  321 
fi- 1  -Methyl-3-indolylalanine, 

structure  of,  324 

tryptophanase,  323-324 
6-Methylisoxanthopterin,    xanthine    oxi- 
dase, 289 
6-Methyl-ll-ketoprogesterones,     Zl*-3-ke- 

tosteroid  reductase,  450 
Methylmaleate,  see  Citraconate 
Methy  Imalonate , 

accumulation  of,  malonate,   145 

biosynthesis  of,  226 

metabolism  of,  234-235 

occurrence  of,  224-225 

renal   action,   2 
Methylmalonyl-CoA  isomerase,  intercon- 

version  of  methylmalonate  and  succi- 
nate, 235 


yS-Methylmaltoside, 
a-glucosidase,  416 
maltose  transglucosylase,  422 

7-Methylmannitylflavin,  flavokinase,  539 

a-Mannoside,  a-mannosidase,  422 

^-Methylmannoside,  a-glucosidase,  416 

Methylmercuric  ion  (MM),  see  also  Mer- 
curials 

anion  complexes  of,  744-745 
cysteine  complex  of,  747 
protein  complexes  of,  748 

iV-Methylnicotinamide, 
NAD  nucleosidase,  493 
renal  transport  of,  dehydroacetate,  625 
urinary  levels  of,  nicotinylalanine,  610 

2-Methylnicotinate,    glucose    dehydroge- 
nase, 501 

a-Methylnorepinephrine,  formation  from 
a-methyldopa,   318 

5-Methylorotate,  dihydroorotate  dehydro- 
genase, 470 

7-Methylpteroate,    dopa    decarboxylase, 
586 

iV-Methylpyridoxal,  pyridoxal  kinase,  564 

co-Methylpyridoxal,  pyridoxal  kinase,  565 

w-Methylpyridoxol, 

glutamate  decarboxylase  in  brain,  569, 

571 

growth  of  microorganisms,  575 

Methylquinolines,  chymotrypsin,  373 

5-Methylresorcinol,  see  Orcinol 

4-Methylthiazoles,  thiaminase,  524 

D-Methyl-DL-thyroxine,  L-thyroxine  deio- 
dinase,  602 

0-Methyltransferase,  see   Catechol-0-me- 
thyl  transferase 

6  -Methyl  tryptazan , 

maltase  biosynthesis,  326 
L-tryptophan:sRNA  ligase  (AMP),  326 

a-Methyltryptophan,  induction  of  tryp- 
tophan pyrrolase,  325 

)3-Methyltryptophan, 
tryptophan  pyrrolase, 
L-tryptophan:sRNA  ligase  (AMP),  326 

4  -  Methy Itryptophan, 
E.  coli  growth,  323 
tryptophan  synthetase,  312 


SUBJECT    INDEX 


1197 


5-Methyltryptophan, 

anthranilate  synthesis,  321 

maltase  biosynthesis,  326 

tryptophan  pyrrolase,  325 

L-tryptophan:sRNA  ligase  (AMP),  326 
6-Methyltryptophan, 

maltase  biosynthesis,  326 

tryptophan  pyrrolase,  325 

L-tryptophan:sRNA  ligase  (AMP),  326 
a-Methyl-m-t>Tosine, 

aminofcacid  decarboxylase  in  vivo,  315- 

317 

catecholamine  levels  in  brain,  315-316 

dopa  decarboxylase,  308-309 

dopamine  /3-hydroxylase,  320 
0-Methyl-L-tyrosine,  tyrosinase,  305 
Methylurates,  uricase,  284 
Methylurea,  urease,  603,  610 
7-Methylxanthopterin,  xanthine  oxidase, 

289 
Mevalonate,  incorporation  into  sterols, 

mercurials,  886-887 
Mevalonate    dehydrogenase,    mercurials, 

847 
Micrococcus    aureus,    see    Staphylococcus 

aureus 
Micrococcus  lactilyticus,  succinate  dehy- 
drogenase, 

malonate,  33,  49 
Micrococcus  lysodeikticus, 

oxidative  phosphorylation,  mercurials, 

873 

respiration  (endogenous),  malonate,  168 

succinate  dehydrogenase,  malonate,  26 
Micrococcus  pyogenes,  see  Staphylococcus 

aureus 
Micrococcus  sodonensis,  citrate  oxidation, 

malonate,  78 
Microsporum  audouini, 

growth  of,  pyrithiamine,  529 

respiration  (endogenous),  malonate,  168 
Microsporum  canis,  respiration  (endoge- 
nous), 

malonate,  169 
Milkweed  bug,  see  Oncopeltus 
Mitochondria, 

Ca++  uptake, 
malonate,  209 


mercurials,  909-910 

K+  uptake, 
malonate,  209 
mercurials,  909,  914 

Mg++  uptake,  mercurials,  909 

Na+  transport,  mercurials,  909 

swelling  of, 

ferricyanide,  678 
iodine,  689 
malonate,  210-211 
Mitomycin,  degradation  of, 

dehydroacetate,  623 
Mitosis,  see  Cell  division 
Mixed  disulfides,  639-640,  661,  663 
Molybdate, 

cofactor  function,  614 

uptake  of,  tungstate,  614 
Molybdenum,  tissue  levels  of, 

tungstate,  614 
Moniezia  benedeni,  succinate  oxidation, 

malonate,  54 
Monoamine  oxidase,  mercurials,  816,  847 
Monoamidines,  structures  of,  361 
Monobenzone   (Benoquin), 

melanin  formation  in  skin,  304 

tyrosinase,  304 
Monoethyl   peroxide,   catalase,   592,   694 
Monofluorophosphate,  phosphorylase,  406 
L-Monoiodotyrosine,  intestinal  transport 

of, 

L-amino  acids,  265 

malonate,  207-208 
Moraxella  Iwoffi, 

malate  oxidation,  malonate,  81 

oxalacetate  oxidation,  malonate,  82 
Morphine  iV^-demethylase,   analogs,   590, 

597,  604 
Mosquito,  see  Aedes 
Mucate,  see  Galactarate 
Mucinase,  iodine,  686 
Mung  bean,  see  also  Phaseolus  aureus, 

Rb"*"  uptake  in  roots  of,  mercurials,  909 
Muscle,  see  also  Diaphragm 

ATP  levels  in, 
aminopterin,  585 
o-iodosobenzoate,  721-722 
mercurials,  876 

citrate  oxidation,  malonate,  79 


1198 


SUBJECT    INDEX 


contractility,  mercurials,  896,  937-940 
contracture, 

o-iodosobenzoate,  723 

mercurials,  896,  938 

tetrathionate,  699 
creatine-P  levels  in,  o-iodosobenzoate, 
704 

cycle  intermediates  concentrations,  89 
excitability, 

o-iodosobenzoate,  723 

mercurials,  927 
fructose- 1,6-diP  levels,  mercurials,  876 
fumarate  oxidation,  malonate,  81 
glycolysis, 

in  vivo,  malonate,  125-126 

mercurials,  875-877 
glycolysis  (aerobic),  malonate,  129-130 
glycolysis    (anaerobic),    o-iodosobenzo- 
ate, 721-722 

GSH  levels,  tetrathionate,  700 
hydrogen  peroxide,  696 
malonate  levels  in  vivo,  100-102 
malonate  metabolism  in,  228,  233 
membrane  potentials, 

malonate,  212 

mercurials,  937-938 
mercurial  levels  in,  959 
Na+  pump,  malonate,  212 
oxidative   phosphorylation,    malonate, 
119,   121, 

pH  changes  in,  mercurials,  876 
pyrithiamine  levels  in  vivo,  528 
pyruvate  oxidation,  malonate,  75,  115 
respiration,  kojic  acid,  350 
respiration  (endogenous), 

dehydroacetate,  623-624,  629 

malonate,  2,  175,  177-179 
respiration  (glucose), 

malonate  on  insulin  stimulation  of, 

164 
respiratory  quotient,  malonate,  185 
succinate  accumulation,  malonate,  94, 
96 
succinate  dehydrogenase, 

dicarboxylate  ions,  35-37 

malonate,  29,  31-32 
succinate  levels  in  vivo,  malonate,  100- 
102 


succinate  oxidation,  malonate,  55-56 
thiamine-PP  levels,  thiamine  analogs, 
525-526 

Mussel,  see  Mytilus 

Mutarotase  (aldose  1-epimerase), 
analogs,  413-414 

relation  of  inhibition  to  cataract  for- 
mation, 414 

Mycobacteria, 

fatty  acid  biosynthesis,  malonate,  148 
glycerol  oxidation,  malonate,  164 
growth  of,  D-cycloserine,  359 
malonate  formation  in,  226 
phospholipid  biosynthesis, 
malonate,  151 
mercurials,  887 
succinate  dehydrogenase,  oxalacetate, 
36 

Mycobacterium  phlei, 

malonate  metabolism  in,  228 
respiration  (endogenous), 
malonate,  168,  237 
malonic  diethyl  ester,  237 
pyridine-3-sulfonate,  504 
respiration  (glycerol), 

malonic  diethyl  ester,  237 
pyridine-3-sulfonate,  504 
respiration  (lactate),  tartronate,  238 

Mycobacterium  tuberculosis, 

growth  of,  dehydroacetate,  632 

infection    by,    malonic    diethyl    ester, 

224 

malonate  metabolism  in,  228-229 

phospholipid  biosynthesis,  ferricyanide, 

678 

respiration,  ferricyanide,  678 

succinate  dehydrogenase,  malonate,  26 

Mycorrhiza, 

respiration  (endogenous),  malonate,  169 
succinate  oxidation,  malonate,  53 

Myelin,  mercurial  incorporation  into,  950 

Myoglobin,  xanthine:  cytochrome  c  oxi- 
doreductase,  603 

Myokinase,  see  Adenylate  kinase 

Myosin,  see  also  Actomyosin, 

association  with  actin,  o-iodosobenzo- 
ate, 723 
mercurials,  938-940 


SUBJECT    INDEX 


1199 


titration  of  SH  groups,  tetrathionate, 

698 
Mytilus  edulis,  succinate  dehydrogenase, 

malonate,  33,  38 
Myxoma  virus,  infectivity  of, 

mercurials,  978 
Myxomycetes,  see  also  Physarum 

respiration  (endogenous),  malonate,  172 

N 

NAD  (nicotine  adenine  dinucleotide), 
analogs  of, 

formation  from  3-acetylp>Tidine,  496 
formation  from  6-aminonicotinamide, 
503-505 

formation  from  isoniazid,  496 
glutamate  semialdehyde  reductase,  507 
NADH  pyrophosphatase,  506,  511 
NADPH:glutathione   oxidoreductase, 
506,  512 

NADPHmitrite  oxidoreductase,  512 
reaction  with   p-mercuribenzoate,    774 
synthesis  of,  pathways,  486 
tissue  levels  of, 

3-acetylpyridine,  495-496 
6-aminonicotinamide,  505 
a-NAD,   NAD:NADP   transhydrogenase, 

510 
NADase,  see  NAD  nucleosidase 
NAD  glycohydrolase,  see  NAD  nucleosi- 
dase 
NADH, 

NAD  kinase,  510 
NAD  nucleosidase,  491,  493 
oxidation  by  ferricyanide,  673 
oxidation  of, 

o-iodosobenzoate,  703 
malonate,  20 
NADH:aldehyde  oxidoreductase,   GSSG, 

662 
NADH:CoQ   oxidoreductase,   mercurials, 

848 
NADH:cytochrome  c  oxidoreductase, 
o-iodosobenzoate,  710 
mercurials,  848-849,  870 
quinacrine,  552 


NADH:DCPIP    oxidoreductase,    mercu- 
rials, 849 
NADH  dehydrogenase,   mercurials,   798, 

870,  872 
NADH: ferricyanide  oxidoreductase, 

mercurials,  849 

nucleotides,  510 
NADH:hydroxylamine   oxidoreductase, 

quinacrine,  552 
NADHrHgOa    oxidoreductase,    hydrogen 

peroxide,  693 
NADH:lipoamide    oxidoreductase    (lipo- 

amide  dehydrogenase), 

o-iodosobenzoate,  708 

mercurials,     coenzyme     displacement, 

804 
NADH:menadione  oxidoreductase, 

mercurials,  850 

nucleotides,  510 

quinacrine,  552-553 
NADH:methylene    blue    oxidoreductase, 

quinacrine,  553 
NADH:nitrate    oxidoreductase,    see   also 

Nitrate  reductase, 

mercurials,  850 
NADH:nitrite    oxidoreductase,    see    also 

Nitrite  reductase, 

quinacrine,  547,  553 
NADH  oxidase, 

ferricyanide,  676 

hydrogen  peroxide,  693 

o-iodosobenzoate,  710-711 

malonate,  63 

mercurials,  847-848,  872 

nicotinamide,  503 

nucleotides,  511 

quinacrine,  547,  553 
NADH  pyrophosphatase,  nucleotides,  506 

511 
NADH:quinone   oxidoreductase, 

ferricyanide,  676 

mercurials,  850 
pH  effects,  793 
stimulation,  817 
NADHrtetrazolium  oxidoreductase, 

mercurials,  850 

quinacrine,  553 
NAD  kinase,  nucleotides,  509-510 


1200 


SUBJECT    INDEX 


NAD:NADP    transhydrogenase,    nucleo- 
tides, 510 
NAD  nucleosidase  (NAD  glycohydrolase, 

NADase,  DPNase), 

analogs,  485-493 

mercurials,  847 

nicotinamide,  485-493 

nicotinamide  analogs,  504 
NADP   (nicotinamide  adenine   dinucleo- 

tide  phosphate), 

glutamate  semialdehyde  reductase,  507 

NAD  nucleosidase,  489 

NADPH:glutathione   oxidoreductase, 

512 
NAD(P), 

coenzyme  functions  of,  analogs,  500-503 

glutamate  dehydrogenase,  508 

NADPH:cytochrome  c  oxidoreductase, 

511 

nicotinamide  deamidase,  512 
NAD(P)diaphorase,  mercurials  mm'i'o, 926 
NADP   glycohydrolase,    mercurials,    850 
NADPH:  cytochrome  c  oxidoreductase, 

mercurials,  815,  817,  850-851 

nucleotides,  511 
NADPH  dehydrogenase,  benzoate,  349 
NADPH  diaphorase  (old  yellow  enzyme), 

iodine,  688 

mercurials,   891 

porphyrindin,  668 

quinacrine,  548 
NAD(P)H  diaphorase,  quinacrine,  553 
NADPH:glutathione  oxidoreductase,  nu- 
cleotides, 506,  512 
NADPH:menadione  oxidoreductase, 

mercurials,  851 

quinacrine,  553 
NADPH:methemoglobin  oxidoreductase, 

mercurials,  774,  851 

riboflavin,  543 
NAD(P)H  :  methemoglobin     oxidoreduc- 
tase, quinacrine,  553 
NADPHmitrate  oxidoreductase,  see  Ni- 
trate reductase 
NADPHmitrite  oxidoreductase,  see  also 

Nitrite  reductase 

mercurials,  851 

NAD,  512 


NADPH  oxidase,  mercurials,  851 

NADPH:trichloroindophenol  oxidoreduc- 
tase, mercurials,  851 

NADP  reductase  (photosynthetic),  mer- 
curials, 891 

NAD  pyrophosphorylase  (NMN  adenyl- 
transferase),  deamino-ATP,  510 

Nalorphine,     morphine     iV-demethylase, 
590,  597,  604 

a-Naphthol,  chymotrypsin,  373 

Naphthoresorcinol,  tyrosinase,  304 

Naphthylamines,  chymotrypsin,  373 

a-Naphthylmethylmalonate,  chymotryp- 
sin, 370 

Naphthylpropionates,  chymotrypsin,  369- 
370 

1-Naphthyl-sulfate,  arylsulfatase,  443 

2-Naphthyl-sulfate,  arylsulfatase,  443 

Neisseria     gonorrhoeae,     a-ketoglutarate 
oxidation, 
malonate,  79 

Newatodirus  filicollis, 

citrate  oxidation,  malonate,  79 
a-ketoglutarate  oxidation,  malonate,  80 
respiration  (endogenous),  malonate,  174 
succinate  accumulation,  malonate,  94 
succinate  oxidation,  malonate,  54 

Neoplectana  glaseri,  respiration  (endoge- 
nous), 
malonate,  174 

Neohydrin,  see  Chlormerodrin 

Neon,  nitrogen  fixation,  291 

Nephrotoxicity, 
malonate,  219-220 
mercurials,  924-925,  985 

Nerve, 

conduction, 

o-iodosobenzoate,  724 
malonate,  211-212 
mercurials,  949-950 
membrane  potentials, 
malonate,  211-212 
mercurials,  949-950 
thiamine  analogs,  531 
respiration  (endogenous),  malonate,  177 

Neuraminidase,  mercurials,  851 

Neuroblasts, 

o-iodosobenzoate,  724 


SUBJECT    INDEX 


1201 


mercurials,  964-965 

porphyrindin,  670 
Neuromuscular    transmission,     thiamine 

analogs,  531-532 
Neurospora  crassa, 

2-deoxyglucose  utilization  by,  387,  400 

germination    of    ascopores,    malonate, 

195 

growth  of,  oxythiamine,  520,  529 

thiosulfate  utilization,  sulfate,  451-452 
Neurospora  crassa  (poky  strain),  succinate 

oxidation, 

malonate,  53 
Neutrophiles,  phagocytosis  by, 

malonate,  203 
Newcastle    virus, 

inactivation  by  mercurials,  978 

infectivity  of,  mercurials,  978 
Nicotinamidase,  see  Nicotinamide  deami- 

dase 
Nicotinamide, 

D-amino  acid  oxidase,  344 

analogs  of,  484-514 

dehydrogenase  inactivation  by,  503 

glucose  dehydrogenase,   500-502 

glucose-6-P  dehydrogenase,  503 

glutamate  dehydrogenase,   863 

heart,  500 

lactate  dehydrogenase,  500-503 

malate  dehydrogenase,   503 

iV-methylnicotinamide  formation  from, 

495 

NAD  degradation,  485-493 

NADH  oxidase,  503 

NAD  nucleosidase,  485-493 

nicotinamide    riboside    phosphorylase, 

487-488 

6-phosphogluconate  dehydrogenase,  503 

respiration  (endogenous),  503 

respiration  (glucose),  500 

respiration  (lactate),  500 
Nicotinamide    deamidase    (nicotinamide 

deaminase,  nicotinamidase), 

3-acetylp5Tidine,  498 

o-iodosobenzoate,  711 

mercurials,  783 

nucleotides,  512 

pyridine-3-sulfonamide,  504 


Nicotinamide  mononucleotide  (NMN), 
NAD  nucleosidase,  489,  493 
NADPH:glutathione   oxidoreductase, 
512 

Nicotinamide  mononucleotide  adenyl- 
transferase,  see  NAD  pyrophosphoryl- 
ase 

Nicotinamide  riboside,  NAD  nucleosidase, 
492 

Nicotinamide  riboside  phosphorylase,  ni- 
cotinamide, 487-488 

Nicotinate, 

D-amino  acid  oxidase,  342,  344,  346 
catechol  oxidase,  298-299,  301 
glucose  dehydrogenase,  500-502 
glutamate  dehydrogenase,  863 
lactate  dehydrogenase,  500-502 
NAD  nucleosidase,  490,  493 
respiration  (glucose),  500 

Nicotine,  NAD  nucleosidase,  491 

Nicotinic  ethyl  ester,  NAD  nucleosidase, 
493 

Nicotinic-hydrazide,  NAD  nucleosidase, 
493 

Nicotinic-hydrazide-NAD, 
alcohol  dehydrogenase,  497 
lactate  dehydrogenase,  497 

Nicotinylalanine, 

formation  from  tryptophan,  610 
kynureninase,  610 
k\Tiurenine  hydroxylase,  610 
urinary  iV-methylnicotinamide,  610 

Nicotinyl-D-phenylalaninamide,  chymo- 
trypsin,  372 

Nicotinyl  -  d  -  tryptophanamide,  chymo- 
trypsin,  371 

Nicotinyl-D-tyrosinamide,  chymotrypsin, 
371 

Nicotinyl-D-tyrosine  ethyl  ester,  chymo- 
trypsin, 371 

Nicotinyl-L-tyrosinemethylamine,  chymo- 
trypsin, 371 

Nikethamide  (Coramine), 
NAD  nucleosidase,  491 
structure  of,  488 

Nitocra  spinipes, 

toxicity  of  Cu++,  962-963 
toxicity  of  mercurials,  962-963 


1202 


SUBJECT    INDEX 


Nitrate, 

creatine  kinase,  446 

I"  transport  by  ciliary  body,  267 

phosphatase  (acid),  441 

tyrosinase,  301 
Nitrate  reductase,  see  also  NAD(P)H:ni- 

trate  oxidoreductase 

o-iodosobenzoate,  711 

malonate,  65 

quinacrine,  547,  554 

tungstate,  615 
Nitric  oxide, 

hydrogenase,  294 

nitrogen  fixation,  291-292 

physical  properties,   295 
Nitrite,  see  also  Nitrous  acid 

amylases,  660 

oxidation  of, 
bromate,  450 
chlorate,  450 
cyanate,  450-451 
iodate,  450 
phosphate,  450 
Nitrite  reductase,  see  also  NAD(P)H:ni- 

trite  oxidoreductase 
Nitrite  reductase,  mercurials,  860 
Nitrobacter,  growth  of, 

chlorate,  450 
m-Nitrobenzoate, 

glutamate  dehydrogenase,  330 

o-  and  p-nitrobenzoate  metabolism,  612 
p  -  Nitrobenzoate,    tjTosine :  a  -  ketogluta- 

rate  transaminase,  306 
Nitrobenzoates, 

D-amino  acid  oxidase,  341,  348 

catechol  oxidase,  298-299 

metabolism  of,  analogs,  612-613 
p-Nitrobenzoyltryptophans,    chymotryp- 

sin,  374 
3-o-Nitrobenzyl-4-methylthiazole,  thiami- 

nase,  524 
5-Nitrofuroate,   glutamate   dehydrogena- 
se, 330 
Nitrogen,  physical  properties,  295 
Nitrogenase,  analogs,  291-296 
Nitrogen  fixation,  analogs,  291-296 
Nitromersol    (Metaphen),    structure    of, 

970 


o-Nitrophenol, 

catechol  oxidase,  298 

m-    and    p-nitrobenzoate    metabolism, 

612 

2?-Nitrophenol, 

D-amino  acid  oxidase,  348 

catechol  oxidase,  298 

o-  and  p-nitrobenzoate  metabolism,  612 

polyphenol  oxidase,  297 

respiration  (endogenous),  297 

p-Nitrophenylacetate,   carboxypeptidase, 
365 

p-Nitrophenylalanine,     tyrosine  :  a  -  keto- 
glutarate  transaminase,  306 

o-Nitrophenylpyruvate,  pyruvate  decar- 
boxylase, 431 

3-Nitropropionate    (hiptagenate),     succi- 
nate dehydrogenase,  244 

Nitroreductase,  quinacrine,  554 

2-Nitroresorcinol,  tyrosinase,  304 

Nitrosoreductase,  quinacrine,  554 

3-Nitro-L-tyrosine,  tyrosinase,  305 

Nitrous  acid,  see  also  Nitrite, 
enzyme  group  oxidation,  657 

Nitrous  oxide, 

nitrogen  fixation,  291-294 
physical  properties,  295 

NMN,  see  Nicotinamide  mononucleotide 

Nocardia  corallina, 

malonate  formation  in,  226 

malonate  occurrence  in,  225 

nitrobenzoate  metabolism,   analogs, 

612-613 

succinate  accumulation,  malonate  90 

Nonanoate,   kynurenine: a-ketoglutarate 
transaminase,  608-609 

Norepinephrine,  see  also  Catecholamines, 
biosynthesis  of,  pathways,  307 
brain  levels, 

dopa  decarboxylase  inhibitors,  316- 
318 

a-methyl-m-tyrosine,  316 
pyrogallol,  611 
heart  levels,  dopa  decarboxylase  inhibi- 
tors,  316-318 

tissue  levels,  a-methyldopa,  315-320 
urinary  excretion,  pyrogallol,  612 

Norleucine,  arginase,  337 


SUBJECT    INDEX 


1203 


Normetanephrine,  urinary  excretion  of, 

pyrogallol,  612 
D-Norvaline,     L-alanine     dehydrogenase, 

354 
L-Norvaline,  arginase,  337 
Nuclei, 

ATP  level,  malonate,  189 

COa  formation  from  glucose,  2-deoxy- 

glucose,  393-394 

glucose  utilization,  dehydroacetate,  624 

respiration,  malonate,  189 

respiration    (glucose),    dehydroacetate, 

624 
Nucleic  acids,  see  also  Deoxyribonucleates 

and  Ribonucleates 

biosynthesis    of,    fluorophenylalanines, 

478-481 

chymotrypsin,  457 

fumarase,  465 

lysozyme,  459 

p-mercuribenzoate  complex  with,  744 

mercuric  complexes  with,  741 
Nucleosidediphosphate :  polynucleotide 

nucleotidyltransferase,  see  Polynucleo- 
tide phosphorylaes 
5 '-Nucleotidase,  analogs,  471-472 
Nucleotide   incorporation   enzyme,    mer- 
curials, 817 


Oats,  see  also  Avena, 

malonate  occurrence  in,  224 

Octanoate,  see  Caprylate 

Oenanthate,  see  Heptanoate 

Old  yellow  enzyme,  see  NADPH  diapho- 
rase 

Oleyl-CoA,  acetyl-CoA  carboxylase,  614 

Oncopeltus  fasciatus,  succinate  dehydro- 
genase, 
malonate,  29 

Onion  roots,  growth  of, 
mercurials,  966 

Ophthalmate,  glyoxylase,  594 

Opsanus  tau  (toadfish)  pancreas,  respira- 
tion (endogenous), 
malonate,   175 

Opsopyrrole-dicarboxylate,  porphobilino- 
gen deaminase,  600 


Orcinol  (5-methylresorcinol), 

catechol  oxidase,  297-298 

structure  of,  296 

tyrosinase,  304 
Ornithine, 

arginase,  335,  337-338 

arginine  uptake  in  ascites  cells,  338 

histidase,  353 

kynureninase,  595 
D-Ornithine,  carbamyl-P:ornithine  trans- 

carbamylase,  592 
Ornithine  carbamyltransferase  (carbamyl- 

P:L-aspartate  carbamyltransferase), 

analogs,  592 

o-iodosobenzoate,  711 

mercurials,  851 

protection  by  substrates,  783 
Ornithine  decarboxylase,  permanganate, 

660 
Orotate,  dihydroorotase,  470 
Orotate  transphosphoribosylase,  analogs, 

473 
Orotidylate  decarboxylase,  analogs,  472- 

473,  479 
Orthosphosphate,  see  Phosphate 
Orthophosphoric   diester   phosphohydro- 

lase,  see  Phosphodiesterase 
Ova  (rabbit),  respiration  (endogenous), 

malonate,  183 
Ovarian  germinal  epithelium,  mitosis  in, 

malonate,  200 
Oxacillin,  penicillinase,  598 
Oxalacetate, 

glyoxylate  reductase,  438 

glyoxylate  transacetatse,  594 

/9-hydroxybutjrrate  dehydrogenase,  594 

keto-enol  tautomerism  of,  39 

malate  dehydrogenase,  596 

malate    dehydrogenase   (decarboxylat- 

ing),  597 

oxalosuccinate  decarboxylase,  597 

oxidation  of,  malonate,  82 

pyruvate  decarboxylase,  430 

succinate  dehydrogenase,  36,  38-42 

utilization  of,  malonate,  66-68 
Oxalacetate  decarboxylase, 

2'-AMP,   507 

analogs,  597 


1204 


SUBJECT    INDEX 


malonate,  63 

Oxalacetate  ethyl  ester,  succinate  dehy- 
drogenase, 39 

Oxalate, 

chelation  with  cations,  12 
glutaniate  decarboxylase,  328 
glycolysis  (anaerobic),  414 
glyoxylate  transacetatase,  594 
D-hydroxy   acid    dehydrogenase,    435- 
437 

intercharge  distance,  6 
ionic  length  and  volume,  188 
kynurenine:a-ketoglutarate     transami- 
nase, 608 

D-lactate:cytochrome  c  oxidoreductase, 
435 

lactate  dehydrogenase,  435-436 
D -lactate  dehydrogenase,  437 
permeability  of  erythrocytes  to,  188 
phosphatase  (acid),  442 
pyruvate  decarboxylase,  430 
succinate  dehydrogenase,  35 
tartronate  semialdehyde  reductase,  602 
tyrosinase,  300 

Oxalate  decarboxylase, 
malonate,  63 
mercurials,  860 

Oxalomalate,  see  also  y-Hydroxy-a-keto- 
glutarate, 

decarboxylation  to  y-hydroxy-a-keto- 
glutarate,  616 

Oxalosuccinate  decarboxylase, 
analogs,  597 
malonate,  63 
mercurials,  852 

Oxamate, 

Crabtree  effect,  435 
glucose  utilization,  435 
glycolysis  (aerobic),  434 
glyoxylate  reductase,  438 
lactate   dehydrogenase,   432-434 
Na+  fluxes  in  HeLa  cells,  434 
pentose-P  pathway,  435 
pyruvate  decarboxylase,  430 
pyruvate  oxidation,  434 
respiration  (glucose),  435 
structure  of,  432 
toxicity,  434 


Oxanilate,  pyruvate  decarboxylase,  431 
Oxidants,  see  also  specific  oxidants 

degree  of  SH  group  oxidation,  656-657 

enzyme    inhibitions,    mechanisms    of, 

657-658 

oxidation  of  protein  groups,  657-658 
Oxidative  phosphorylation, 

benzoate,  348 

Ca++,  453 

dehydroacetate,  623 

galactoflavin,  544 

iodide,  689 

iodine,  688-689 

o-iodosobenzoate,  722 

malonate,    118-122 

mercurials,  872-874,  927 

quinacrine,  556-557 

thiophosphate,  447-448 
Oxidized  glutathione  (GSSG),  see  Gluta- 
thione (oxidized) 
u-Oximinoglutarate,  glutamate  decarbo- 
xylase, 327 
<5-Oximinolevulinate,  aminolevulinate  de- 

hydrase,  591 
2-Oxoglutarate,  see  a-Ketoglutarate 
2-Oxo-4-imidazolidinecaproate,   see   Des- 

methyldesthiobiotin 
Oxybiotin,  structure  of,  588 
Oxybiotinsulfonate,  yeast  fermentation, 

588-589 
Oxygen, 

hydrogenase,  293 

inactivation  of  enzymes,  658-659 

nitrogen  fixation,  291-294 

physical  properties,  295 

pyruvate  oxidase,  659 

respiration  of  brain,  658 

SH  group  oxidation  by,  658 

uptake,  see  Respiration 
Oxypyrithiamine,  structure  of,  517 
Oxythiamine, 

blood  lactate,  520 

blood  pressure,  532 

blood  pyruvate,  520,  527 

central  nervous  system,  527,  530 

glycogen  in  liver,  520 

heart  rate  in  vivo,  527 

miosis,  532 


SUBJECT    INDEX 


1205 


nerve  membrane  potentials,  531 

Neurospora  growth,  520,  529 

phosphorylation  of,  519 

pyruvate  accumulation,  520 

pjTuvate  oxidation  in  vivo,  529-521 

rat  growth,  527 

structure  of,  517 

thiaminase,  523-524 

thiamine  deficiency,  530-532 

thiamine  kinase,  522-523 

thiamine  levels  in  tissues,  525-527 

toxicity,  516,  530-531 

transketolase  in  vivo,  522 

urinary  thiamine,  525 

Vibrio  growth,  522 
Oxythiaminediphosphate, 

acetoin  formation  from  pyruvate,  519 

pyruvate  decarboxylase,  518 

transketolase,  519 
Oxythiaminetriphosphate, 

pyruvate  decarboxylase,  518 

pyruvate  oxidase,  519 
Oxyurea,  urease,  603,  610 
Oyster,  see  also  Crassostrea  and  Saxostrea 
Oyster  eggs, 

glycolysis,  mercurials,  874-875,  884 

respiration    (endogenous),    mercurials, 

882,  884 

succinate  oxidation,  malonate,  22 
Oyster  muscle,  succinate  oxidation, 

malonate,  22 
Oyster  spermatozoa, 

glycolysis,  mercurials,  884 

respiration    (endogenous),    mercurials, 

882,  884 


Palmitate,  oxidation  of, 

2-deoxyglucose,  397 
Palmityl-CoA,  citrate  synthetase,  614 
Pancreas,  respiration  (endogenous), 

malonate,  175,  177 
Pantetheine,  structure  of,  587 
Pantoate:^-alanine  ligase  (AMP), 

acetate,  597 

^-alanine  analogs,  597-598 
Pantothenate, 

analogs  of,  586-588 


acetyl  transfer,  587 

bacterial  growth,  587 

choline  acetylase,  587 

coenzyme  A  formation,  586-587 

pantothenate  biosynthesis,  588 
biosynthesis  of,  analogs,  588 
conversion  to  coenzyme  A,  586-587 
structure  of,  587 
Pantoylaminoethanethiol, 

coenzyme  A  formation  from  pantethe- 
ine, 587 

structure  of,  587 
sulfonamide  acetylation,  587 
Pantoyltaurine, 

bacterial  growth,  587 
choline  acetylase,  587 
structure  of,  587 
pApA,  phosphodiesterase,  473 
Papain, 

analogs,  375 
ferricyanide,  673,  676 
hydrogen  peroxide,  691,  693-694 
iodine,  682-683,  686 
mercurials,  769-770,  804 

crystalline    mercuric    ion    complex, 

769-770 

relation  to  SH  groups,  804 
porphyrindin,  667-668 
succinyl  peroxide,  694 
Papilloma,  malonate  and  succinate  levels 

in  vivo,  102 
Paracentrotus  lividus  eggs,  development 
of, 

o-iodosobenzoate,    726-727 
Paramecium  caudatum, 

cycle  intermediate  oxidations  in,  ma- 
lonate, 79-82 

mercurial  toxicity,  981-982 
motility  of, 

<rans-aconitate,  274 

malonate,  203 

mercurials,  981 
pyruvate  oxidation,  malonate,  74 
respiration  (endogenous), 

iraws-aconitate,   273 

malonate,  173 
succinate  dehydrogenase,  malonate,  28, 
50 


1206 


SUBJECT    INDEX 


Parasorbic  acid,  617 

Paris  daisy  stem  cultures,  growth  of, 

malonate,  197 
Pasteurella  muUocida, 

fumarate  oxidation,  malonate,  81 

glutamate  oxidation,  malonate,  187 

isocitrate  oxidation,  malonate,  79 

pyruvate  oxidation,  malonate,  74 
Patulin  (clavacin),  617 
Pea  leaves, 

malonate  occurrence  in,  224 

succinate  accumulation,  malonate,  91 
Peanut  cotyledons,  fatty  acid  oxidation, 

malonate,  136 
Peanut  mitochondria, 

butyrate  oxidation,  malonate,   137 

malonate  metabolism  in,  228,  231 

phospholipid    biosynthesis,    malonate, 

151 

propionate  metabolism,  malonate,  145 
Pea  roots,  mitosis  in, 

malonate,  197 
Peas,  sea  also  Pisum  sativum, 

respiration  (endogenous),  malonate,  170 
Pea  seedlings, 

a-ketoglutarate  utilization,  malonate,  84 

oxidative   phosphorylation,    malonate, 

122 
Pea  stems,  growth  of, 

mercurials,  966 
Peloscolex    velutinus,    respiration    (endo- 
genous), 

malonate,  174 
Penicillic  acid,  617 
Penicillinase, 

analogs,  598-599,  615,  688 

configurational  changes,  analogs,  249 

ferricyanide,  676 

iodine,  688 

o-iodosobenzoate,  711,  717 
pH  effects,  717 

mercurials,  pH  effects,  793 
Penicillin-G,  renal  transport  of, 

dehydroacetate,  625 
Penicillium  chrysogenum, 

acetate  oxidation, 
malonate,  77 
malonic  diethyl  ester,  237 


lactate  oxidation,  malonate,  78 
respiration  (endogenous),  malonate,  169 
succinate  dehydrogenase, 
malonate,  26 
oxalacetate,  36 

Penicillium  citrinum,  growth  of, 
dehydroacetate,  633 

Penicillium  cyclopium,   malonate   occur- 
rence in,  226,  228 

Penicillium   digitalum,   growth   of, 
dehydroacetate,  632 

Penicillium  expansum.,  growth  of, 
dehydroacetate,  632 

Penicillium    funiculosum,    malonate    oc- 
currence in,  225 

Penicillium  notatum,  growth  of, 
mercurials,  973 
resistance  to  mercurials,  983 

Penicillium    oxalicum,    respiration    (glu- 
cose), 
malonate,  133-134 

Penicillium  roqueforti,  tolerance  to  mer- 
curials, 983 

Pentanoate,  see  Valerate 

Pentose-phosphate    isomerase,    o-iodoso- 
benzoate, 711 

Pentose-phosphate  pathway, 
2-deoxyglucose,  393-394 
malonate,    130-132 
mercurials,  885-886 
oxamate,  435 

pyrithiamine-resistant  S.  aureus,  529 
quinacrine,  560 

Pepsin 

dehydroacetate,  622 
iodine,  680,  682-683,  688 
macroions,  457-458 
mercurials,  860 
permanganate,  657 

Peptidases,  mercurials,  860 

Periodate, 

chymotrypsin,  657 
/S-fructofuranosidase,  660 
ovalbumin  SH  groups,  657 
protein  oxidation  by,  657 

Periplaneta  americana,  succinate  dehydro- 
genase, 
malonate,  29 


SUBJECT    INDEX 


1207 


Periwinkle  stem  cultures,  growth  of, 

malonate,  197 
Permanganate, 

enzyme  inhibitions,  659-660 

Fiisarium    conidial   growth,   660 

insulin,  657 

oxidation, 

of  amino  acids,  657 
of  casein,  657 
of  proteins,  657 

pepsin,  657 
Permeability, 

to  dicarboxylate  ions,  188 

to  malonate,   186-192 

to  mercurials,  900-907 
Peroxidase, 

jj-coumarate  analogs,  599 

iodine.  682,  686 

mercurials,  860 
Peroxidases,  see  also  Hydrogen  peroxide, 

Monoethyl  peroxide,  and  Succinyl  per- 
oxide, 

acetylcholine  response,  696 

chemical  properties,  690-691 

enzyme  inhibitions,  691-694 

glycolysis,  695 

intestine,  696 

muscle,  696 

respiration,  695 

toxicity,  696 

tumor  growth,  695 
pH,  effects  of, 

ADP  on  ATPase,  445 

carbobenzoxyglutamate  on  papain,  375 

copper  on  ^-glucuronidase,  795 

dehydroacetate, 

on  growth  of  microorganisms,  633 
on  succinate  dehydrogenase,  622 

fatty  acids  on   kynurenine    transami- 
nase, 609 

ferricyanide  oxidation  of  hemoglobin, 

671 

hydrogen  peroxide  on  ATPase,  691 

iodine, 

bactericidal  action,  690 
on  enzymes,  688 
oxidation  of  cysteine,  680 

o-iodosobenzoate, 


on  enzymes,  716-717 

oxidation  of  SH  groups,  657,  702 
macroionic   inhibitions,   454-457,    461, 
464 
malonate, 

on  cardiac  metabolism,  214-215 

on  succinate  oxidation,  51,  56 

permeability  to,  189-192 
mercurials, 

on  ATPase,  867 

on  enzymes,  790-797 

on  K+  efflux,  898 

on  luminescence,  889 

reactions  with  proteins,  760-761 

reactions  with  SH  groups,  749-750 

titration  of  3-phosphoglyceraldehyde 

dehydrogenase,  806 

toxicity  to  heart,  944-945 
mercuric  ion, 

activation  of   glycerate-2,3-diphos- 

phatase,  820 

formation  from  organic   mercurials, 

931-933 
metal  ion  reactions  with  thiols,  638 
methylurea  on  urease,  610 
oxidation  of  enzyme  SH  groups,  664 
quinacrine   on   D-amino   acid   oxidase, 
557-558 

silver  on  y3-glucoronidase,  795 
Phage  (coli),  see  Coliphage 
Phage  (staphylococcal),  see  Staphylococ- 
cal phage 
Phagocytosis, 

o-iodosobenzoate,  728 
kojic  acid,  350 
malonate,  223 
Phaseolus  aureus,  see  also  Mung  bean, 
citrate  oxidation,   malonate,   78 
a-ketoglutarate  oxidation,  malonate,  80 
oxidative    phosphorylation,    malonate, 
120 
Phaseolus  coccineus  (runner  bean), 
malonate  occurrence  in,  225 
respiration,  malonate,  225 
Phaseolus  seeds,  glutamate  metabolism, 

pyrithiamine,  522 
Phaseolus  vulgaris  (bush  bean), 

malonate  metabolism  in,  226,  228,  232 


1208 


SUBJECT    INDEX 


malonate  occurrence  in,  225-226 

succinate  dehydrogenase,  malonate,  33 
2-Phenantryl-sulfate,  arysulfatase,  443 
Phenbenicillin,   penicillinase,   598 
Phenethicillin,   penicillinase,   598 
Phenol, 

D-amino  acid  oxidase,  348 

dehydroshikimate  reductase,   606 
Phenol-a-glucoside, 

/3-glucosidase,  417 

taka-/3-glucosidase,  271 
Phenol  oxidase,  see  also  Catechol  oxidase 

analogs,  296-302 
Phenol  red,  renal  transport  of, 

dehydroacetate,  626 

malonate,  205 

mercurials,  921 
Phenol  sulfokinase,  3'-AMP-5'-P,  473 
Phenolsulfonphthalein,  renal  transport  of, 

dehydroacetate,  625-626 
2-Phenoxyethanol,  chymotrypsin,  370 
Phenylacetamide,  chymotrypsin,  372 
Phenylacetate, 

D-amino  acid  oxidase,  342,  346 

ammonia  formation  in  kidney,  348 

carboxypeptidase,   363-366 

catechol  oxidase,  298 

chymotrypsin,  370,  372 

dopa  decarboxylase,  312 

glutamate  decarboxylase,  329 

p-hydroxyphenylpuruvate  oxidase,  306 

tyrosinase,  300-301 
Phenyl-iV-acetylglucosaminide,  iV-acetyl- 

/3-galactosaminidase,  420 
Phenylalaninamides,  cathepsin  C,  375 
Phenylalanine, 

analogs  of,  dopa  decarboxylase,  308 

blood  serotonin,  325 

brain   serotonin,   325 

dipeptidase,  367 

glutamate  decarboxylase,  329 

histidine  decarboxylase,   353 

pyruvate  decarboxylase,  430 

tryptophan  hydroxylase,  325 
D  -Phenylalanine , 

carboxypeptidase,  365,  367 

E.  coli  growth,  268 


L-Phenylalanine, 

amino  acid  transport  by  brain,  266 

arginase,  337 

carboxypeptidase,  366 

cathepsin  C,  375 

tyrosinase,  305 

tyrosine:a-ketoglutarate  transaminase, 

306 
Phenylalanine  activating  enzyme,  see  l- 

PhenylalanineisRNAligase  (AMP) 
Phenylalanine  deaminase, 

analogs,  355 

mercurials,  852 
Phenylalanine     ^-hydroxylase,     analogs, 

354,  599-600 
L-Phenylalanine:sRNA  ligase  (AMP),  ana- 
logs, 354-355 
Phenylalanylglycine,  cathepcin  C,  375 
Phenylbutyramide,  chymotrypsin,  372 
a-Phenylbutyrate, 

cholesterol  biosynthesis,  614 

fatty  acid  biosynthesis,  614 
y-Phenylbutyrate, 

carboxypeptidase,  365-366 

chymotrypsin,  370,  372 

kynurenine:a-ketoglutarate     transami- 
nase, 608-609 
Phenylethylam  ine , 

dopa  decarboxylase,  308 

dopamine  ^-hydrolase,  320 

phenylalanine  ^-hydroxylase,  600 
Phenyl-a-glucopyranoside,  a-glucosidase, 

416,  423 
Phenylglucosides, 

a-amylase,  420 

maltose  transglucosylase,  415 
Phenylglycine,  derivatives  of, 

dopa  decarboxylase,  312 
Phenylglyoxylate,   glyoxylate  reductase, 

438 
Phenylisocyanate,  urease,  649 
Phenylketonuria   (phenyl pyruvate   oligo- 
phrenia), 

inhibition  of  amino  acid  transport  into 

brain,  266 

inhibition  of  dopa  decarboxylase,  314 

inhibition  of  glutamate  decarboxylase, 

329 


SUBJECT    INDEX 


1209 


inhibition    of    pyruvate    metabolism, 

429-430 

inhibition  of  tryptophan  hydroxylase, 

325-326 

inhibition  of  tyrosinase,  305 

Phenyllactate, 

glutamate  decarboxylase,  329 
pyruvate  decarboxylase,  430 
tryptophan  hydrolase,  325 

Phenylmercuric  acetate,  see  Phenylmer- 
curic  ion 

Phenylmercuric  ion  (PM),  see  also  Mer- 
curials, 
amino  acid  complexes  of,  744 

3-Phenyl-4-methylthiazole,   thiaminase, 
524 

Phenylphosphate,  arylsulfatase,  443 

Phenylpropionamide,  see  Hydrocinnamide 

y3-Phenylpropionate,  see  Hydrocinnamate 

Phenylpyruvate, 

accumulation  in  phenylketonuria,  429- 

430 

acetoin  formation  from  pyruvate,  430 

dopa  decarboxylase,  312-314 

glutamate  decarboxylase,  329 

glycerate  dehydrogenase,  430 

33-hydroxyphenylpyruvate   oxidase, 

305-306 

lactate  dehydrogenase,  437 

melanin  formation,  305 

metabolism  of,  429-430 

pyruvate  decarboxylase,  431-432 

pyruvate  metabolism,  430 

pyruvate  oxidase,  430 

tryptophan  hydroxylase,  325 

tyrosinase,  305 

Phenylpyruvate  oligophrenia,  see  Phenyl- 
ketonuria 

/3-Phenylserine, 

phenylalanine  deaminase,  355 
L-phenylalanine:sRNA    ligase    (AMP), 
354 

Phenylsulfate,  arylsulfatase,  443 

Phloretin-phosphate  polyesters,  hyaluro- 
nidase,    461 

Phlorizin, 

mitochondrial  swelling,  210 
renal  transport  of  PAH,  205 


Phloroglucinol, 

D-araino   acid   oxidase,   344 

catechol  oxidase,  297 

tyrosinase,  304 
Phloroglucinol-phosphate   polymer,   hya- 

luronidase,  461 
Phosphatase,  see  also  Phosphatase  (acid) 

and  Phosphatase  (akaline), 

analogs,  439-443 

arsenate,   439-440 

borate,  439-440 

ferricyanide,  676 

iodine,  686 

mercurials  in  vivo,  926-927 

oxidation  of,  657 

permanganate,  660 

silicate,   439-440 

substrate  inhibition,  439 

tartrates,  440-442 
Phosphatase  (acid), 

active  center  of,  440-442 

alginate,   464 

anionic  polymers,  443 

o-iodosobenzoate,  711,  718 

macroions,  464-465 

malonate,   63 

mercurials,  772-773,  778,  817 

Mg++  analogs,  452-453 

polyestradiol- phosphate,  464 

polyphloretin-phosphate,    464 

tartronate,  238 
Phosphatase  (alkaline), 

5-fluorouracil,   formation  of  abnormal 

enzyme,  479 

o-iodosobenzoate,  711 

mercurials,  772-773,  817,  860,  926 
Phosphate, 

active  transport  of,  malonate,  209-210 

arylsulfatase,  443-444 

carbamyl  phosphatase,  439 

choline  sulfatase,   444 

creatine  kinase,  446 

glucose  dehydrogenase,  501 

glucose-6-P  dehydrogenase,   411 

glycolysis  (anaerobic),  414 

nitrite  oxidation,  450 

phosphatases,  439 

phosphodeoxyribomutase,  412 


1210 


SUBJECT    INDEX 


3-phosphoglyceraIdehyde      dehydroge- 
nase, 409 

phosphopentose  isomerase,  411 

pyrophosphatase,  439 

renal  transport  of, 
dehydroacetate,  625 
malonate,  205 

ribulose-P  carboxylase,  412 

serum  level  of,  malonate,  219 

transaldolase,  412 

transketolase,  412 

triose-P  isomerase,  412 

uptake    by    erythrocytes,    mercurials, 

910 

uptake  by  staphylococci, 
analogs,  267 

mercurials,  910,  912-913 
urinary  excretion  of,  malonate,  206 
yeast  level  of,  mercurials,  885 
Phosphatidate  phosphatase,  cystine,  662 
3'-Phosphoadenosine-5'-phosphosulfate 

reductase,  quinacrine,  554 
Phosphoarabinose  isomerase,  analogs,  411 
Phosphodeoxyribomutase, 

analogs,  413 

phosphate,  412 
Phosphodiesterase    (orthophosphoric    di- 

ester  phosphohydrolase),  analogs,  473 
Phosphoenolpyruvate  carboxylase, 

o-iodosobenzoate,  711 

mercurials,  852 
Phosphoenolpyruvate  carboxytransphos- 

phorylase,  mercurials,  852 
Phosphofructokinase, 

cycle  ingermediates,  385-386 

cycHc  3  ,5 -AMP,  474 

mercurials,  852 
Phosphoglucomutase, 

l,5-anhydro-D-glucitol-6-P,  379 

D-glucosamine,  382 

glucose-6-P,  413 

GSSG,  662 

o-iodosobenzoate,  711,  716 

malonate,   130 

mercurials,  804,  810, 
rate  of  inhibition,  810 
relation  to  SH  groups,  804 

oxygen  inactivation  of,  659 


6-Phosphogluconate,  phosphoglucose  iso- 
merase, 406 
Phosphogluconate   dehydrogenase, 
2'-AMP,  507 
o-iodosobenzoate,  711 
nicotinamide,  503 
Phosphoglucose  isomerase, 
analogs,  406-407 
l,5-anhydro-D-glucitol-6-P,  379 
ATP,  474 

2-deoxyglucose-6-P,  390 
o-iodosobenzoate,   711 
mercurials,  839 
3-Phosphoglyceraldehyde,  anolase,  409 
3-Phosphoglyceraldehyde   dehydrogenase 
(triose-P  dehydrogenase), 
active  center  of,  641 
6-aminonicotinamide  in  vivo,  505 
analogs,  408-409 
cystamine  monosulfoxide,  663 
ferricyanide,  676-677 
p-fluorophenylalanine,   replacement  of 
phenylalanine  in,  351 
GSSG,  661-662 
hydrogen  peroxide,   694-695 
iodine,  683,  687 
o-iodosobenzoate,  704,  714 
malonate,  63 

mercurials,  650,  766,  770,  776,  783, 
785-786,  788,  802,  804,  806,  810,  812, 
817,  824,  826-827,  852,  876 

coenzyme  displacement,  785-786 

denaturation,  788 

protection,  783 

rate  of  inhibition,  810 

relation  to  SH  groups,  802,  804,  812 

reversal  by  cysteine,  824,  826 

serum  effect,   776 

spontaneous  reversal,  827 

stimulation,    817 

titration  of  SH  groups,  806 
p-mercuribenzoate,  650,  766,  852 

titration  of  SH  groups,  766 
mercuric  ion,   crystalline   complex   of, 
770 

oxygen  inactivation  of,  659 
porphyrexide,  668 
pterin-6-aldehyde,  288 


SUBJECT    INDEX 


1211 


succinyl  peroxide,  694 

tetrathionate,  699 
2-Phosphoglycerate,  glycerate-2,3-diphos- 

phatase,  413 
3  -Phosphogly  cerate , 

enolase,  410 

glycerate-2,3-diphosphatase,  413 
Phosphoglycerate     kinase,     o-iodosoben- 

zoate,  711 
Phosphoglycerate  mutase,  mercurials,  852 
Phosphoglycerol  dehydrogenase,  o-iodoso- 

benzoate,  711 
Phosphohalidase,  see  DFPase 
Phospholactate,  enolase,  410 
Phospholipase,   mercurials,   860 
Phospholipids,  biosynthesis  of, 

malonate,   151 

mercurials,  887 
Phosphomevalonate  kinase, 

o-iodosobenzoate,   711 

mercurials,  852 
Phosphomonoesterase,  mercurials,  788 
Phosphonoacetate,  succinate  dehydroge- 
nase, 243 
/S-Phosphopropionate, 

intercharge  distance,  7 

succinate  dehydrogenase,  242-243 
Phosphopentose  isomerase,  see  also  Phos- 

phoarabinose  isomerase 

analogs,  411 
Phosphopyruvate  carboxylase,  see  Phos- 

phoenolpyruvate    carboxylase 
Phosphopyruvate  hydratase,  see  Enolase 
Phosphoribomutase,   2,3-diphosphoglyce- 

rate,  413 
5-Phosphoribonate, 

phosphopentose  isomerase,  411 

phosphoribulokinase,  413 
Phosphoribosyl-ATP  pyrophosphorylase, 

feedback  inhibition  by  histidine,  351 

trypsin  inactivation  of,  351 
Phosphoribosylpyrophosphate         amido- 

transferase, 

ADP  and  ATP,  474 

mercurials,  853 
Phosphoribulokinase,  5-phosphoribonate, 

413 


Phosphorylase, 

analogs,  405-406 

D-glucosamine,  382 

mercurials,  648-649,  789,  803-804,  808, 

811-813,  852-853 

rate  of  inhibition,  811-812 

relation  to  SH  groups,  803-804,  808, 

813 

splitting  into  subunits,  789 

a-methylglucoside,  271-272 

SH  groups  of,  649 
Phosphorylase  kinase,  Ca++,  453 
Phosphorylphosphatase,   mercurials.   817 
0-Phosphoserine  phosphatase, 

alanine,  270-271 

mercurials,   853 

serine,  270 
Phosphothreonine,     L-threonine    synthe- 
tase, 357 
Phosphotransacetylase, 

lipoate  analogs,  590 

Na+  and  Li+,  452 
Phosphotransferases,  see  also  Kinases, 

analogs,  444-447 
Photophosphorylation , 

mercurials,  892 

quinacrine,  557 
Photosynthesis , 

a  -  hydroxy  -  2  -  pyridinemethanesulfo- 

nate,  439 

malonate,  163-164 

mercurials,  891-892 

threose-2,4-diphosphate,   409 

xylose,  414 
Phthalate, 

D-amino  acid  oxidase,  341,  344 

aspartate :  a  -  ketoglutarate      transami- 
nase, 334 

glutamate  dehydrogenase,  331 

intercharge  distance,  6 

ionization  constants,  8 

kynurenine:a-ketoglutarate     transami- 
nase, 607-608 

succinate  dehydrogenase,  37 

tyrosinase,  300 
m-Phthalate,  see  Isophthalate 
p-Phthalate  see  Terephthalate 
Phycomyces  blakesleeanus, 


1212 


SUBJECT    INDEX 


carbohydrate     metabolism,     thiamine 

analogs,  529 

resistance  to  pyrithiamine,  529 

Physarum  polycephalum ,  succinate  dehy- 
drogenase, 
malonate,  28 

Picolinate,  NAD  nucleosidase,  488 

Pimelate, 

aspartate :  a  -  ketoglutarate      transami- 
nase, 334 

glutamate  decarboxylase,  328 
ionization  constants,  8 
kynurenine:a-ketoglutarate     transami- 
nase, 595,608 

Pink  disease,  see  Acrodynia 

Pinus  lambertiana  (sugar  pine),  succinate 
dehydrogenase, 
malonate,  27 

Pinus  taeda  (loblolly  pine)  roots,   phos- 
phate uptake, 
malonate,  209 

iV-Piperidinomethylnicotinamide,   inhibi- 
tion of  various  oxidations,  503 

Pisuni  sativum,  succinate  dehydrogenase, 
malonate,  27 

Pituitary,  succinate  dehydrogenase, 
malonate,  31 

Placenta, 

acetate  oxidation,  malonate,  78 
lipid  biosynthesis,  malonate,  147,  149 
malonate  metabohsm  in,  228,  233 
pyruvate  oxidation,  malonate,  77 
respiration  (endogenous),  malonate,  179 
succinate  oxidation,  malonate,  55 

Plants,  growth  of, 
malonate,  196-197 
mercurials,  965-968 

Plasma,    see    Blood 

Plasmodium  berghei,  glucose  utihzation, 
quinacrine,  560 

Plasmodium  gallinaceum,  succinate  accu- 
mulation, 
malonate,  93 

Plasmodium  lophurae,  respiration, 
quinacrine,  560 

Plasmodium  vivax, 

lactate  oxidation,  malonate,  78 
pyruvate  oxidation,  malonate,  74 


succinate  oxidation,  53 
Pleuropneumonia  virus,  infectivity  of, 

mercurials,  979 
Plumaria  elegans,  mercurial  toxicity,  967 
PM,  see  Phenylmercuric  ion 
Pneumococcus  polysaccharide,  lysozyme, 

459 
Pneumonitis  virus,  proliferation  of, 

D-cycloserine,  360 

malonate,   194 
Poliomyelitis  virus,  infectivity  of, 

mercurials,  979 
Pollen,  malonate  metabolism  in,  228 
Polyacrylate, 

phosphatase  (acid),  465 

trypsin,  457 
Polyadenylate,  polycytidylate  phosphory- 

lase,  463 
Polyalanine,   pepsin,   458 
Polyaspartate, 

pepsin,  458 

ribonuclease,  462 
Polyestradiol-phosphate,     phosphatase 

(acid),  464 
Polyethylenesulfonate,  phosphatase  (acid) 

464-465 
Polygalacturonase,  galacturonate,  421 
Polyglucose  (oxidized),  lysozyme,  459 
Polyglucose-sulfate, 

lipoprotein  lipase,  463 

lysozyme,  459 

ribonuclease,  462 
Polyglutamate, 

lysozyme,  459 

pepsin,  458 

ribonuclease,  462 

trypsin,  456-457 
Polyhydroquinone,    phosphatase    (acid), 

464-465 
Polylysine, 

lipoprotein  lipase,  463 

pepsin,  457-458 

trypsin,  456-457 
Polymeric  sulfonates, 

lysozyme,  459 

ulcer  reduction,  458 
Polymers,  see  Macroions 
Polymethacrylate,  hyaluronidase,  459 


SUBJECT    INDEX 


1213 


Polynucleotide  nucleotidyltransferase,  see 
Polynucleotide  phosphorylase 

Polynucleotide  phosphorylase, 
analogs,  474 
macroions,  463 

Polyornithine,  pepsin,  457-458 

Polyphenol  oxidase,  see  Catechol  oxidase, 
Phenol  oxidase,  and  Tyrosinase 

Polyphloretin-phosphate,    hyaluronidase, 
461 

Polyphloroglucinol-phosphate,  hyaluroni- 
dase, 461 

Polystyrenesulfonate,  hyaluronidase,  459 

Polyuridylate,  polyadenylate  phosphory- 
lase, 463 

Polyvinyl-sulfate,  ribonuclease,  463 

Polyxenyl-phosphate,  phosphatase  (acid), 
443,  464 

Porphobilinogen  deaminase,  analogs,  600 

Porphyra  perforata, 

K+-Na+  transport,  mercurials,  908-909 
respiration    (endogenous),    mercurials, 
881 

Porphyrexide, 

chemical  propeties,  664-666 
enzyme  inhibitions,  667-669 
oxidation  of  amino  acids,  666 
oxidation  of  proteins,  666-667 
oxidation  of  SH  groups,  666 
oxidation-reduction  potential,  665 
paramagnetism,   664-665 
spiro  analogs,  666 
stability,  665 

Porphyridium  cruentum,  respiration  (en- 
dogenous), 
malonate,  169 

Porphyrin,  biosynthesis  of, 
ferricyanide,  678 
malonate,  158-163 
mercurials,  888 
pathways  of,  159 

Porphyrindin, 

chemical  properties,   664-666 

determination  of  SH  groups,  666-667 

enzyme  inhibitions,  667-669 

heart,  669 

neuroblastic  damage,  670 

oxidation  of  proteins,  666-667 


oxidation  of  SH  groups,  666 

oxidation-reduction   potential,   665 

paramagnetism,  664-665 

Sarcoma  37,  670 

skin,  669-670 

spiro   analogs,   666 

stability,  665 

urease,  643 
Potassium, 

active  uptakes,  mercurials,  908-909 

barley  root  uptake,  malonate,  209-210 

erythrocytic  efflux,  mercurials,  903-905 

mitochondrial  uptake, 
o-iodosobenzoate,  722 
malonate,  209 

plasma  level,  malonate,  206 

renal  transport, 

dehydroacetate,  626 
mercurials,  921,  928,  936 

renal  uptake,  malonate,  205-206 

yeast  efflux,  mercurials,  898-900 
Potato, 

amino  acid  levels  in,  malonate,  106 

amino  acid  metabolism,  154 

carbohydrate   biosynthesis,    malonate, 

106 

citrate    formation    from    glucose,    ma- 
lonate, 105-106 

glucose  metabolism,  106,  HI,  132-133 

lipid  biosynthesis,  malonate,  149 

pentose-P  pathway,  malonate,  132 

respiration  (endogenous), 

y-hydroxy-a-ketoglutarate,  616 
malonate,  172,  182 

succinate  accumulation,  malonate,  91, 

106 

succinate  dehydrogenase,  19,  27 

sucrose  biosynthesis,  malonate,  132 
Potato  virus  X,  splitting  into  subunits  by 

mercurials,  980 
Proflavine, 

Lactobacillus  growth,  537 

structure  of,  537 
Prolidase  (imidodipeptidase),  mercurials, 

protection  by  Mn++,  783 
D-Proline,  Zl^-pyrroline-5-carboxylate  re- 
ductase, 355 
L-Proline, 


1214 


SUBJECT    INDEX 


L-amino  acid  oxidase,  340 
arginase,  337 

A  i-pyTroline-5-carboxylate   dehydroge- 
nase, 336,  355 

A  ^  -  pyrroline - 5 - carboxylate   reductase, 
355 

Proline  oxidase,  kojic  acid,  350 

Propane-tricarboxylate, 
aconitase,  240 

isocitrate  dehydrogenase,  240 
a-ketoglutarate  oxidation,   240 
pyruvate  oxidation,  240-241 
succinate  dehydrogenase,  240-241 

PropicilHn,  penicilHnase,  598 

Propionate, 

acetate  metabolism,  613-614 
acetyl-CoA  synthetase,  613 
carboxypeptidase,  366 
fatty  acid  synthesis  from  acetate,  613 
formation  from  lactate,  malonate,  165 
glutamate  decarboxylase,  328 
kynurenine:  a-ketoglutarate     transami- 
nase, 608 

lactate  dehydrogenase,  436 
lipid  metabolism,  613-614 
metabolism  of,  malonate,  144-145 
pantoate:/?-alanine  ligase,  598 
pyruvate  decarboxylase,  430 

Propionibacterium,  methylmalonate  in, 
224 

Propionibacterium   pentosaceum,   glycerol 
fermentation, 
malonate,  164 

Propionibecterium  shermanii,  methylma- 
lonyl-CoA  and  succinyl-CoA  intercon- 
version  in,  235 

Propionyl-CoA,  fatty  acid  biosynthesis, 
613-614 

Propionyl-CoA  carboxylase,  mercurials, 
853 

Prostate, 

malonate  metabolism  in,  228 
respiration  (endogenous),  malonate,  179 
succinate  oxidation,  malonate,  55 

Protease  (Aspergillus), 
hydrogen  peroxide,  693 
iodine,  687 

Protease  (Etroplus),  o-iodosobenzoate,  712 


Protease  (Rastrelliger),  o-iodosobenzoate, 

712 
Protease  (Trifolium),  hydrogen  peroxide, 

693 
Protection,  see  also  specific  enzymes  and 
inhibitors, 

against  o-iodosobenzoate,  717 
against  mercurials,  778-779,  783-784 
against  SH  reagents,  41-42,  650-651 
Proteinase  (Aspergillus), 
cystine,  662 
iodine,  687 
Proteinase  (Clostridium), 
iodine,  687 
mercurials,  793 
Proteinase  (lens), 
GSSG,  662 

o-iodosobenzoate,  712 
Proteinase  (mackerel),  ferri cyanide,  676 
Proteinase  (Pseudomonas),  permanganate, 

660 
Proteinase  (yeast),  mercurials,  791,  793 
Protein  disulfides  reductase, 
mercurials,  853,  926 
quinacrine,  554 
riboflavin,    543 
Proteins, 

biosynthesis  of, 
amethopterin,  585 
a-amino-/3-chlorobutyrate,  351 
jff-azaguanine,  478 
2-deoxyglucose,  399 
p-fluorophenylalanine,  351 
5-fluorouracil,  479 
folate  analogs,  585 
malonate,  155-157 
mercurials,  887-888 
chymotrypsin  inhibition,  457 
oxidation   of, 

ferricyanide,  671-672 
porphyrexide,  666-667 
porphyrindin,   666-667 
tetrathionate,   697-698 
reactions  with, 
iodine,  680-682 
o-iodosobenzoate,  703-704 
mercurials,  751-768 


SUBJECT    INDEX 


1215 


Proteus  morganii,  infections  by, 

malonate,  221-222 
Proteus  vulgaris, 

acetate  accumulation,  benzoate,  349 

citrate  oxidation,  malonate,  78,  86 

glucose  metabolism,  benzoate,  349 

growth  of,  mercurials,  972 

isocitrate  oxidation,  malonate,  79,  86 

pyruvate  metabolism,  benzoate,  349 

succinate  oxidation,  malonate,  52 
Protocatechuate,   see  3,5-Dihydroxyben- 

zoate 
Protoporphyrin,  biosynthesis  of, 

arsenite,  162 

2,4-dinitrophenol,  162 

fluoroacetate,    162 

malonate,  162 
Pseudomonas, 

fumarate  oxidation,  difluoromalonate, 

239 

malonate  inhibition,  effect  of  drying, 

187 

malonic    semialdehyde    occurrence    in, 

225 

succinate  oxidation,  difluoromalonate, 

239 
Pseudomonas   aeruginosa, 

enzyme  induction  in,  mercurials,  888 

growth  of,  dehydroacetate,  632 

malonate  metabolism  in,  228-229 

succinate  oxidation,  malonate,  52 
Pseudomonas  fluorescens, 

growth  of,  malonate,  195 

malonate  metabolism  in,  228,  230-231 
Pseudomonas  hydrophila,  pentose  oxida- 
tion, 

malonate,  132 
Pseudomonas  saccharophila, 

a-amylase  synthesis  in,   D-asparagine, 

269 

fumarate  oxidation,  malonate,  81 

pyruvate  oxidation,   malonate,   74 

succinate  oxidation,  malonate,  52 
Psicofuranine,   xanthosine-5'-P   aminase, 

476,  481 
9  -  D  -  Psicofuranosyl  -  6  -  aminopurine,    see 

Psicofuranine 
Psittacosis  agent. 


inactivation  of, 

o-iodosobenzoate,  728 
mercurials,   979-980 

proliferation  of, 
D-cycloserine,  360 
o-iodosobenzoate,  728 
malonate,  194 
mercurials,  981 
Psittacosis-lymphogranuloma    virus,    in- 

fectivity  of, 

mercurials,  979 
Pteridines, 

structures   of,   287 

xanthine  oxidase,  285-289 
Pteridylaldehyde,    see    Pterin-6-aldehyde 
Pterin,  structure  of,  287 
Pterin-6-aldehyde, 

glucose  oxidase,  288 

guanase,  288 

oxidation  by  xanthine  oxidase,  288 

3  -  phosphoglyceraldehyde     dehydroge- 
nase, 288 

potentiation   of  8-azaguanine  carcino- 

stasis,  288 

quinine  oxidase,  288 

structure  of,  287 

urate  levels  in  tissues  in  vivo,  288 

uricase,  288 

xanthine  oxidase,  285-289 
Pterin-6-carboxylate,    xanthine    oxidase, 

289 
Pteroate,  xanthine  oxidase,  289 
Pteroylaspartate,     dopa     decarboxylase, 

586 
Pteroyglutamate,  see  Folate 
pTpTpTpT,  phosphodiesterase,  473 
Puccinia  (stem  rust), 

germination  of  uredospores,  malonate, 

195-196 

leaf  infections  by,  oxythiamine,  529 

respiration  (endogenous),  malonate,  169 
Pullularia  pullulans, 

acetate  oxidation,  malonate,  77 

malonate  metabolism  in,  190,  228 

pyruvate  oxidation,  malonate,  74 
Pupillary  size,  thiamine  analogs,  532 
Purines, 

ionization  of,  280 


1216 


SUBJECT    INDEX 


riboflavin  complexes  of,  545 
Putrescine, 

kynureninase,  595 
structure   of,   361 
Pyrazoloisoguanine, 
structure  of,  280 
xanthine  oxidase,  281-282 
Pyridine, 

glucose  dehydrogenase,  501 
NAD  nucleosidase,  488 
Pyridine-2-carboxylate,    NAD    nucleosi- 
dase, 491 
Pyridine-2,6-dicarboxylate, 

diaminopimelate  decarboxylase,  593 
glutamate  dehydrogenase,  331 
Pyridine-3-suIfonamide, 
bacterial  growth,   504 
glucose  dehydrogenase,  500-502 
lactate  dehydrogenase,  500-502 
NAD  nucleosidase,  491,  504 
nicotinamide  deaminase,  504 
Pyridine-3-sulfonate, 

alcohol  dehydrogenase,  504 
bacterial  growth,  504 
glucose  dehydrogenase,  500-502,  504 
lactate  dehydrogenase,  500-502,  504 
NAD  nucleosidase,  491 
respiration  of  mycobacteria,  504 
sulfite  oxidase,  451 
toxicity,  504 
Pyridoxal, 
analogs  of, 

active  transport,  574-575 
bacterial  growth,  575-576 
blood  urea,  572-573 
carcinostasis,  576-577 
central  nervous  system,  573-574 
embryogenesis,  576 
enzyme  inhibitions,  564-566,  569-572 
fatty  acid  biosynthesis,  574 
GABA  level  in  brain,  573-574 
pyridoxal  kinase,  564-565 
pyridoxine  deficiency,  562,  577-578 
pyridoxine  levels  in  tissues,  566-569 
pyridoxine   metabolism,   564 
structures  of,  563 
toxicity,   573-574,   577-578 
tjrpes  of,  562 


glucose  dehydrogenase,  501-502 

metabolism  of,  pathways,  561-562 

pjrridoxamine-P  oxidase,   566 

pyridoxol-P  oxidase,   566 
Pyridoxal  azine,  pjTidoxal  kinase,  564 
Pyridoxal  kinase, 

aminooxyacetate,  358-359 

analogs,  564-565 

ATP  analogs,  465,  477 

toxopyrimidine,  578 
Pyridoxal  oxime, 

pyridoxamine-P  oxidase,  566 

pyridoxol-P  oxidase,   566 
PjTidoxal-phosphate,  metabolic  functions 

of,  561 
Pyridoxal   semicarbazone,   pyridoxal   ki- 
nase, 564 
Pyridoxamine, 

pyridoxamine-P  oxidase,  566 

pyridoxamine-P     oxidative     deamina- 

tion,  564 

pyridoxol-P  oxidase,   566 
Pyridoxamineroxalacetate    transaminase, 

see  Transaminases 
Pyridoxamine-phosphate,    oxidative    de- 

amination  of, 

pyridoxamine,  564 
Pyridoxamine-phosphate    oxidase,    ana- 
logs, 566 
4-Pyridoxate, 

glucose  dehydrogenase,  501-502 

pyridoxamine-P  oxidase,  566 

pyridoxol-P  oxidase,  566 
4-Pyridoxate-phosphate, 

pyridoxamine-P  oxidase,  566 

pyridoxol-P  oxidase,  566 
Pyridoxol, 

oxidation  of,  analogs,  564 

pyridoxamine-P  oxidase,  566 

pyridoxol-P  oxidase,   566 
Pyridoxol  dehydrogenase,  mercurials,  817 
Pyridoxol-phosphate     oxidase,     analogs, 

564,  566 
Pyridoxyl  -  L  -  alanine,     alanine :  pyruvate 

transaminase,  569 
2-Pyridylalanine,    L-phenylalaninersRNA 

ligase  (AMP),  355 
4(5)-3'-Pyridylglyoxaline, 


SUBJECT    INDEX 


1217 


NAD  nucleosidase,  489,  491 

structure  of,  488 
PjTimethamine  (Daraprim), 

folate  deficiency,  584 

folinate  formation,  584 

uptake  by  bacteria,  584 
Pyrimidines,  metabolism  of, 

fluoropyrimidines,  478-481 
Pyrithiaminase,   528-529 
Pyrithiamine    (neopyrithiamine), 

bacterial  growth,  528-530 

bacterial  resistance  to,  528-529 

blood  pyruvate,  520,  527 

central  nervous  system,  527,  530 

fungal  growth,  516,  528-529 

glutamate  metabolism  in  seeds,  522 

heart  rate  in  vivo,  527 

miosis,  532 

nerve  membrane  potentials,  531 

neuromuscular  block,  531-532 

phosphorylation  of,  519,  527 

pyruvate  decarboxylase  in  vivo,  531 

pyruvate  dismutation  in  vivo,  521 

pyruvate  oxidation  in  vivo,  520-521 

rat  growth,  527 

respiratory  quotient  in  rats,  521-522 

structure  of,  517 

thiaminase,  523-524 

thiamine  deficiency,  516,  530-532 

thiamine  kinase,  522-523 

thiamine  levels  in  tissues,  525-527 

thiamine-PP  levels  in  tissues,  521 

tissue  levels  of,  527-528 

toxicity,  530-531 

Vibrio  growth,  530 
Pyrithiamine-diphosphate, 

pyruvate  decarboxylase,  518 

pjTuvate  oxidase,  519 
Pyrocatechase,  mercurials,  853 
PjTOgallol, 

adrenergic  potentiation,  611 

D-amino  acid  oxidase,  344 

blood  pressure,  611 

catecholamine  metabolism,  611-612 

catechol-0-methyl  transferase,  595, 611- 

612 

dehydroshikimate  reductase,  593 

epinephrine  responses,  611 


histidine  decarboxylase,  352 
norepinephrine  level  in  brain,  611 
urinary  catecholamines,  611-612 
Pyrophosphatase, 

o-iodosobenzoate,  712 
malonate,  64 

mercurials,  relation  to  SH  groups,  802 
nucleotides,  475 
phosphate,  439 
P>Tophosphate, 

glutamate  semialdehyde  reductase,  507 
ionization  constants,  242 
NAD  nucleosidase,  492 
oxidative  phosphorylation,  448 
phosphatases,  439 
pyridoxal  kinase,  477 
succinate  dehydrogenase,  243 
tyrosinase,  301 
Pyrophosphite,    oxidative    phosphoryla- 
tion, 448 
Pyrrole-2-carboxylate,  D-amino  acid  oxi- 
dase, 342,  346 
A 1  -  Pyrroline  -  5  -  carboxylate    dehydroge- 
nase, analogs,  336,  355 
A 1  -  Pyrroline  -  5  -  carboxylate     reductase, 

analogs,  355 
a-Pyrrone-5-carboxylate,   see   Coumalate 
Pyruvate, 

accumulation  of,  ferricyanide,  677 
D-amino  acid  oxidase,  340 
blood  levels  of, 

malonate,  219 

thiamine  analogs,  520,  527 
decarboxylation  of, 

2-deoxyglucose,  396 

lipoate  analogs,  590 
dismutation  of,  pyrithiamine,  521 
glycerate  dehydrogenase,   430 
glyoxylate  reductase,  438 
lactate  dehydrogenase,  437 
malate  dehydrogenase,  596 
oxidation  of, 

acetylene-dicarboxylate,  240-241 

analogs,  429-432 

benzoate,  349 

trans  -  cyclopentane  -1,2-  dicarboxy- 

late,  241 

2-deoxyglucose,  391-392,  397,  399 


1218 


SUBJECT    INDEX 


D-glucosamine,  383 

lipoate  analogs,  590 

malonate,  69,  74-77,  128,  135 

mercurials,  878-879 

oxamate,  434 

oxygen,  658-659 

propane-tricarboxylate,  240-241 

pyrithiamine  in  vivo,  520-521 

quinacrine,  560 

thiamine  analogs,  519 
phosphatase  (acid),  442 
tartronate  semialdehyde  reductase,  602 
Pyruvate   carboxylase   (pyruvateiCOa  li- 


ferricyanide,  676 

mercurials,  853 
PyruvaterCOj  ligase,  see  Pyruvate  carbo- 
xylase 
Pyruvate  decarboxylase, 

acetaldehyde,  432,  600 

acetamide,  430 

iodine,  682,  687 

o-iodosobenzoate,  712 

mercurials,  775,  783,  810,  853 
mutual  depletion  behavior,  775 
protection  by  pyruvate,  783 
rate  of  inhibition,  810 

oxythiamine-PP,  518 

porphyrexide,  668 

porphyrindin,  668 

pyrithiamine  in  vivo,  521 

pyrithiamine-PP,  518 

pyruvate  analogs,  600 

thiamine  analogs,  516,  518,  521 
Pyruvate  dehydrogenase,  mercurials,  854 
Pyruvate  kinase, 

6-aminonicotinamide-NAD,  505 

malonate,  64,  129 

mercurials,  825,  854 
Pyruvate  oxidase, 

ferricyanide,  676 

malonate,  64 

mercurials,  751,  774-775,  783,  854 
mutual  depletion  behavior,  775 
protection  by  thiamine-PP,  783 
type  of  inhibition,  774 

mercuric  ion,  antagonism   by  hpoate, 

751 


oxygen  inactivation  of,  659 
phenylpyruvate,  430 
thiamine  analogs,  519 
Pyruvic  ethyl  ester,  pyruvate  decarboxy- 
lase, 430 


Quinacrine    (Atabrine,     Atebrin,    Mepa- 

crine) 

bacterial  growth,  546 

binding  to  cell  components,  546 

competition  with  FAD  or  FMN,  546- 

547,  556 

distribution  in  tissues,  546 

enzyme  inhibitions,  546-559 

flavin  complexes  with,  556 

germination   of  spores,   546 

glucose  utilization,  560 

nucleotide  complexes  with,  556 

oxidative  phosphorylation,  556-557 

pH  effects,  557-558 

photophosphorylation,  557 

respiration,  559-560 

structure  of,  546 

toxicity,  546 

use  to  detect  flavin  components,  559 
Quinine,  D-amino  acid  oxidase,  557 
Quinine  oxidase,  pterin-6-aldehyde,   288 
Quinolinate,  NAD  nucleosidase,  488 
Quinolines,   chymotrypsin,   373 
Quinone   reductase,   see   NADH:quinone 

oxidoreductase 

R 

Rafiinose, 

galactosidases,  418 

a-glucosidase,  416 
Rainbow  trout,  fatty  acid  oxidation, 

malonate,  137 
Reactivation,  see  Reversal 
Relative  /IF  values,  calculation  of,  254- 

255 
Renilla  reniformis  (sea   pansy),  lumines- 
cence, 

mercurials,  891 
Resistance, 

to  mercurials,  983-985 

to  thiamine  analogs,  528-529 


SUBJECT    INDEX 


1219 


Resorcinol, 

catechol  oxidase,  296-298 

respiration  (endogenous)  of  apple  skin, 

296 

tyrosinase,  304 
Resorcinol  monobenzoate,  tyrosinase,  304 
Respiration, 

trans -aconitate,  273-274 

kojic  acid,  350 

mercurials,  879-886,  984 

quinacrine,  559-560 
Respiration   (acetate),   see   also   Acetate, 

oxidation   of, 

dehydroacetate,  626 
Respiration  (endogenous),  see  also  speci- 
fic inhibitors, 

6-aminonicotinamide,  504 

benzoate,  348 

dehydroacetate,  623-624 

2-deoxyglucose,  391-392,  396-397 

y-hydroxy-a-ketoglutarate,  616 

o-iodosobenzoate,  721 

malonate,  166-186 

nicotinamide,  503 

p-nitrophenol,  297 

resorcinol,  296 
Respiration   (galactose),   see  also   Galac- 
tose, oxidation  of, 

deoxygalactose,  391-392 
Respiration   (glucose),   see   also   Glucose, 

oxidation  of, 

l,5-anhydro-D-glucitol-6-P,  379 

2-deoxyglucose,  391-394 

ferricyanide,  677 

hydrogen  peroxide,  695 

o-iodosobenzoate,  722 

malonate,  123-125,   127,   133-135 

mercurials,  885,  889-891,  893-894,  898, 

927-928,   948 

nicotinamide,  500 
Respiratory  quotient  (R.  Q.), 

dependence  on  substrate,  184-185 

malonate,    184-185 

pyrithiamine,  521-522 
Reticulocytes, 

glucose  utilization,  quinacrine,  560 

iron  incorporation  into  heme,  malonate, 
163 


leucine  incorporation  into  protein,  mer- 
curials, 887 

protein  biosynthesis,  mercurials,  887 
respiration  (endogenous),  malonate,  179 
succinate  oxidation,  malonate,  55 
valine   incorporation   into   protein,    a- 
amino-^-chlorobutyrate,  351 

Retina, 

glutamate  uptake,  malonate,  153 
a-ketoglutarate  oxidation,  malonate,  81 
K+  uptake,  malonate,  153 
respiration  (endogenous),  malonate,  176 
183 

Retinene  oxidase,  quinacrine,  554 

Reversal, 

of  enzyme  inhibitions,  analysis  of,  821- 
823 

of  o-iodosobenzoate  inhibitions,  718 
of  mercurial  inhibitions,  different  meth- 
ods for,  825 
of  SH  reagent  inhibitions,  650-651 

Rhamnose,  a-mannosidase,  422 

Rhizobium  japonicwm, 

succinate  dehydrogenase,  malonate  K^, 

33 

succinate  oxidation,  malonate,  53 

Rhizoctonia  solani, 

growth  of,  mercurials,  973 
sucrose  uptake,  mercurials,  911 

Rhizopus  nigricans,  growth  of, 
dehydroacetate,  632 

Rhodanese,  see  Thiosulfate  transulfurase 

Rhodopseudomonas  spheroides,  porphyrin 
biosynthesis, 
malonate,  162-163 
mercurials,  888 

Rhodospir ilium    rubrum, 

acetate  oxidation,  malonate,  77 
fumarate  oxidation,  malonate,  81 
hydrogen  evolution,  nitrogen,  293 
lactate  oxidation,  malonate,  78 
photophosphorylation, 
malonate,  163 
mercurials,  892 
propionate  oxidation,  malonate,  146 
respiration  (endogenous),  malonate,  168 
succinate  oxidation,  malonate,  53 


1220 


SUBJECT    INDEX 


Rhodotorula  gracilis,  cycle  substrate  oxi- 
dations, 

malonate,  53,  74,  77,  79 
Rhubarb  leaves, 

respiration  (endogenous),  malonate,  171 
respiratory  quotient,  malonate,  185 
Ribityllumazine,  riboflavin  transglucosi- 

dase,  453 
Riboflavin, 

D-amino  acid  oxidase,  540-541 
L-amino  acid  oxidase,  540-542 
analogs  of, 

carcinostasis,  538 

enzyme  inhibitions,  540-545 

FAD  levels  in  tissues,  539-540 

flavoenzymes,  540-545 

growth  of  microorganisms,   537-538 

metabolism  of,  539-540 

molecular  complexes  between,  544 

quinacrine,  545-561 

riboflavin  metabolism,  539-540 

structures  of,  535-537 

types  of,  535 
biosynthesis  of,  analogs,  539 
FAD  pyrophosphorylase,  542 
galactono-y-lactone    dehydrogenase, 
540,  542 

D-gluconate  oxidase,   542 
glutamate  recemase,  542,  544 
D-lactate    oxidase,    542 
L-lactate  oxidase,  542-543 
NADPH :  methemoglobin     oxidoreduc- 
tase,  543 

protein  disulfide  reductase,  543 
purine  complexes  of,  545 
succinate  oxidase,  543 
Riboflavin  kinase,  see  Flavokinase 
Riboflavin-5 '-sulfate, 

D-amino  acid  oxidase,  540-541,  544 
NADPH  diaphorase,  544 
Riboflavin  synthetase,  analogs,  543 
Riboflavin  transglucosidase, 
analogs,  543 
quinacrine,  554 
Ribonuclease, 

configurational    changes,    disulfide   re- 
duction, 650 
DNA,  462 


hydrogen  peroxide,  693 

o-iodosobenzoate,  712,  715-716 
kinetics,  715-716 
pH  effects,  716 

macroions,  461-463 

mercurials,    750,    791,    793,    815,    817- 

818,  820,  860 

pH  effects,  791,  793 
reaction  with  substrate,  815 
stimulation,  815,  817-818,  820 

p-mercuribenzoate,  splitting  of  disulfide 

bonds,  750 

nucleotides,  475 
Ribonuclease  T2,  mercurials,  817 
Ribonucleates,  see  also  Nucleic  acids, 

biosynthesis  of,  mercurials,  969 

deoxyribonuclease,  462 

glycolysis  (anaerobic),  414,  465 

polycytidylate  phosphorylase,  463 
5'-Ribonucleotide  phosphohydrolase,  see 

5'-Nucleotidase 
Ribose, 

a-mannosidase,  422 

metabolism  of,  malonate,  132 

NAD  nucleosidase,  492 

phosphoarabinose  isomerase,  411 
Ribose  isomerase,  mercurials,  854 
Ribose-3-phosphate,  phosphopentose  iso- 
merase, 411 
Ribose-5-phosphate, 

D-amino  acid  oxidase,  545 

glucose  dehydrogenase,  410 

hexokinase,  379 

phosphoarabinose  isomerase,  411 
Ribose-6-phosphate,     phosphodeoxyribo- 

mutase,  413 
Ribose-phosphate  isomerase, 

cystine,  662 

mercurials,  854-855 
Ribulose-phosphate    carboxylase,    phos- 
phate, 412 
Ribulose-5-phosphate  kinase,  mercurials, 

855 
Rigor,  see  Contracture 
RNA,  see  Ribonucleates 
RNAase,  see  Ribonuclease 
RNA  nucleotidyltransferase,  quinacrine, 

555 


SUBJECT   INDEX 


1221 


Rose  petals,  respiration  (endogenous), 

malonate,   173,   182 
Rubidium,  uptake  by  roots, 

mercurials,  909 
Runner  bean,  see  Phaseolus  coccineus 


Saccharo-l,4-lactone,  see  Glucaro-l,4-lac- 

tone 
Saccharomyces  cerevisiae,  see  Yeast 
Sake,  malonate  occurrence  in,  224 
Salicylamide, 

glucose  dehydrogenase,  501 

lactate  dehydrogenase,  501 

sulfanilamide  acetylase,  601 
Salicylate, 

D-amino  acid  oxidase,  348 

dehydroshikimate  reductase,  606 

glucose  dehydrogenase,  501 

lactate  dehydrogenase,  501 

oxidative  phosphorylation,  348 

tricarboxylate  cycle,  348 

tyrosine:a-ketoglutarate  transaminase, 

306 
Salicyloyl-^-alanide, 

choline  acetylase,  587 

structure  of,  587 
Salmonella  enteritidis,  infection  by, 

malonate  on  antibacterial  activity  of 

blood,    223-224 
Salmonella  paratyphi,  growth  of, 

mercurials,  972 
Salmonella  pullorum, 

growth  of, 

dehydroacetate,  632-633 
mercurials,  972 

infection  by,  malonate,  224 

resistance  to  mercurials,  983,  983 
Salmonelle  schotmillleri,  growth  of, 

mercurials,  972 
Sahnonella  typhosa, 

growth  of, 

dehydroacetate,  632-633 
mercurials,  972 

resistance  to  mercurials,  983,  985 
Salmonella  typhimurium, 

infection  by,  malonate,  221-223 


malonate  metabolism  in,  228 
Salyrgan,   see  Mersalyl 
Samia  cecropia,  succinate  dehydrogenase, 

malonate,  29 
Sarcina  lutea,  respiration  (glucose) 

malonate,  124 
Sarcoma,  ATP  levels  in, 

aminopterin,   585 
Sarcoma  (mouse),  growth  of, 

2-deoxyglucose,  400 
Sarcoma  (rat),  growth  of, 

thiophene-2,5-dicarboxylate,  415 
Sarcoma  37, 

ADP-ATP    levels    in,    2-deoxyglucose, 

395 

blebbing  of,  malonate,  201 

porphyrindin,  670 
Sarcosine  oxidase,   methoxyacetate,   601 
Saxostrea  commercialis  (oyster)  eggs, 

respiration  (endogenous),  malonate,  174 
Saxostrea  commercialis  muscle, 

respiration  (endogenous),  malonate,  174 

succinate  dehydrogenase,  malonate,  28 
Scallop  muscle,  glycolysis, 

o-iodosobenzoate,    721 

mercurials,  876 
Scarlet  fever  toxin,  porphyrindin  inacti- 

vation  of,  667 
Scenedesmus  obliquus, 

glucose  uptake, 

D-glucosamine,  382 
a-L-sorbose-1-P,  379 

photosynthesis,  mercurials,  892 

respiration    (endogenous),    mercurials, 

881 
Schistocera  gregaria  fat  body, 

acetate  oxidation,  malonate,  77 

succinate  oxidation,  malonate,  54 
Schizophyllum  commune, 

citrate  accumulation,  malonate,  104 

cycle  intermediate  oxidations,  78-81 

pyruvate  oxidation,  malonate,  74 

respiration  (glucose),  mercurials,  881 

succinate  oxidation,  malonate,  53 
Sclerotinia  fructicola, 

growth  of,  mercurials,  973 

resistance  to  mercurials,  983 
Sea  hare,  see  Aplysia 


1222 


SUBJECT    INDEX 


Sea  pansy,  see  Renilla  reniformis 

Sea   urchin,   see   also   Arbacia,   Echinus, 

Lytechinus,  Strongylocentrotus,  and  Trip- 

neustes 
Sea  urchin  eggs, 

cleavage  and  development, 
2-deoxyglucose,  400 
o-iodosobenzoate,  726-727 
malonate,  117 

glycolysis,  mercurials,  875 

respiration     (endogenous),     malonate, 

174-175 

respiration    (glucose),    2-deoxyglucose, 

391 
Sea  urchin  spermatozoa,  respiration  (en- 
dogenous), 

o-iodosobenzoate,  721-722 
Sebacate,     kynurenine  :  a  -  ketoglutarate 

transaminase,  608-609 
Sedoheptulose- 1 ,7-diphosphate,  2-keto-3- 

deoxy  -  d  -  arabo  -  heptonate-7-P  synthe- 
tase, 413 
Sedoheptulose-7-phosphate,    2-keto-3-de- 

oxy-D-ara6o-heptonate-7-P  synthetase, 

413 
Semicarbazide,  diamine  oxidase,  362 
Seminal  vesicle, 

respiration  (endogenous),  malonate,  176 

succinate  dehydrogenase,  malonate,  31 
Sequential  inhibition, 

5-azaorotate  and  6-azauridine,  480 

malonate  and  fluoroacetate,  112 
Serine, 

biosynthesis  of, 
aminopterin,  585 
2-deaminofolate,  585 
deoxypyridoxol,  570-571 

dipeptidase,  367 

0-phosphoserine  phosphatase,  270 
D-Serine,  L-alanine  dehydrogenase,  354 
L- Serine, 

L-amino  acid  oxidase,  340 

arginase,  337 

phosphatase  (acid),  441 

L-threonine  dehydrase,  357 
Serine  deaminase, 

analogs,  357 

o-iodosobenzoate,  712 


L-Serine  hydro-lyase,  see  Tryptophan  syn- 
thetase 
Serotonin  (5-hydroxytryptamine), 

brain    levels    of,    a-methyl-m-tyrosine, 

316 

metabolism  of, 

deoxypyridoxol,  574 
galactoflavin,  544 

tissue  levels  of, 

a-methyldopa,   315-320 
phenylalanine,  325 

tryptophan  pyrrolase,  324-325 
Serratia  marcescens, 

fatty  acid  oxidation,  malonate,  137 

succinate  oxidation,  malonate,  52 
SH  groups,  see  Sulfhydryl  groups 
Shigella,  succinate  accumulation, 

malonate,  90 
Shikimate  dehydrogenase,  benzoate,  349 
SH  reagents,  see  Sulfhydryl  reagents 
Silicate,   phosphatases,  439-440 
Silkworm,  see  also  Cecropia  and  Samia, 

larvae  of,  aminomalonate  decarboxyla- 
tion in,  239 
Silver 

/S-glucuronidase,  795 

glutamate  dehydrogenase,  863 
Sinapate,   peroxidase,   599 
Skatole, 

chymotrypsin,  374 

histamine  release,  374 
Skin, 

electrical  potential  of, 

iodine,  690 

mercurials,  950 

porphyrindin,  669-670 

mercurial   levels  in,   930 

mercurial  vesication,   950 

mitosis  in,  mercurials,  968 

permeability  to  Cl~,  mercurials,  912 

pyruvate     oxidation,     2-deoxyglucose, 

391-392 

respiration  (endogenous), 
mercurials,  882,  912 

respiration  (fructose),   2-deoxyglucose, 

391-392 

respiration  (galactose),  2-deoxyglucose, 

391-392 


SUBJECT    INDEX 


1223 


respiration    (glucose),    2-deoxyglucose, 

391-392 

respiration  (mannose),  2-deocyglucose, 

391-392 
Smooth  muscle,  see  specific  muscles 
Snail,  see,  Helix 
Sodium, 

erythrocyte    transport    of,    malonate, 

209 

intestinal  transport  of,  mercurials,  916 

phosphotransacetylase,  452 

plasma  levels  of,  malonate,  206 

renal  transport  of, 
dehydroacetate,  626 
malonate,  206 
mercurials,  918-920 

urinary  excretion  of,  malonate,  206 
Solanain,  iodine,  687 
Sorbitol,  mutarotase,  413-414 
Sorbitol  dehydrogenase, 

o-iodosobenzoate,  712 

mercurials  in  vivo,  926 
Sorbitylflavin,  flavokinase,  539 
a  -  L  -  Sorbopyranose  - 1  -  phosphate,  struc- 
ture of,  377 
a-L-Sorbose, 

formation  from  glyceraldehyde,  377 

hexokinase,  377 
L-Sorbose-l,6-diphosphate,  aldolase,  407 
L-Sorbose- 1  -phosphate, 

aldolase,  407 

glucose  uptake  by  Scenedesmus,  379 

hexokinase,    377-379 

structure    of,    378 
L-Sorbose-6-phosphate,    hexokinase,    377 
Soybean  nodules, 

hydrogen  evolution,  nitrogen,  293 

nitrogen  fixation,  analogs,  292,  294 
Spermatozoa, 

acetate  oxidation,  malonate,  77 

glycolysis, 

malonate,  128 
mercurials,  884 
tetrathionate,  699 

motility  of, 
malonate,  203 
mercurials,  964 
tetrathionate,  699 


pyruvate  oxidation,  malonate,  77,  87 
respiration  (endogenous), 
o-iodosobenzoate,  721-722 
malonate,  176 
mercurials,  882,  884 
respiration  (glucose),  malonate,  124 

Spermine,  structure  of,  361 

Spinach  chloroplasts, 

COj  photochemical  fixation,  mercurials, 

892 

fatty  acid  biosynthesis  from  malonate, 

149 

Hill  reaction,   mercurials,  891 

malonate  incorporation  into,  232 

NADP  photoreduction,  mercurials,  891 

photophosphorylation,  mercurials,  892 

protein  biosynthesis,  mercurials,  887 

Spinach  levaes, 

respiration  (endogenous), 
malonate,   170-171,   181 
p-nitrophenol,   297 
succinate  accumulation,  malonate,  94 

Spindle  formation,  see  also  Cell  division 
mercurials,  969 

Spiroporphyrexide,  666 

Spirophorphyrindin,  666 

Spleen, 

amino  acid  accumulation,  malonate,  103 
ATP  levels  in,  aminopterin,  585 
citrate  levels  in,  sequential  inhibition 
by  malonate  and  fluoroacetate,  112 
glycolysis,  mercurials,  875 
malonate  decarboxylation  in,  232 
malonate  levels  in  vivo,  102 
mercurial  levels  in  vivo,  930,  959-960 
succinate  accumulation,  malonate,  102 
succinate  levels  in  vivo,  malonate,  102 

Spore  germination,  see  Germination 

Squalene,  formation  from  mevalonate, 
mercurials,  886 

Staphylococcal    phage,    inactivation    by 
mercurials,  976,  979-980 

Staphylococcus  aureus, 

acetate  oxidation,  malonate,  77 
growth   of, 

dehydroacetate,  632-633 
dichlororiboflavin  analog,  537 
mercurials,  972-973,  975 


1224 


SUBJECT    INDEX 


infection  by,  malonate,  221-222 

o-iodosobenzoate  killing  of,  727 

phosphate  transport, 
analog  ions,  267 
iodine,  690 
mercurials,  910,  912-913 

pyruvate  oxidation,  malonate,  74 

resistance, 

to  D- cycloserine,  359 
to  mercurials,  983-984 
to  pyrithiamine,  529 

respiration  (glucose),  mercurials,  880 
Stearyl-CoA,  acetyl-CoA  carboxylase,  614 
Stemphyllium   sarcinaeforme,   growth   of, 

mercurials,  973 
Zl^-Steroid    dehydrogenase,    o-iodosoben- 
zoate, 712 
zl*-5a-Steroid    dehydrogenase,    o-iodoso- 
benzoate, 713 
Steroid  hydroxylase, 

o-iodosobenzoate,  713 

quinacrine,  555 
Sterol    ester    hydrolase,    see    Cholesterol 

esterase 
Sterols,  see  also  Cholesterol, 

biosynthesis  of, 
malonate,  149-150 
mercurials,  886-887 
Stomach,  see  Gastric  acid  secretion  and 

Gastric  mucose 
Strain  L  cells,  galactose  uptake, 

2-deoxyglucose,  394 
Streptococcus  faecalis, 

acetoin  formation  from  pyruvate,  phe- 

nylpyruvate,  430 

folate  metabolism,  analogs,  582 

growth  of,  pyridoxal  analogs,  575 

pyruvate  oxidation,  mercuric  ion,  751 
Streptococcus  hemolyticus,  growth  of, 

mercurials,  973 
Streptococcus  plantarum,  growth  of, 

dichlororiboflavin  analog,  537 
Streptococcus  pyogenes,  infection  by, 

malonate,  221 
Streptomyces  coelicolor, 

a-ketoglutarate    oxidation,    malonate, 

79,  84 

succinate  levels  in,  malonate,  97 


Streptomyces  olivaceus, 

malonate  metabolism  in,  228 
succinate  oxidation,  malonate,  53 

Strigonionas  oncopelti,  respiration  (endo- 
genous), 
malonate,  173 

Strong ylocentrotus  purpuratus  eggs, 
development  of,  malonate,   198-199 
respiration      (endogenous),     malonate, 
175-176,   181 

Stylonychia  pusulata,  succinate  dehydro- 
genase, 
malonate,  28 

Suberate, 

aspartate  :  a  -  ketoglutarate      transami- 
nase, 334 

glutamate  decarboxylase,  328 
kynurenine:  a-ketoglutarate     transami- 
nase, 608 

Substitution  analog,  definition  of,  257 

Succinate, 

accumulation  of,  malonate,  70-71,  90- 
104 

aspartate :  a  -  ketoglutarate      transami- 
nase, 334 

citrate  sjnithetase,  597 
fumarase,  275,  277 
glutamate  decarboxylase,  328 
glutamate  dehydrogenase,  330,  332 
intercharge  distance,  5-6 
ionic  species,  pH  effect,  10 
ionization  of,  8 

kynurenine:a-ketoglutarate     transami- 
nase, 608 

D -lactate  dehydrogenase,  437 
oxidation  of, 
benzoate,  348 
malonate,  50-58 
mercurials,  879 
phosphofructokinase,  385 
urinary  excretion  of,   dehydroacetate, 
628 

Succinate  dehydrogenase, 

acetylene-dicarboxylate,  240-241 
active  center  of,  18,  40-42,  44-45 
arsonoacetate,  243 
1,4-butane-disphophonate,  243 
caffeate,  314 


SUBJECT    INDEX 


1225 


<raw5-cyclopentane-l,2-dicarboxylate, 

241 

cysteine,  659 

dehydroacetate,  620-622 

dicarboxylate  ions,  34-40 

difluoromalonate,  239 

ethane- 1,2-disulfonate,   242-243 

ferricyanide,  676 

fluoromalonate,  239 

GSSG,  661-662 

hydrogen  peroxide,  693 

3-hydroxycinnamate,  314 

hypophosphate,  243 

intercharge  distance  in  active  site,  42- 

44 

iodine,  682,  687 

o-iodosobenzoate,   713,   715-718 

itaconate,  601 

malonate, 

absorption  spectrum  changes,  18-19 

activation  of  enzyme,  45-46 

binding  energy,  42 

competitive  nature,  21-25 

dependence  on  electron  acceptor,  18- 

20 

discovery  of  inhibition,  2 

effect  of  ATP,  48 

effect  of  Ca++,  46-48 

effect  of  liquid  nitrogen,  187 

effect  of  Mg++,  48 

effect  of  osmolarity,  46-47 

effect  of  succinate  concentration,  25 

effect  of  temperature,  46 

interaction  distance,  42-43 

KJKi  ratio,  33 

reversibility,  24-25 

species  variation,  49-50 

malondialdehyde,  40-41 

malondiamide,  41 

malonic  diethyl  ester,  236 

mercurials,  773,  779,  783-784,  787-788, 

802,  810,  825,  855-856,  870-872, 925-926 
acceleration  of  nonheme  Fe  chela- 
tion, 788 

in  vivo  inhibition,  925-926 
loss  of  nonheme  Fe,  787 
protection  by  XAD,  783 
protection  by  oxalacetate,  779 


protection  by  succinate,  783-784 

rate  of  inhibition,  810 

relation  to  SH  groups,  802 

reversal   by  different  methods,   825 

type  of  inhibition,  773 
methane-diphosphonate,   243 
methionate,  243 

methods  for  measuring  activity  of,  18- 
20 

3-nitropropionate,  244 
oxygen  inactivation  of,  659 
phosphonoacetate,  243 
^-phosphonopropionate,  242-243 
porphyrindin,  668 
propane-tricarboxylate,  240-241 
pyrophosphate,  243 
quinacrine,  555 
succinyl  peroxide,  694 
o-sulfobenzoate,  243 
/S-sulfopropionate,  242-243 
tetrathionate,  696,  700 
Succinate:  methylmalonate  isomerase,  ma- 
lonate, 64 
Succinate  oxidase, 
composition  of,  16-17 
dehydroacetate,  620-622 
FAD,  543 

fiavinmononucleotide,  543 
y-hydroxy-a-ketoglutarate,  616 
o-iodosobenzoate,  715-718 

kinetics,  715 

protection  by  succinate,  717 

reversal  by  thiols,  718 

temperature  effects,  715 
kojic  acid,  350 

malonate,  localization  of  site  of  action, 
18,  20-21 

mercurials,  817,  826,  870-872 
A"-piperidinomethylnicotinamide,  503 
quinacrine,  547,  555,  560 
riboflavin,  543 
riboflavin  analogs,  543 
Succinate   semialdehyde   dehydrogenase, 
j3-hydroxybenzaldehyde,  601 
mercurials,  856 
Succinyl-CoA  deacylase,  ATP,  475 
Succinyl-/S-ketoacyl-CoA    transferase, 
transfer  of  Co  A  to  malonate,  234 


1226 


SUBJECT    INDEX 


Succinyl  peroxide,  enzyme  inhibitions  by, 

694 
Sucrose, 

a-amylase,  420 

/S-amylase,  421 

a-dextran-l,6-glucosidase,  417 

a-glucosidase,  416-417,  423 

a-mannosidase,  422 

uptake  by  Fhizoctonia,  mercurials,  911 
Sucrose  transfructosylase,  analogs,  421 
Sugar  cane,  glucose  uptake, 

mercurials,  910-911 
Sugar  pine  seedlings,  see  also  Pinus  lam- 

bertiana, 

tricarboxylate    cycle    oxidations,    ma- 

lonate,  74,  80,  82 
Sulfamate,  phosphatase  (acid),  441 
Sulfanilamide,  acetylation  of, 

mercurials,  927 
Sulfanilamide  acetylase, 

analogs,  601 

mercurials,  750 
Sulfanilate,  D-amino  acid  oxidase,  343 
Sulfatases, 

analogs,  443-444 

mercurials,  860 
Sulfate, 

aldolase,  414 

arylsulfatases,  443-444 

choline  sulfatase,   444 

creatine  kinase,  446 

glycolysis  (anaerobic),  414 

thiosulfate  utilization  by  Neurospora, 

451-452 

tyrosinase,  301 

uptake  by  yeast,  thiosulfate,  267 
Sulfated  hyaluronate,  hyaluronidase,  459 
Sulfated  pectate,  hyaluronidase,  459 
Sulfated  polysaccharides, 

lipoprotein  lipase,  463 

ribonuclease,  462-463 
Sulfenyl  iodine  groups,  formation  by  io- 
dine, 680-681 
Sulfhydryl  enzymes, 

definition  of,  635,  647 

oxygen  inactivation  of,  659 

reactivity  of  SH  groups,  factors  deter- 
mining, 643-647 


Sulfhydryl  groups, 

addition  to  double  bonds,  643 

alkylation  of,  642 

bond  characteristics,  639 

chemical  properties,  637-642 

determination   of,    640-642,    667,    669, 

671-672,    680-681,    697-698,    702-704, 

752-757,  762-768,  798-809 

hydrogen  bonding,  640 

ionization  of,  638 

oxidation  of,  642,  655-700 

oxidation-reduction  potentials,  656 

reactivity  of,  factors  determining,  643- 

647 

role  in  cellular  function,  637 

role  in  metabolism,  636-637 
Sulfhydryl  reagents,  635-653 

configurational    changes    in    enzymes, 

649-650 

enzyme  inhibitions, 

interpretation  of,  647-650 
mechanisms  of,  647-648 
protection  against,  650-651 
reversal  of,  650-651 

reaction  with  disulfides,  640 

reaction  with  SH  groups,  pH  effects, 

638 

specificity  of,  652-653 

types  of  reactions  with  enzymes,  642- 

643 

uses    of,   general    considerations,    651- 

653 
Sulfite, 

arylsulfatase,  443-444 

choline  sulfatase,  444 

oxidation  of,  thiosulfate,  451 
Sulfite  oxidase, 

analogs,  451 

quinacrine,  555 
Sulfoacetate,  aspartase,  355 
o-Sulfobenzoate, 

intercharge  distance,   7 

succinate  dehydrogenase,  243 
Sulfocholine,   thetin:homocysteine  trans- 

methylase,  356 
Sulfoglycolate, 

glycolate  oxidase,  438 

structure  of,  438 


SUBJECT    INDEX 


1227 


Sulfonamides,  see  also  Sulfanilamide, 

acetylation  of,  folate  analogs,  586-587 
/9-Sulfopropionate, 

intercharge  distance,  7 

succinate  dehydrogenase,  242-243 
Sunflower  stems, 

auxin  transport,  mercurials,  967 

growth  of,  malonate,  197 

respiration    (endogenous),    mercurials, 

881 
Sweet  potato,  see  also  Ipomea, 

citrate  oxidation,  mercurials,  878 

oxidative  phosphorylation,  mercurials, 

873 
Sympathetic  ganglia,  transmission, 

malonate,  211 


Taka-/?-glucosidase,   phenol-a-glucoside, 

271 
Talose,  fructokinase,  376 
Tapeworms,  see  Echinococcus  and  Hyme- 

nolepis 
Tartrate, 

fumarase,  275,  277-279 

ionization  constants,   8 

lactate  dehydrogenase,  437 

permeability  of  erythrocytes  to,  188 

phosphatases,  440-442 
D-Tartrate, 

L-tartrate  dehydrase,  601 

meso-tartrate  dehydrase,  601 
L-tartrate,  D-tartrate  dehydrase,  601 
meso-Tartrate, 

a-ketoglutarate  oxidation,  432 

malate  dehydrogenase,  596 

phosphatases,  440-441 

pyruvate  oxidation,  432 
D-Tartrate  dehydrase,  analogs,  601 
L-Tartrate  dehydrase,  analogs,  601 
meso-Tartrate  dehydrase,  analogs,  601 
Tartronate    (hydroxy malonate), 

glycolysis,  238 

D-a-hydroxy  acid  dehydrogenase,  437 

intercharge  distance,  6 

ionization  constants,  8 

ketogenic  activity,  237 


lactate  dehydrogenase,  237-238, 436-437 

malate  dehydrogenase,  237,  596 

malate    dehydrogenase   (decarboxylat- 

ing),  238,  597 

occurrence  in  tissues,  225,  238 

phosphatases,  238 

respiration,  238 

succinate  dehydrogenase,   36,  40,   237 

D-tartrate  dehydrase,  601 

weso-tartrate  dehydrase,  601 

tartronate  semialdehyde  reductase,  602 

Tartronate  semialdehyde  reductase,  ana- 
logs, 602 

Terephthalate, 

D-amino  acid  oxidase,  341,  344 
aspartate :  a  -  ketoglutarate      transami- 
nase, 334 

glutamate  dehydrogenase,  331 
intercharge  distance,  6 
kynurenine:  a-ketoglutarate     transami- 
nase, 607-608 

succinate  dehydrogenase,  37 
tyrosinase,  300 

Testis, 

citrate  formation  in,  malonate,  105,  108 
malonate  decarboxylation  in,  232 
respiration  (endogenous),  malonate,  175 
transaminases  in  vivo,  deoxypyridoxol, 
570 

Tetradecane-l,14-dicarboxylare,   kynure- 
nine:a-ketoglutarate  transaminase,  608 

Tetraethylammonium  ion,  renal  transoprt 
of, 

dehydroacetate,  626 
malonate,  204 
mercurials,  921 

Tetrahydrofolate,  biosynthesis  of, 
analogs,  581-584 

Tetrahymena  geleii, 

acetate  utilization,  <raw5-cyclopentane- 
1,2-dicarboxylate,  241 
succinate   dehydrogenase,   irans-cyclo- 
pentane-l,2-dicarboxylate,   37 

Tetrahymena   pyriformis,   glycolysis   (an- 
aerobic), 
oxamate,  434 

Tetraiodothyroacetate,  thyroxine  deiodi- 
nase,  602 


1228 


SUBJECT    INDEX 


3,5,3',5'-Tetraiodothyropropionate,     thy- 
roxine deiodinase,  602 
Tetrathionate,  696-700, 
blood  GSH,  696 
chemical  properties,  697 
conversion  to  thiosulfate,  699-700 
cyanide  poisoning,   696 
cytochrome  reduction  by,  696 
determination   of  protein   methionine, 
696 

enzyme  inhibitions,  698-699 
GSH  levels  in  tissues,  700 
nephrotoxicity,   700 
oxidation  of  protein  SH  groups,  697- 
698 

oxidation  of  thiols,  697-698 
oxidation-reduction  potential,  697 
reaction  with  SH  groups,  697 
thromboangiitis  obliterans,  696 
toxicity,  700 
Tetrolate,  fatty  acid  biosynthesis,  614 
Tetrolyl-CoA,    fatty    acid    biosynthesis, 

613-614 
Tetrose-diphosphate,    see    D-Threose-2,4- 

diphosphate 
Thea  pollen,  malonate  metabolism  in,  228 
Theobromine,  NAD  nucleosidase,  492 
Theophylline,  NAD  nucleosidase,  492 
Thetin:homocysteine   transmethylase,   a- 

nalogs,   356-357 
Thiaminase,  analogs,  523-525 
Thiamine, 

analogs  of,  see  also  Oxythiamine,  Pyri- 
thiamine,  and  others,  514-534, 

central  nervous  system,  527,  530-531 
enzyme  inhibitions,  518-519 
growth  of  microorganisms,   528-530 
mechanisms  of  action,  532-534 
metabolic  disturbances  by,  533-534 
neuromuscular  function,  531 
phosphorylation  of,  519 
pyruvate  accumulation,  520-521 
resistance  to,  528-529 
sites  of  action,  516 
structures  of,  517 
summary  of  actions,  532 
thiaminase,  523-525 
thiamine  deficiency,  530-531 


thiamine  kinase,  522-523 
thiamine  levels  in  tissues,  525-526 
thiamine-PP  levels  in  tissues,  525-527 
tissue  levels  of,  527-528 
toxicity,  530-531 
types  of,  516-518 
metabolism  of,  pathways,  514-515 
reaction  with  mercurials,  774 
urinary  excretion  of,  oxythiamine,  525 

Thiamine  disphosphatase,  mercurials,  860 

Thiamine  disulfide, 
structure  of,  517 
thiamine  kinase,  523 

Thiamine-diphosphate, 

metabolic  functions  of,  514-516 
tissue  levels  of,  analogs,  525-527 

Thiamine   kinase, 
analogs,  522-523 

role  of  inhibition  in  toxicity  of,  532- 
533 
nucleotides,  475 

Thiazole-pyrophosphate,  see  also  4-Me- 
thyl-5-hydroxyethylthiazole-PP,  struc- 
ture of,  517 

Thiazolidine  rings,  reactivity  of  enzyme 
SH  groups,  644,  646 

Thiazoline  rings,  reactivity  of  enzyme  SH 
groups,  644,  646 

/3-2-Thienylalanine, 

phenylalanine  hydroxylase,  354 
L-phenylalanine:sRNA    ligase    (AMP), 
354 

Thienylglycine,  glycine  uptake  by  ascites 
cells,  265 

Thimerosal  (Merthiolate),  structure  of, 
970 

Thiobacillus  thiooxidans, 

CO2  fixation,  mercurials,  892 
respiration  (endogenous),  malonate,  168 

Thioesters,  splitting  by  mercurials,   751 

Thioglycolate,  glutathione  oxidase,  593 

Thioglycosidase,  glucono-l,4-lactone,  429 

6-Thioinosinemonophosphate,  IMP  dehy- 
drogenase, 471,  481 

Thiol-disulfide  equilibria,  639,  656 

Thiols,  see  also  Sulfhydryl  groups, 
equilibria  with  disulfides,  639,  656 
role  in   metabolism,   636-637 


SUBJECT    INDEX 


1229 


Thiomerin,  see  Mercurin 
Thionicotinamide-NAD, 

alcohol  dehydrogenase,  497 

lactate  dehydrogenase,   497 
Thiophene-2,5-dicarboxylate, 

carcinostasis,   415 

glucose  metabolism  in  tumors,  415 
Thiophosphate,    oxidative    phosphoryla- 
tion, 447-448 
Thiosulfate, 

formation  from  tetrathionate,  699-700 

sulfate  uptake  by  yeast,  267 

sulfite  oxidase,  451 

utilization  by  Neurospora,  sulfate,  451- 

452 
Thiosulfate  reductase,  quinacrine,  555 
Thiosulfate  transulfurase, 

o-iodosobenzoate,  713 

mercurials,  relation  to  SH  groups,  803 
Thiothiamine,  thiamine  kinase,  523 
Thiourea,  urease,  603,  610 
L-Threonine, 

L-amino  acid  oxidase,  340 

aspartokinase,   356 

homoserine  kinase,  357 
Threonine  aldolase,  o-iodosobenzoate,  713 

718 
Threonine  dehydrase, 

mercurials,  coenzyme  displacement,  787 

serine,  357 
Threonine  dehydrogenase, 

o-iodosobenzoate,  713 

mercurials,  856 
Threonine  synthetase,  analogs,  357 
D-Threose-2,4-diphosphate, 

glycolysis,  409 

3  -  phosphoglyceraldehyde     dehydroge- 
nase, 408-409 

photosynthesis,  409 
Thrombin,  tosylagmatine,  375 
Thromboangiitis  obliterans,  use  of  tetra- 
thionate in,  696 
Thymidine, 

aspartate  carbamyltransferase,  468 

5'-nucleotidase,  472 
Thymidinemonophosphate  (TMP), 

aspartate  carbamyltransferase,  468 

phosphodiesterase,  473 


Thymidylate  kinase,  deoxy-GMP,  475 
Thymidylate    synthetase,    analogs,    476, 

479 
Thymine,  NAD  nucleosidase,  492-493 
Thymocytes,  glycolysis, 

malonate,   126 
Thymus, 

amino    acid    accumulation,    malonate, 

103 

citrate  levels  in,  sequential  inhibition 

by  malonate  and  fluoroacetate,  112 

malonate  levels  in  vivo,  102 

succinate  dehydrogenase,  32 

succinate  levels  in  vivo,  malonate,  102 
Thymus  nuclei, 

ATP  levels  in,  malonate,  189 

glucose    metabolism, 
dehydroacetate,  624 
2-deoxyglucose,  393-394 

respiration,  malonate,  189 

respiration    (glucose),    dehydroacetate, 

624 
Thyone  muscle,  glycolysis, 

mercurials,  876 
Thyroid, 

glucose  metabolism,  malonate,  131 

glycolysis  (aerobic),  malonate,  128 

iodide  uptake, 
malonate,  209 
mercurials,  910 

pyruvate  oxidation,  malonate,  76 

respiration  (endogenous),  malonate   176 

respiration  (glucose),  mercurials,  883 
Thyronine,  thyroxine  deiodinase,  603 
D-Thyroxine,  L-thyroxine  deiodinase,  602 
Thyroxine  deiodinase, 

analogs,  602-603 

mercurials,  778 

quinacrine,  555 
Tibial  condyles,  chondroitin  sulfate  bio- 
synthesis, 

malonate,   166 
Titration  of  enzyme  SH  groups, 

o-iodosobenzoate,  714-715 

mercurials,  766,  798-809 

porphyrindin,  667 

tetrathionate,  699 
Titration  of  protein  SH  groups. 


1230 


SUBJECT    INDEX 


ferricyanide,    672 
iodine,  680-681 
o-iodosobenzoate,  703-704 
mercurials,  762-766 
tetrathionate,  697 

TMP,  see  Thymidinemonophosphate 

Toadfish,  see  Opsanus  tau 

Tobacco, 

citrate  formation,  malonate,  105 

citrate  oxidation,  malonate,  79,  87 

cycle  intermediate  levels  in,  malonate, 

107,  111 

glycolate  levels  in,  a-hydroxysylforates, 

439 

growth  of  stem  cultures,  malonate,  197 

a-ketoglutarate  oxidation,  malonate,  80 

malate  oxidation,  malonate,  82 

malonate, 

metabolism  in,  228 
occurrence  in,  225 
pH  effects,   190 
NADP  reduction  by  hexose-P's,  mer- 
curials, 885 

pentose-P  pathway,  mercurials,  885 
protein  biosynthesis,   107,  156 
respiration  (endogenous),  malonate,  172 
succinate  accumulation,  malonate,  91 
succinate  dehydrogenase,  28 

Tobacco   mosaic   virus, 

inactivation  by  mercurials,   976,   979- 
980 

infectivity  of,  mercuric  ion,  741 
proliferation  of, 
malonate,  194 
mercurials,  976,  979 
titration  of  SH  groups, 
iodine,  681 
porphyrindin,  667 

Toluates,  see  Methylbenzoates 

j)-Toluenesulfonate,  D-amino  acid  oxidase, 
343 

p-Tolylalanine,  L-phenylalanine:sRNA  li- 
gase  (AMP),  355 

m-Tolyl-sulfate,  arylsulfatase,  443 

Tomato, 

pyruvate  oxidation,  malonate,  74 
respiration  (endogenous),  malonate,  171 
succinate  dehydrogenase,  malonate,  27 


Tosylagmatine, 

blood  clotting,  375 

thrombin,  375 

trypsin,  375 
Toxicity,  see  also  Lethal  doses, 

3-acetylpyridine,  489,  494 

alkylmalonates,  2 

6-deoxy-6-fluoroglucose,  404-405 

2-deoxyglucose,  401 

fluoromalonate,  239 

hydrogen  peroxide,  696 

o-iodosobenzoate,  725 

kojic  acid,  349 

malonate,   1-2,  217-221 

mercurials,  924-925,  950-957 

oxamate,  434 

pyridoxal  analogs,  562,  573-574,  577- 

578 

quinacrine,  546 

tetrathionate,  700 

thiamine  analogs,  530-531 
Toxoflavin, 

bacterial  growth,  538 

structure  of,  537 
Toxoplasma  gondii, 

infection  by,  deoxypyridoxol,  576 

respiration  (glucose),  mercurials,  882 
Toxopyrimidine, 

central  nervous  system,  578 

enzyme  inhibitions,   578 

glutamate  decarboxylase  in  brain,  571 

structure  of,  563 

toxicity,  562,  564 
Toxopyrimidine-phosphate,   tyrosine  de- 
carboxylase, 578 
TPN,  see  NADP 

TpTpTpT,  phosphodiesterase,  473 
Transaldolase,  phosphate,  412 
Transaminases, 

alanine:  a-ketoglutarate, 
aminoxyacetate,  358 
analogs,  334 
L-cycloserine,  360 
deoxypyridoxol  in  vivo,  569 
malonate,  64 
mercurials,  856-857 

alanine:pyruvate,  deoxypyridoxol,  569 

y-aminobutyrate:  a-ketoglutarate, 


SUBJECT    INDEX 


1231 


aminooxyacetate,  358-359 

D-cycloserine,  359 

malonate,  64 

mercurials,  857 
6-  amino  valerate :  a  -ketoglutarate,    mer- 
curials, 857 
asparagine :  a  -  ketoglutarate, 

cycloserines,  360 
asparagine: pyruvate,  mercurials,  857 
aspartate:a-ketoglutarate, 

analogs,  334,  355 

ferricyanide,  675 

malonate,  64 

mercurials,  808,  827,  857 
glutamine:pyruvate, 

GSSG,  662 

mercurials,  857 
glycine:  a-ketoglutarate,  mercurials,  857 
kynurenine:  a-ketoglutarate, 

acetate,  608 

analogs,  595,  607-610 

malonate,  64 
pyridoxal  analogs,  569-570 
pyridoxamine-oxalacetate,  oxalacetate 
analogs,  600 
toxopyrimidine,  578 
tryptophan:a-ketoglutarate,  mercurials 
857 
tyrosine:a-ketoglutarate, 

analogs,  305-306 

o-iodosobenzoate,  713 

mercurials,  857 
Transfer  RNA,  biosynthesis  from  nucleo- 
tides, 

mercurials,  820 
Transhydrogenase,  see  NAD:NADP  trans- 

hydrogenase 
Transketolase, 
mercurials,  857 
oxythiamine  in  vivo,  522 
phosphate,  412 
thiamine  analogs,  519,  522 
Transmembrane  potentials,  see  Membrane 

potentials 
Transphosporylases,  analogs,  444-447 
Trehalose,  a-glucosidase,  416 
Treponema  pallidum,  resistance  to  mer- 
curials, 983-985 


2, 4, 7  -  Triamino  -  6  -  o  -  methylphenylpteri- 
dine,  folinate  formation,  582 

Tribromophenol,   D-amino   acid   oxidase, 
344 

Tricarboxylate  cycle, 
ferrocyanide,  677-678 
y-hydroxy-a-ketoglutarate,  615-616 
intermediates     of,     concentrations     in 
cells,  88-90 

limiting  reactions  in,  70 
malonate,  69-90 
mercurials,  877-879 

Trichloroacetate, 

pantoate:/3-alanine  ligase,  598 
tyrosinase,  300 

a,a,/3-Trichloropropionate,    pantoate:/5-a- 
lanine  ligase,  598 

2,6,8-Trichloropurine,  uricase,  284 

Trichomonas  foetus,  respiration  (endoge- 
nous), 
malonate,  173 

Trichomonas    suis,    respiration    (endoge- 
nous), 
malonate,  173 

Trichomonas  vaginalis,   respiration   (glu- 
cose), 
mercurials,  882 

Trichophyton  interdigitale,  growth  of, 
dehydroacetate,  632 

Trichophyton  mentagrophytes,  growth  of, 
dehydroacetate,  632 

Trichophyton  rubrum,  respiration  (endo- 
genous), 
malonate,   169 
mercurials,  880 

Trichophyton  schoenleini,  respiration  (en- 
dogenous), 
malonate,  169 

Triethylsulfonium    ion,    thetin: homocys- 
teine transmethylase,  356 

Trifluoroacetyltryptamine,  chymotrypsin, 
371 

Trifluoroacetyltryptophanaraide,   chymo- 
trypsin, 371 

Trifluoroacetyltyrosinamide,  chymotryp- 
sin, 371 

Trifluorothiamine, 
Bacillus  growth,  531 


1232 


SUBJECT    INDEX 


thiamine  deficiency,  531 

tumor  growth,  531 
Trigonelhne, 

glucose  dehydrogenase,  501-502 

lactate  dehydrogenase,  501-502 

NAD  nucleosidase,  488,  491 

structure  of,  488 
Trihydrobenzoates,  dopa  decarboxylase, 

312 
Triiodophenol,  D-amino  acid  oxidase,  344 
Triiodothyroacetate,  intestinal  transport 

of, 

mercurials,  911,  913-914 
Triiodothyronines,  thyronine  deiodinase, 

602-603 
Trimesate, 

glutamate  dehydrogenase,  330 

structure  of,  330 
Trimetaphosphimate,  oxidative  phospho- 
rylation, 448 
Trimethylacetate,  tyrosinase,  300 
Trimethylammonium    ion,    thetin:homo- 

cysteine  transmethylase,  356 
Trimethylenediamine,    diamine    oxidase, 

362 
1,3,7-Trimethylurate,  uricase,  283-284 
1,3,9-Trimethylurate,  uricase,  283-284 
2,4,6-Trinitrophenol,    D-amino   acid   oxi- 
dase, 348 
Triose-phosphate  dehydrogenase,   see  3- 

Phosphoglyceraldehyde  dehydrogenase 
Triose-phosphate    isomerase,    phosphate, 

412 
Tri peptidase,  o-iodosobenzoate,  713 
Triphosphate,  oxidative  phosphorylation, 

448 
Triphosphoinositide   phosphodiesterase, 

mercurials,  858 
Triphosphoinositide     phosphomonoester- 

ase,  mercurials,  858 
Tripneustes  esclulentus,  development  of, 

mercurials,  964 
Tripolyphosphate, 

creatine  kinase,  446 

hexokinase,  383 

yeast  fermentation,  383 
True  inhibitor  constant,  definition  of,  252 
Trypanosoma  cruzi, 


glucose  utilization,  malonate,  127 

succinate  accumulation,  malonate,  91, 

127 

succinate  dehydrogenase,  malonate,  28 
Trypanosoma  hippicum,  respiration  (endo- 
genous), 

malonate,  173 
Trypanosomes, 

growth  of,  quinacrine,  559 

respiration,  quinacrine,  559 
Trypsin, 

dehydroacetate,  622 

inactivation  of  phosphoribosyl-ATP  py- 

rophosphorylase,  351 

macroions,  456-457 

mercurials,  797-798 

tosylagmatine,  375 
Tryptamine, 

chymotrypsin,  371,  374 

dopa  decarboxylase,  308 

tryptophan  pyrrolase,  325 

L-tryptophan:sRNA  ligase  (AMP),  326 
Tryptazan, 

incorporation  into  enzymes,  326 

maltase  biosynthesis,  326 

tryptophan  pyrrolase,  325 

L-tryptophan:sRNA  hgase  (AMP),  326 
Tryptophan, 

biosynthesis  of,  analogs,  321,  323 

histidine  decarboxylase,  352-353 

metabolism  of, 
analogs,  321-326 
deoxypyridoxol,  572 
pathways  of,  322 
D -Tryptophan, 

chymotrypsin,  374 

tryptophan  pyrrolase,  325 

L-tryptophan:sRNA  ligase  (AMP),  326 
L-Tryptophan, 

arginase,  337 

chymotrypsin,  374 

dipeptidase,  367 

feedback    inhibition    of    anthranilate 

synthesis,  321 

intestinal  transport  of  iodotyrosine,  265 

tyrosine:a-ketoglutarate  transaminase, 

306 

uptake  by  ascites  cells,  analogs,  265-266 


SUBJECT    INDEX 


1233 


Tryptophanamides,    chymotrypsin,    271, 

371 
L-Tryptophanase, 

analogs,  323-324 

mercurials,  858 

methylindoles,  321 

toxopyrimidine,  578 

D-tryptophan,  268 
Tryptophan  hydroxylase  (phenylalanine 

hydroxylase),  analogs,  325-326 
Tryptophan :  a  -  ketoglutarate     transami- 
nase, see  Transaminases 
Tryptophan  peroxidase,  see  Tryptophan 

pyrrolase 
Tryptophan  pyrrolase  (tryptophan  pero- 
xidase), 

hematin  analogs,  603 

induction  of,  /?-azaguanine,  478 

mercurials,  817 

tryptophan  analogs,  324-325 
L-Tryptophan :  sRNA  ligase  (AMP),  ana- 
logs, 326 
Tryptophan  synthetase, 

analogs,  321 

ferricyanide,  676 
Tumor  2146  (mouse),  growth  of, 

2-acetamido-2-deoxygluconolactone, 

428 

glucaro-l,4-lactone,  428 
Tumors,  see  also  specific  tumors, 

blebbing  of,  see  Sarcoma  37 

DNA  levels  in,  2-deoxyglucose,  399 

fatty  acid  biosynthesis,  malonate,  148- 

149 

glucose  metabolism,  thiophene-2,5-  di- 

carboxylate,  415 

glycolysis, 

2-deoxyglucose,  392 
hydrogen  peroxide,  695 

glycolysis  (anaerobic),  D-glucosone,  385 

growth  of, 

6-azauracil,  478 
2-deoxyglucose,  400-401 
malonate,  200-202 
pyridoxal  analogs,  576-577 
riboflavin  analogs,  538 
trifluoro thiamine,  531 

mercurial  uptake  in  vivo,  969-970 


metabolism  in,  ferricyanide,  677 

protein  biosynthesis,  malonate,  156 

respiration    (endogenous),    trans-aconi- 

tate,  273 

transaminases  in  vivo,  deoxypyridoxol, 

570 
Tungstate, 

Agrohacterium  growth,  615 

Aspergillus  growth,  614 

Azotobacter  growth,  614 

molybdate  uptake,  614 

molybdenum  deficiency,  614-615 

molybdenum  levels  in  tissues,  614 

nitrate  reductase,  615 

xanthine  oxidase  in  vivo,  614 
Turanose,  a-glucosidase,  416,  423 
Turnip  yellow  mosaic  virus,  splitting  into 

subunits  by  mercurials,  980 
Tyramine, 

dopa  decarboxylase,  308 

dopamine  /^-hydroxylase,  320 

oxidation  of,  kojic  acid,  350 

tyrosine:a-ketoglutarate  transaminase, 

306 

L-tyrosine:sRNA  ligase  (AMP),  307 
D-Tyrosinamide,  chymotrypsin,  271 
L-Tyrosinamide, 

cathepsin  C,  375 

L-tyrosine:sRNA  ligase  (AMP),  307 
Tyrosinase, 

acetate,  300-301 

analogs,  300-305 

benzoate,  349 

cafFeate,  314 

fluoride,  300 

maleate,  301 

mercurials,  794 

oxalate,  300 

pyrophosphate,  301 

succinyl  peroxide,  694 

D-  and  L-tyrosines,  268 
)3-Tyrosinase,  ferricyanide,  676 
Tyrosine, 

histidine  decarboxylase,  352-353 

metabolism  of, 
analogs,  302-320 
pathways  of,  303 

phenylalanine  deaminase,  355 


1234 


SUBJECT    INDEX 


D-Tyrosine,    L-tyrosine:a-ketoglutarate 

transaminase,  306 
Tyrosine  decarboxylase, 

analogs,  306-307 

cafFeate,   314 

folate  analogs,  586 

permanganate,  660 

toxopyrimidine,  578 

toxopyrimidine-P,  578 
Tyrosinehydroxamide,  chymotrypsin,  371 
Tyrosine:a-ketoglutarate   transaminase, 

see  Transaminases 
L-Tyrosine:sRNA  ligase  (AMP),  analogs, 

307 

u 

UDP,  see  Uridinediphosphate 
UDPgalactose-4-epimerase,    mercurials, 

858 
UDPglucose  dehydrogenase, 

GSSG,  662 

o-iodosobenzoate,  713 

mercurials,  858-859 
UDPglucose:a-l,4-glucan-a-4-glucosyl- 

transferase, 

2-deoxyglucose-6-P,  391 

UDP  and  UMP,  476 
UDPglucose  -  glycogen    glucosyltransfer- 

ase,  D-glucosamine,  382 
UDPglucose  pyrophosphorylase, 

galactose- 1-P,  603 

D-glucosamine,  382 
UDPglucose-starch  glucosyltransferase, 

see  UDPglucose:a-l,4-glucan-a-4-gIuco- 

syltransferase 
UDP  glucuronyltransferase, 

o-iodosobenzoate,  713 

mercurials,  817,  859 
Ulcers,  treatment  with  pepsin  inhibitors, 

458 
Ulothrix  zonata,  succinate  dehydrogenase, 

malonate,  27 
UMP,  see  Uridinemonophosphate 
Undecane  -1,11-  dicarboxylate,     kynuren- 

ine:a-ketoglutarate   transaminase,   608 
Uracil, 

D-amino  acid  oxidase,  545 

intestinal  transport  of,  mercurials,  911 


NAD  nucleosidase,  493 

5'-nucleotidase,  472 
6-Uracilmethylsulfone,  orotate  transphos- 

phoribosylase,  473 
6-Uracilsulfonamide,    orotate   transphos- 

phoribosylase,  473 
6-Uracilsulfonate,  orotate  transphospho- 

ribosylase,  473 
Urate, 

diamine  oxidase,  365 

oxidation  of,  kojic  acid,  350 

uptake  by  erythrocytes,  hypoxanthine, 

267 
Urate  oxidase,  see  Uricase 
Urate  riboside,  inosine  hydrolase,  471 
Urea, 

formation  of, 

deoxypyridoxol,  572-573 
malonate,   157-158 
Urea  cycle,  deoxypyridoxol,  572-573 
Urease, 

acetamide,  603 

analogs,  603,  610 

cystine,  662 

dehydroacetate,  621-622 

ferricyanide,  673,  676 

iodine,  682-683,  687 

malonate,  64 

mercurials,  773,  778,  794,  859 
pH  effects,  794 
protection  by  ascorbate,  778 
type  of  inhibition,  773 

methylurea,  610 

phenylisocyanate,  649 

porphyrindin,  669 

SH  groups  of,  titration  of,  643 

succinyl  peroxide,  694 
Urechis  caupo  eggs,  development  of, 

ferricyanide,  678 
Urechis  unicinctus  eggs,  development  of, 

o-iodosobenzoate,  727 
y-(3,4-Ureylenecyclohexyl)butyrate, 

structure  of,  588 

yeast  fermentation,  588-589 
Uricase, 

hydrogen  peroxide,  691,  693 

o-iodosobenzoate,  713 

mercurials,  859 


SUBJECT    INDEX 


1235 


pterin-6-aldehyde,  288 

purine  analogs,  283-286 
Uridine, 

aspartate  carbamyltransferase,  468 

5 '-nucleotidase,  472 
Uridinediphosphate  (UDP), 

fructose- 1,6-diphosphatase,  470 

orotidylate  decarboxylase,  473,  479 

UDPglucose:a-l,4-glucan-a-4-glucosyl- 

transferase,  476 
Uridinemonophosphate  (UMP), 

adenylosuccinate  synthetase,  467 

aspartate  carbamyltransferase,  468 

deoxycytidylate  deaminase,  469 

5 -nucleotidase,  471 

orotidylate  decarboxylase,  473,  479 

phosphatase,  439 

pyrophosphatase,  475 

UDPglucose:a-l,4-glucan-a-4-glucosyl- 

transferase,  476 
Uridinetriphosphatase  (UTPase), 

ADP,  446 

IDP,  446 
Uridinetriphosphate  (UTP), 

aspartate  carbamyltransferase,  468 

isocitrate  dehydrogenase,  509 

NADH  oxidase,  511 

orotidylate  decarboxylase,  473,  477 
Urine, 

ethylmalonate  in,  225 

malonate  in,  225-226 

methylmalonate  in,  224 

tartronate  in,  238 
Urinary  flow, 

dehydroacetate,  625 

malonate,  206 

mercurials,  917-918 
Urinary  pH,  malonate,  206 
Urocanase, 

hydrogen  peroxide,  693 

o-iodosobenzoate,  713 

mercurials,  protection  by  urocanate,  783 

permanganate,  660 
Urocanate  oxidase,  malonate,  64 
Uroporphyrinogen    decarobxylase,    mer- 
curials, 859 
Ustilago  maydis,  respiration  (glucose), 

malonate,  133-134 


Uterus, 

contracture,  o-iodosobenzoate,  724 
motility,  o-iodosobenzoate,  724 
UTP,  see  Uridinetriphosphate 
UTPase,  see  Uridinetriphosphatase 


Vaccinia  virus, 

infectivity  of,  mercurials,  979-980 

proliferation  of,  malonate,  193 
Valerate  (pentanoate), 

carboxypeptidase,  366 

glutamate  decarboxylase,  328 

kynurenine:a-ketoglutarate     transami- 
nase, 608-609 

lactate  dehydrogenase,  436 

leucine  decarboxylase,  352 
f5-Valerolactam,    A  i-pyrroline-5-carboxy- 

late  dehydrogenase,  336 
Valine, 

alanine:  ct-ketoglutarate     transaminase, 

334 

dipeptidase,  367 

incorporation  into  proteins,  a-amino-^- 

chlorobutyrate,   351 
D-Valine,  L-alanine  dehydrogenase,  354 
L-Valine,  L-amino  acid  oxidase,  340 
Valylleucine,  penicillinase,  599 
Valylvaline,  penicillinase,  599 
Vanillin,     dehydroshikimate     reductase, 

593,  604-605 
Vascular  smooth  muscle,  malonate,  212 
Ventricle,  see  Heart 
Venturia    inaegualis,    ascosporulation- 

malonate,  195 
Veratroyl-^-glucoronide,    glucuronidases, 

426-427 
Vessels,  see  Vascular  smooth  muscle 
Vetch  leaves,  malonate  occurrence  in,  224 
Vibrio  cholera, 

amino  acid  accumulation,   deoxypyri- 

doxol,  576 

growth  of, 

dehydroacetate,  632 
deoxypyridoxol,  576 
thiamine  analogs,  522,  530 
Vibrio  metchnikowii,  growth  of, 

dehydroacetate,  532 


1236 


SUBJECT    INDEX 


Vigna  sinensis,  glutamate  metabolism, 

malonate,  153 
Viruses,  see  also  specific  viruses, 
proliferation  of, 

o-iodosobenzoate,  728 
malonate,  192-194 
mercurials,  976-981 
Vitamin  B,,  see  FAD  and  Riboflavin 
Vitamin  B^,  see  Thiamine 
Vitamin  Bg,  see  Pyridoxal  and  derivatives 
Vitamin  Bjj,  see  Cyanocobalamin 
Vitamin   Kj   reductase,   o-iodosoenzoate, 
713 

W 

Walker  carcinosarcoma, 
2-deoxyglucose,    386 
glutamate  metabolism,  malonate,  152 
glycolysis,  ferricyanide,  677 
malonate  levels  in  vivo,  102 
protein  biosynthesis,  malonate,  156 
pyruvate  oxidation,  mercurials,  878 
succinate  levels  in  vivo,  malonate,  102 

Water, 

corneal  transport  of,  mercurials,  91 1 
intestinal  transport  of,  mercurials,  916 
renal  transport  of,  mercurials,  917-918 

Western    equine   encephalitis    virus,    in- 
fectivity  of, 
mercurials,  979 

Wheat, 

malonate  occurrence  in,  224-225 
respiration  (endogenous),  malonate,  170 
184 

Woodroach,  see  Leucophaea 


Xanthine, 

NAD  nucleosidase,  492 

oxidation  in  liver,  kojic  acid,  350 

uricase,  285 
Xanthine:cytochrome   c   oxidoreductase, 

myoglobin,  603 
Xanthine  dehydrogenase,  mercurials,  859 
Xanthine  oxidase, 

analogs,  279-289 

8-azaguanine,  477 


hydrogen  peroxide,  693-694 

9-hydroxyethylriboflavin  analog,  544 

iodine,  687 

o-iodosobenzoate,  669,  704,  713,  718 

mercurials,  783,  796-797,  803,  807,  814 
pH  eifects,  796-797 
potentiation   of  inhibition    by   sub- 
strate, 807 

protection  by  hypoxanthine,  783 
relation  to  SH  groups,  803 
spontaneous  reversal,  814 

oxygen  inactivation  of,  659 

porphyrindin,    668-669 

pteridine  analogs,  285-289 

purine  analogs,   280-283 

quinacrine,  555 

tungstate  in  vivo,  614 
Xanthomonas  phaseoli,  succinate  dehydro- 
genase, 

malonate,   19,  26 
Xanthopterin, 

guanase,  288 

structure   of,   287 

xanthine  oxidase,  289 
Xanthopterin-7-carboxylate,  xanthine  o- 

xidase,  289 
Xanthosine,  inosine  hydrolase,  471 
Xanthosine-5'-phosphate  aminase  (GMP 

synthetase),  psicofuranine,  476,  481 
Xanthurenate,  urinary  excretion  of, 

deoxypyridoxol,  572 
Xanthyl- cytochrome  c,  cytochrome  c,  592 
Xylanase,  mercurials,  859 
Xylenesulfonate,  arylsulfatase,  444 
Xylitol,  mutarose,  413-414 
Xylono- 1 ,4-lactone, 

a-glucuronidase,  426 

structure  of,  425 
Xylose, 

a-amylase,  420 

fructokinase,  376 

galactose  transport  by  intestine,  263 

/3-galactosidase,  418 

a-glucosidase,  423 

a-mannosidase,  422 

mutarotase,  413-414 

phosphoarabinose  isomerase,  411 

photosynthesis,  414 


SUBJECT    INDEX 


1237 


uptake  by  diaphragm,  mercurials,  911- 
912 
Xylulokinase, 

o-iodosobenzoate,  713 
mercurials,  859 


Yeast, 

acetate  oxidation,  malonate,  77,  116 
catalase  induction  in,  8-azaguanine,  478 
citrate    formation    from    acetate,    ma- 
lonate, 105 

coenzyme  A  levels  in,  mercurials,  750, 
885 

cycle   intermediate   concentrations   in, 
89 

2-deoxyglucose  uptake,  387 
ethanol  oxidation,  2-deoxyglucose,  395- 
396 

fermentation  (fructose),  6-deoxy-6-fluo- 
roglucose,  404 
fermentation  (glucose), 

biotin  analogs,  588-589 

6-deoxy-6-fluoroglucose,  404 

D-glucosone,  384-385 

iodine,  689 

mercurials,  875,  884 

tripolyphosphate,  383 
fumarate  oxidation,  malonate,  81 
glycolysis,  2-deoxyglucose,  392 
growth  of, 

benzoate,  349 

dehydroacetate,  632 

o-iodosobenzoate,    727 

malonate,   195 

mercurials,  971,  973-974 

cy-methylpyridoxol,  575 
growth  of  (menadione-stimulated),  fer- 
ricyanide,  678 

K+  fluxes,  mercurials,  898-900,  908 
maltase  biosynthesis,  p-fluorophenyla- 
lanine,  351 


mercurial  uptake  by,  898-900,  974 

nucleotide   levels  in,   mercurials,   884- 

885 

pyruvate  oxidation,  benzoate,  349 

resistance  to  mercurials,  983-985 

respiration  (endogenous), 

malonate,   169 

mercurials,  879-881 
respiration  (glucose), 

malonate,  124 

mercurials,  880-881,   884-885 
succinate  accumulation,  malonate,  92- 
93 
succinate  dehydrogenase, 

fumarate  K,,  38 

malonate,  27,  33,  187 
succinate  oxidation,  malonate,  22 
sulfate  uptake,  thiosulfate,  267 
Yoshida  ascites  hepatoma,  glycolysis  (ae- 
robic), 

tartronate,  238 
Yoshida    sarcoma,    respiration    (endoge- 
nous), 
malonate,  179 


Zahdel  hepatoma,  ADP-ATP  levels  in, 

2-deoxyglucose,  395 
Zinc, 

complexes  with  di-  and  tricarboxylates, 

12 

glutamate  dehydrogenase,  863 

Lactobacillus  growth,  452 

phosphatase  (acid),  452-453 
Zygorrhynchus  moelleri, 

acetate     oxidation,     malonic     diethyl 

ester,  236-237 

glucose     oxidation,     malonic     diethyl 

ester,  236-237 

succinate  oxidation,  malonate,  51,  53, 

187