Skip to main content

Full text of "Rotorcraft dynamics"

See other formats


NASA SP-352 



JL\.\^ir JL \«^r JL\.\_^A\.^ljLJL JL I *P JL J. ^I JT ^JLVJLJL\_aJ' 



A conference held at 

AMES RESEARCH CENTER 

Moffett Field, California 

February 13-15, 1974 




US. ft- 



NATIONAL AERONAUTICS AND SPACE ADMINISTRATION 



NASA SP-352 



ROTORCRAFT DYNAMICS 



A conference sponsored by 

Ames Research Center and the American Helicopter Society 

and held at Ames Research Center, Moffett Field, California 

February 13-15, 1974 



Prepared at Ames Research Center 




Scientific and Technical Information Office 1974 

NATIONAL AERONAUTICS AND SPACE ADMINISTRATION 

Washington, D.C. 



For sale by the National Technical Information Service 
Springfield, Virginia 22151 
Price $8.00 



PREFACE 



Events of recent years have clearly identified rotorcraft dynamics as one of the most critical technologies 
pacing the helicopter industry's efforts to develop new rotor concepts. And as rotary wing program 
investments have escalated, the financial stakes in the technical risks have become tremendous. The rotor 
dynamicist is suddenly very much in the critical path. 

Fortunately Research and Development efforts by industry in rotorcraft dynamics have been greatly 
augmented in recent years by stepped up in-house efforts on the part of NASA and the U. S. Army. New 
computer tools and more complete experimental data coming from many quarters^are bringing us to the 
threshold of a far more complete understanding of the problem. This increased R&D is reflected in the 
number of high quality papers that have exceeded the capacity of the dynamics session at recent AHS 
Annual Forums. 

With these thoughts in mind, the Specialists' Meeting on Rotorcraft Dynamics was organized to provide an 
opportunity for the principal investigators in the field to dialogue in greater depth than is possible at the 
American Helicopter Society Annual Forum. This is the first such meeting in the U. S. since the 
CAL/TRECOM Helicopter and V/STOL Dynamic Loads Symposium in 1963. 

This volume contains the formal presentations of the first four sessions of the meeting. Presentations of the 
fifth session and transcriptions of questions and panel discussions are contained in the supplement to this 
volume. 

E. S. Carter, Jr. 

Meeting General Chairman 



ill 



ORGANIZING COMMITTEE 

General Chairman 

Edward S. Carter, Jr., Sikorsky Aircraft 

Technical Chairman 

Robert A. Ormiston, U. S. Army Air Mobility R&D Laboratory, 
Ames Directorate 

Administrative Chairman 

James C. Biggers, NASA-Ames Research Center 

Chairman, AHS Dynamics Technical Committee 
E. Roberts Wood, Lockheed-California Company 



Session Chairmen and Co-Chairmen 

Session I 

Kurt H. Hohenemser, Washington University 
James R. Neff , Hughes Helicopters 

Session II 

E. Roberts Wood, Lockheed-California Company 
G. Alvin Pierce, Georgia Institute of Technology 

Session HI 

Peter J. Arcidiacono, Sikorsky Aircraft 
William E. Nettles, U. S. Army Air Mobility R&D Laboratory 
Eustis Directorate 

Session IV 

James J. CLeary, Boeing Vertol Company 

William G. Flannelly, Kaman Aerospace Corporation 

Session V 

Troy M. Gaffey, Bell Helicopter Company 



xv 



CONTENTS 

Paper Page 

Preface iii 

Organizing Committee iv 

SESSION I - ROTOR SYSTEM DYNAMICS 

1 Hingeless Rotor Frequency Response with Unsteady Inflow 

D.A.Peters 1 

2 Dynamic Stall Modeling and Correlation with Experimental Data 
on Airfoils and Rotors 

R. G. Carlson, R. H. Blackwell, G. L. Commerford and P. H. Mirick 13 

3 Computer Experiments on Periodic Systems Identification Using 
Rotor Blade Transient Flapping-Torsion Responses at High 
Advance Ratio 

K. H. Hohenemser and D. A. Prelewicz 25 

4 Dynamic Analysis of Multi-Degree-of-Freedom Systems Using 
Phasing Matrices 

R. L. Bielawa 35 

5 Some Approximations to the Flapping Stability of Helicopter Rotors 

J. C. Biggers 45 

6 Flap-Lag Dynamics of Hingeless Helicopter Blades at Moderate and 
High Advance Ratios 

P. Friedmann and L. J. Silverthom 55 

SESSION II - HELICOPTER VIBRATION AND LOADS - THEORY 

7 Correlation of Finite-Element Structural Dynamic Analysis with 
Measured Free Vibration Characteristics for a Full-Scale 
Helicopter Fuselage 

I. J. Kenigsberg, M. W. Dean and R. Malatino 67 



8 Coupled Rotor/ Airframe Vibration Prediction Methods 

J. A. Staley and J. J. Sciarra 81 

9 Helicopter Gust Response Characteristics Including Unsteady Aerodynamic 
Stall Effects 

P. J. Arcidiacono, R. R. Bergquist and W. T. Alexander, Jr 91 

10 Application of Antiresonance Theory to Helicopters 

F. D. Bartlett, Jr. and W. G. Flannelly 101 



Paper Page 

1 1 The Effect of Cyclic Feathering Motions on Dynamic Rotor Loads 

K. W. Harvey , 107 

12 Control Load Envelope Shaping by live Twist 

F. J. Tarzanin, Jr. and P. H. Mirick 115 

13 Application to Rotary Wings of a Simplified Aerodynamic Lifting 
Surface Theory for Unsteady Compressible Flow 

B. M. Rao and W. P. Jones 127 

SESSION III - ROTOR/VEHICLE DYNAMICS 

14 Rotor Aeroelastic Stability Coupled with Helicopter 
Body Motion 

Wen-Liu Miao and H. B. Huber 137 

15 An Application of Floquet Theory to Prediction of Mechanical 
Instability 

C. E. Hammond 147 

16 Theory and Comparison with Tests of Two Full-Scale Proprotors 

W. Johnson 159 

17 Experimental and Analytical Studies in Tilt-Rotor Aeroelasticity 

R. G. Kvaternik 171 

18 Comparison of Flight Data and Analysis for Hingeless Rotor 
Regressive Inplane Mode Stability 

W. D. Anderson and J. F. Johnston 185 

19 Hub Moment Springs on Two-Bladed Teetering Rotors 

W. G. O. Sonneborn and J. Yen 199 

20 Open and Closed Loop Stability of Hingeless Rotor Helicopter Air 
and Ground Resonance 

M. I. Young, D.J. Bailey and M.J. Hirschbein 205 

SESSION IV -HELICOPTER VIBRATION AND LOADS - APPLICATIONS 

21 Vertical-Plane Pendulum Absorbers for Minimizing Helicopter 
Vibratory Loads 

K. B. Amer and J. R. Neff 219 

22 The Evaluation of a Stall-Flutter Spring-Damper Pushrod in 
the Rotating Control System of a CH-54B Helicopter 

W. E. Nettles, W. F. Paul and D. 0. Adams 223 

23 Multicyclic Jet-Flap Control for Alleviation of Helicopter Blade 
Stresses and Fuselage Vibration 

J.L.McCloudIIIandM.Kretz . 233 



vi 



Paper Page 

24 Identification of Structural Parameters from Helicopter Dynamic 
Test Data 

N.Giansante and W. G. Flannelly 239 

25 Engine/Airframe Interface Dynamics Experience 

C. A. Fredrickson 249 

26 Hingeless Rotor Theory and Experiment on Vibration Reduction by 
Periodic Variation of Conventional Controls 

G. J. Sissingh and R. E. Donham 261 

SUPPLEMENT 

Foreword to Supplement 279 

Welcome 

C. A. Syvertson 280 

Opening Remarks 
E. S. Carter . . 280 

Dinner Address - What Can the Dynamicist Do for Future Army Aircraft? 
P. F. Yaggy 281 

Session V — Application of Dynamics Technology to Helicopter Design 283 

Panel 1 : Prediction of Rotor and Control System Loads 

Panel Members 283 

Comparison of Several Methods for Predicting Loads on a Hypothetical Helicopter Rotor 
R. A. Ormiston , 284 

Discussion, Panel 1 303 

Prepared Comments 

W. D. Anderson 303 

P. J. Arcidiacono 304 

R. L. Bennett 306 

W. Johnson 307 

A. Z. Lemnios 307 

R. H. MacNeal • • • 308 

F. J. Tarzanin 309 

R. P. White 311 

Survey of Panelists 313 

Questions and Answers, Panel 1 314 



Vil 



Page 

Panel 2 : Control of 1 /Rev Vibration 

Panel Members 315 

The User's Problem 

R. J. van der Harten 316 

D..F, Benton 320 

Technical Aspects of 1 /Rev Vibration 

W. F. Wilson . . , 331 

Discussion, Panel 2 . 322 

Questions and Answers, Panel 2 326 

Panel 3 : Integrating Dynamic Analysis and Helicopter Design 

Panel Members 327 

Discussion, Panel 3 328 

Prepared Comments 

R. W. Balke 328 

R. Gabel 329 

J. F. Johnston 332 

J. R. Neff 334 

W. F. Paul 335 

Questions and Answers, Panel 3 / 338 

Questions and Answers — Sessions I— IV 

Session I — Rotor System Dynamics 339 

Session II — Helicopter Vibration and Loads — Theory 347 

Session HI - Rotor/Vehicle Dynamics 355 

Session IV — Helicopter Vibration and Loads — Applications 361 

List of Attendees 367 



Vlll 



HIHGELESS ROTOR FREQUENCY RESPONSE WITH UNSTEADY INFLOW 



U.S. 



David A. Peters 
Research Scientist 

Ames Directorate 
Army Air Mobility R&D Laboratory 
Moffett Field, Calif. 94035 



Abstract 

Hingeless rotor frequency response calcula- 
tions are obtained by applying a generalized har- 
monic balance to the elastic blade flapping equa- 
tions. Nonuniform, unsteady induced flow effects 
are included by assuming a simple three-degree- 
of-freedom description of the rotor wake. Results 
obtained by using various models of elastic blade 
bending and induced flow are compared with exper- 
imental data obtained from a 7.5-ft diameter wind 
tunnel model at advance ratios from 0.0 to 0.6. 
It is shown that the blade elasticity and nonuni- 
form, unsteady induced flow can have a signifi- 
cant effect on the transient response character- 
istics of rotor systems. Good correlation be- 
tween theory and experiment is obtained by using: 
(i) a single rotating mode shape description of 
the elastic blade bending, (ii) an empirical form- 
ula for the quasi-steady induced flow behavior, 
and (iii) the apparent mass terms from potential 
flow for the unsteady induced flow characteris- 
tics. 

Notation 



a jn> b jn 

b 

B 

c 

Cj 

c M 



e 
e pc 

EI 

{£} 
F.G 

go>gs»Sc 

i 

lA 

[I] 

J 



two-dimensional lift-curve slope, 

rad -1 

harmonics of jth flapping mode 

number of blades 

tip loss factor 

blade chord, ft 

steady value of thrust coefficient, 

steady thrust/pirn 2 R l * 

harmonic perturbation of thrust 

coefficient 

harmonic perturbation of roll moment 

coefficient « roll moment/pir$2 2 R 5 , 

positive advancing blade down 

harmonic perturbation of pitch 

moment coefficient = pitch moment/ 

pirfi 2 R 5 , positive nose up 

dimensionless flapping hinge offset 

dimensionless radius of pocket 

cutout 

rotor blade bending stiffness, . 

lb-ft 2 

generalized response vector 

aerodynamic and inertial forces per 

unit blade span, lb/ft 

nondimensional harmonics of inertial 

forcing function, Eq. (12) 

/T 

apparent inertia of air, slug-ft 2 

identity matrix 

index referring to mode number 



J 

[K] 
[L] 
[L E ] 



ffi.niyy.myj 



M,[M'] 

dm 
dm 

m A 

n 

N 

[0] 

P 

r 
R 
S(0,i{i) 

H 



{u} 
Up,U T 

U„o,V„„ 



v(o,<|0 

w 
[W] 
{x} 
[Y] 



number of flap bending modes 
nondimensional apparent mass and in- 
ertia of impermeable disk 
control feedback matrix 
nonuniform induced flow matrix 
empirical value for quasi-steady 
portion of [L] 

rotor blade mass distribution, slug/ft 
nondimensional blade parameters 
R -R 



pacR^ 



pacR d 



m dr, 



pacR 1 * 



mr^ dr, 



mr<j>, dr 



rotor response matrix open loop, 

closed loop 

elemental apparent mass, slugs 

elemental mass flow, slugs/sec 

apparent mass of air, slugs 

index referring to harmonic number 

number of azimuthal harmonics 

null matrix 

first flap frequency divided by Q 

generalized coordinates 

steady values of q^ 

rotor blade radius coordinate, ft 

rotor blade radius, ft 

blade root moment, ft-lb 

blade parameter 



pacR2 



R 



m *j 



dr 



Presented at the AHS/NASA-Ames Specialists' Meet- 
ing on Rotor craft Dynamics, February 13-15, 1974. 



generalized control vector 

<e o e s e c8o8sSc A o x s^c> 

perpendicular, tangential components of 

air speed in undef ormed blade coordi- 
nate system, ft/sec 
freestream airspeed perpendicular and 
parallel to rotor shaft (V» positive 
down), ft /sec 
induced flow parameter «■ 

[y 2 + A(X + v)]/(y 2 + T 2 ) 1/2 

blade root shear, lb 

rotor blade flap deflection, ft 

frequency transform, Eq. (23) 

physical control vector <@ o s c z<f>a> 

control coupling matrix 

hub plunge deflection divided by R, 

positive down 

hub pitch angle, positive nose up, 

rad 



Y 
Y* 

e 



e o» 6 s» 6 c 
X 



Lock number, 1/myy 

equivalent Lock number, Eq. (35) 

blade pitch angle = 5 + (6 + 6 S 

sin iji + 6 C cos ^)e ±la ^ 

steady collective pitch angle 

rotor pitch perturbations 

total inflow (including induced flow) 



= X + 



X b f sln 



* 



X c f cos 



* 



iuf 



^o»^s»^c 



steady inflow ratio = V^/QR + v 
inflow perturbations (including in- 
duced flow), Eq. (10) 
advance ratio = U<»/fiR 
total induced flow = 



v + 



K 



+ v 



S R 



sin iji 
itoijj 



v o» v s» v c 

p 

a 


r T ,T S 


♦J 

* 


u 
SI 


8' 



+ v c — cos i|> I e' 

induced flow due to steady rotor 
thrust 

induced flow perturbations 
air density, slug/ft 3 
rotor solidity, bc/ffR 
induced flow time constants, rad~l 
hub roll angle, positive advancing 
blade down, rad 
orthogonal functions 
rotor blade azimuth position, non- 
dimensional time, rad 
' excitation frequency divided by Si 
rotor blade angular velocity, 
rad /sec 
3/3r 
3/3* 



The dynamic response characteristics of 
hingeless rotors are dependent upon the distrib- 
uted structural properties of the rotor blades, 
the local aerodynamic properties of the blade 
sections, and the detailed description of the 
aerodynamic environment. It is generally be- 
lieved i however j that reasonable predictions of 
rotor thrust and moments at low lift can be ob- 
tained by using some appropriately simplified 
models for the blade structure, section aero- 
dynamics^ and inflow distribution. The develop- 
ment of these simplified rotor models is useful 
for gaining insight into the basic dynamic mech- 
anisms of rotor response. Detailed calculations 
of dynamic airloads, necessary for many applica- 
tions, are usually too complex for use in basic 
dynamic research or preliminary design calcula- 
tions. 

The formulation of a minimum complexity 
rotor response model is the subject of several 
recent papers. One area of interest is the ef- 
fect of mode shape and mode number on rotor 
flapping response. Shupe 1 addresses the effects 
of the second flap mode, Ormiston and Peters 2 
compare various mode shape models for first and 
second flap modes, and Hohenemser and Yin 3 con- 
sider the effect of using rotating rather than 
nonrotating modes as generalized degrees of 
freedom. The fundamental conclusion, as 



clarified in Reference 3, is that for u < 0.8 a 
single rotating mode shape is adequate for model- 
ing the steady rotor flapping response. 

A second area of interest is the effect of 
induced flow perturbations on rotor flapping 
response. In Reference 1, a simple momentum 
theory predicts a significant effect of induced 
flow on steady rotor response. In Reference 2', a 
comparison of steady experimental and theoretical 
results indicates that, although there is a sig- 
nificant effect due to induced flow, momentum 
theory is inadequate for predicting this effect in 
forward flight. Alternate induced flow models are 
introduced and compared with the data, but no 
clear choice for the best model is found. In 
Reference 4, an unsteady momentum theory is used 
in hover to improve correlations with experimental 
frequency response data. ' 

The work in References 1 through 4 indicates 
that a minimum complexity analytic model for 
rotor dynamics must include appropriate degrees of 
freedom for both structural and induced flow per- 
turbations (certain flight dynamics programs 
presently include a simplified dynamic treatment 
of the induced f low^) . Unfortunately, while some 
success has been achieved using simple models of 
the rotor induced flow in hover, a completely 
satisfactory induced flow model for forward flight 
has not been found, not even for the condition of 
steady response. In addition, neither the phy- 
sical values of the induced flow time constants 
nor the frequency range in which they are impor- 
tant is known. The unsteady behavior of the in- 
duced flow contributes directly to the low fre- 
quency rotor control characteristics and to the 
coupled rotor/fuselage aeroelastic stability. In 
particular, induced flow perturbations contribute 
to the rotor damping available in pitch and roll 
(which is important for ground and air resonance 
calculations) . It is consequently important to 
understand the dynamic characteristics of the in- 
duced flow. 

The purpose of this paper is to provide ad- 
ditional insight into the question of rotor 
structural and induced flow modeling. To this 
end, experimental rotor frequency response data 
in hover and in forward flight are compared with 
theoretical results that are calculated by using 
several different models for the elastic blade 
bending and induced flow. The frequency response 
data provide a broad base of comparison so 'that 
the effects of mode shape and induced flow model 
can be clearly determined throughout the fre- 
quency range of interest. 



Basic Equations 



Analysis 



The mathematical technique used here is a 
further generalization of the harmonic balance 
approach of Reference 2. In addition to an arbi- 
trary number of bending modes (with an arbitrary 
number of azimuthal harmonics for each mode), the 
generalized harmonic balance allows for a 



rational treatment of reversed flow aerodynamics 
and the possibility of harmonically oscillating 
control inputs. 

The linear equation of motion for the de- 
flection of an elastic beam subject to distrib- 
uted aerodynamic and inertial loadings F(r,i)0 
and G(r,ijO is 6 

(EIw") + m« 2 w + n 2 (mrw' - w" f mr dr 1 



are obtained for the aj n and bj n . Solution of 
these equations, followed by a substitution of Eq. 
(5) into Eqs. (2) and (3), results in the phase 
and magnitude of all desired harmonics of the 
flapping deflections and hub forces and moments. 

Blade Loading 

The aerodynamic loading of each Blade is 
given by 



F(r,i|0 + G(r,<J;) 



(1) 



pac 



l u T Kv- u p) 



(6) 



The associated expressions for bending moment and 
shear at the blade root are 

S(0,*) » J (F + G - mfl 2 w - mfl 2 w)r dr (2) 



where 



U = S2r + fiRu sin <fi 



(7) 



V(0,*) 



,R 
J (F + G - mC 2 w)dr 



(3) 



U = Qw + SiRX + fiRuw 1 cos ii 



(8) 



The blade root bending moment is transformed into 
a stationary coordinate system to yield the pitch 
and roll moment of the rotor. The solution of 
Eq. (1) yields directly the blade deflections, 
and substitution into Eqs. (2) and (3) then 
yields the forces and moments. 

Application of the harmonic balance involves, 
first of all, an orthogonal expansion of w: 



R = 2v 
J-l 



qjOlOfjCr) 



(4) 



For the present analysis, the <f>* are taken to 
be the exact mode shapes of the rotating beam 
without aerodynamics. Galerkin's method is then 
used to transform Eq. (1) into J ordinary dif- 
ferential equations (with periodic coefficients) 
for the modal coordinates qj.^ When the forcing 
terms contain a steady portion superposed onto 
periodic functions that are modulated by an ex- 
citation frequency u (cycles per revolution) , 
Floquet's theorem implies that the qj have a 
solution of the form' 



q J = q 3 + 



jo + S [ a jn cos(n "' ) 



n=l 



+ b. sin(ni()) 



•] 



*«>* 



(5) 



where q. are the steady coning displacements 
and the aj n and bj n are complex quantities 
indicating the magnitude and phase shift of each 
modulated harmonic of the perturbation response. 
The harmonic balance approach entails substitut- 
ing Eq. (5) into the J ordinary differential 
equations for qj and setting coefficients of 
like harmonics equal. When n is truncated at 
the highest harmonic of interest N, then 
(2 • N + 1) • J linear algebraic equations 



Eq. (8) contains the primary contributions of mode 
shape and induced flow to the flapping equations. 
The details of blade mode shape become important 
as u increases because Up depends upon both 
the blade deflection w and its first derivative 
w" . The induced flow is important because first 
order perturbations to the inflow X create first 
order changes to Up and F. 

Although the inflow is in general a compli- 
cated function of radius and azimuth, as a first 
approximation, the total inflow can be represented 
by 



A + 



A + X 



f sin 



ijj + 



c R 



cos ifi e 



lWTp 



(9) 



The steady portion of the total inflow X con- 
tains contributions from the f reestream velocity 
V„/nR and from the steady induced flow due to 
rotor thrust v. The unsteady inflow components 
X ,X S ,X C contain contributions from harmonic 
plunging ze iu *, rolling ()>e iw *, and pitching 
ae"' of the shaft, as well as contributions 
from the unsteady induced flow components 
v ,v s ,v due to perturbations in rotor thrust and 
moments : 



X = -iwz + v 
o o 



-ioxji + v 
-iuio + v 



ua 



(10 a-c) 



c c 

The blade pitch angle 9 is given by 



= 6 + 



8 + 9 sin ifj + 8 cos ty 



iuif. 



(11) 



where 8 is the steady value of 6 and 
8 ,8 S ,8 are control system perturbations. The 
inflow perturbations X ,X g ,X are assumed to be 



small compared with unity. This implies that the 
induced flow perturbations v ,v s ,v c and the con- 
trol perturbations 9 ,6 s ,8 c ,z,<j>,a are also small 
quantities yielding linear perturbation equa- 
tions. 



by 



The inertial loading of each blade is given 



moments influence the induced flow. The induced 
flow, therefore, is a feedback loop of Eq. (15) , 
causing the uj to depend upon the f 4 . 

From the standpoint of calculation, it is con- 
venient to express the coupling relation (between 
the generalized controls, the physical controls, 
and the rotor response) in matrix form: 



G = -mfi 2 R[g + g s | sin * + g c | cos * e*" r (12) 



iuij, 



{u> = me*} + [K]{f> 



(17) 



where 



8 o 


— g 


U) 2 Z 




g s 


= 


to 2 * + 


2iua 


S c 


s 


-2iu<(> 


+ <o 2 a 



(13 a-c) 



The inertial loading is a result of centrifugal, 
Coriolis, and gyroscopic forces which occur in 
the rotating reference frame of the blade due to 
hub motions z,<(i,a in the inertial reference 
frame. 

When Eqs. (6) through (12) are combined and 
appropriately integrated in Eqs. (1), (2), and 
(3), the steady deflections and forces qj.Oj/aa 
are obtained as linear functions of the steady 
inputs 6, A; and the perturbation blade deflec- 
tions and hub forces and moments are obtained as 
linear combinations of the generalized control 
variables 



<u> =(ee e gggHi \ 

N o s c £> o°s B c o s c S 



(14) 



Although g ,g s ,g c are simply related to the 
shaft motion through Eq. (13), they are retained 
as generalized controls so that the generalized 
controls can be separated into physical, in- 
ertial, and aerodynamic groupings. This will 
facilitate the calculation of rotor response when 
induced flow is included later. 

Interpretation of Results 

The results of the harmonic balance can be 
expressed in matrix form as 



{f} - [M]{u} 



(15) 



where {f } represents the perturbation harmon- 
ics of thrust, moments, and generalized coordi- 
nates. The elements of [M] , therefore, have 
direct physical significance. They are the par- 
tial derivatives of each of the response harmon- 
ics taken with respect to each of the generalized 
controls uj. The generalized control variables 
are in turn functions of the physical controls 



<x> "(s 8 6 zAaS 

^ O S C r 



(16) 



Eq. (17) is simply a set of linear equations de- 
scribing: (i) the generalized control perturba- 
tions due to application of the physical controls 
[Y] and (ii) the generalized control perturba- 
tions due to the effect that rotor response has 
on the induced flow [K] . The matrices [Y] and 
[K] will be obtained later by using an appropri- 
ate induced flow model. It follows that the par- 
tial derivatives of the f j with respect to the 
physical controls -x.± can be found (including in- 
duced flow effects) from Eqs. (15) and (17). The 
derivative matrix is designated [M'] and has the 
properties 



{f} - [M']{xl 



[M'l = [[I] - [M][K] 



Mm 



(18) 
(19) 



Although the higher harmonics are often necessary 
in the harmonic balance calculation of [M] , the 
subsequent calculation of [M'] by Eq. (19) may 
be performed for only those response and inflow 
harmonics of interest. In this paper, five har- 
monics are used in the calculation of [M] , but 
only first harmonics are retained in Eqs. (18) and 
(19), so that the f*. are taken to be 



f j 



<f> / C T C L C M „ \ 

<f> = \«M« a JoVw 



(20) 



Induced Flow 



Form of Induced Flow Model 

A useful form of the induced flow model is 
given by* 



v o 

V 

s 

V 

. c 


Et 


L 


Cj/aa 
C L /aa 
C M /aa 



aerodynamic only 



(21) 



Although not completely general, Eq. (21) can 
accommodate a variety of induced flow models. 
Only aerodynamic contributions are included on 
the right-hand side, because they are the only 
loads which produce reaction forces on the rotor 
wake. Using Eqs. (2) and (3), these aerodynamic 
forces and moments can be expressed in matrix 
form as 



as evidenced in Eqs. (10) and (13). 



The generalized control variables uj axe 
also coupled to the fj, because the thrust and 



LC m /caJ 

1 I 

(C /era) 

J 1 

/C„/aa\ 
^ M ^aero- 
dynamic 



where 



iC T /aa, 
?C L /aa( 



0a 1 



V 



- [W] 






1 

- — m 
2 yy 



1 








-2 m yy 



J 1 

-Kj 






1 

2 yj. 



ft 

?! 

a ji 

(22) 



directly in Eq. (19) to obtain the complete rotor 
response to physical" control inputs . 

Unsteady Momentum Theory 

An approximation of the induced flow that is 
suitable for Eq. (21) can be obtained as an ex- 
tension of the momentum theory used in Reference 
2. The differential force on an elemental area of 
rotor disk is written as 



dF = 2£2Rvdm + fl 2 Rvdm 



(26) 



where 2QRv is the total change in velocity nor- 
mal to the disk, dm is the differential mass flow 
through the element, dm is the # apparent mass 
associated with the flow, and v is the time 
derivative of v in the nonrotating system. The 
differential mass flow relation 



w=- 





<u 2 
-2iu) 




2i<u 

U, 2 



(23) 



With the induced flow v described by Eqs. (21) 
and (22), the inflow relation follows directly 
from Eqs. (10), (13), and (21). The matrices 
[Y] and [K] of Eq. (17) may then be identified 
as 



m - 



C"j, 



°3*3 [t] [W] 





Ky 
o 








iu 





y 





- 





lu 





Hy 










iu 



(24) 



[K] 



3*3J 



ai, 



-[L] [W] 



J 1 







Hi 



J J 3*3J 



(25) 



Eq. (24) represents the control coupling be- 
tween the physical controls Xj and the general- 
ized controls u-j. The presence of [L] in this 
matrix indicates that the X's are indirectly 
coupled (through the induced flow), as well as 
geometrically coupled [Eqs. (10) and (13)] to the 
rotor plunge, pitch, and roll motions. Eq. (25) 
represents the induced flow caused by the de- 
pendence of X upon the thrust and first harmon- 
ic flapping. If a suitable approximation to the 
inflow can be modeled in the form of Eq. (21), 
then Eqs. (24) and (25) may be substituted 



dm = pflR Vu z + A 2 r dr diji 



(27) 



can be used to integrate the first term of Eq. 
(26) over the disk to obtain a quasi-steady in- 
duced flow relation for rotors that have combined 
conditions of thrust and forward speed. The eval- 
uation of the second term in Eq. (26) (the un- 
steady effect) requires the additional knowledge 
of the apparent mass dm associated with the flow. 

An approximation to the apparent mass terms 
of a lifting rotor can be made in terms of the 
reaction forces (or moments) on an impermeable 
disk which is instantaneously accelerated (or ro- 
tated) in still air. This approximation was used 
in Reference 8, giving good agreement with trans- 
ient thrust measurements for an articulated rotor. 
The reactions on such an impermeable disk are 
given from potential flow theory in terms of el- 
liptic integrals which are evaluated in the lit- 
erature. 9 They result in apparent mass and in- 
ertia values 



m A = f pR3 , 



^-if* 8 



(28) 



(For 



v_ r/R, a radial velocity distribution, 



m A = pR 3 .) These values represent 64 percent of 
the mass and 57 percent of the rotary inertia of 
a sphere of air having radius R. It is empha- 
sized that they are only approximations to the 
actual values for a lifting rotor. 

Using this approximation, the steady induced 
flow equation and the unsteady induced flow per- 
turbation equations can be derived from Eqs. (26) 
through (28): 



2v \/u 2 + X 2 - C T 
Vo + 2vV o " C T 
Kl v s +|vv 8 --C L 



(29a) 



(29b-d) 



where 



v= u 2 + X(X + v) 



vV + X 2 



(29e) 



and 



K = __A_ = JL = 0.8488 
m pnR 3 3ir 

A 16 
^ s — &- = -^2- = 0.1132 



(30a-b) 



pirR 5 45ir 



Eq. (29a) expresses the nonlinear relation be- 
tween the steady thrust and the steady induced 
flow v. Eqs. (29b-d) are then the linear per- 
turbation equations for small changes in thrust, 
moments, and induced flow. In order for the per- 
turbation equations to be valid, it is assumed 
that v ,v s ,v c are much smaller than Cli 2 + X 2 )\ 

The time constants associated with the induc- 
ed flow model in Eq, (29) are 



(24) , (25) , and (19) to obtain the rotor response 
that includes inflow. 

Empirical Model 

Experimental data have shown that momentum 
theory, although particularly simple to use, is 
qualitatively inaccurate for certain steady re- 
sponse derivatives in forward flight. Reference 
2 introduces an alternate induced flow model for 
forward flight in which the elements of [L] 
(with u = 0) are chosen to give the best fit of 
experimental response data for several configura- 
tions at conditions of near zero lift. If this 
empirical inflow model, [Lg] , is taken for the 
quasi-steady portion of the induced flow law, and 
if the theoretical apparent mass terms (from po- 
tential flow) are taken as a model for the un- 
steady portion of the induced flow law, then a 
complete induced flow equation can be expressed as 



1_ 
oa 



K 



"0 



m 

-K. 



I 
-K, 




(33) 



t T = -^ = 0.4244/v (for v q ) 



2K I 



(31a-b) 



T = = 0.2264/v (for v ,v ) 

S v s' c 



In Reference 4, the steady induced flow v and 
the time constant for v s ,v c are obtained by 
correlating experimental hover frequency response 
data. Two operating conditions are considered, 
and the best fit in these cases is found to be 
v - .014, t s = 8 (with 8 = 2°) and v = .028, 



l s 



4 (with 8=8°). From the 



values indi- 



cated for these cases, it can be shown that each 
t s implies the same value of Kj = 0.112. Thus, 
there is some experimental evidence that the po- 
tential flow value 
valid . 



Kj; = 0.113 is approximately 



By assuming simple harmonic motion, Eqs. 
(29b-d) can be brought into the form of Eq. (21) , 
yielding the components of [L] for unsteady 
momentum theory. 



<3& 



[L] 



2v + K iw 



-era 



v/2 + Rjiio 




v/2 + ICj.iu) 



(32) 



(L22 ana L 33 differ by a factor of 4/3 from Ref- 
erence 2, because v s and v c are taken uniform 
with r in that reference, whereas they are 
taken linear with r here.) The matrix [L] 
from Eq. (32) may now be substituted into Eqs. 



The assumption that the apparent mass terms 
may be superposed on the quasi-steady terms is not 
rigorous, but it can be considered analogous to 
unsteady wing theory in which the apparent mass 
terms are theoretically independent of the free- 
stream velocity. Under the superposition assump- 
tion, the empirical inflow model modified for the 
unsteady case is 



[L] 



m 











~ K T 











-K 



I- 1 



aa 



I 1-1 

L, 



-1 



(34) 



Although this particular formulation of [L] is 
valuable for predicting the effects of induced 
flow, ultimately a more consistent formulation of 
[L] should be made, as discussed in Reference 10. 

Equivalent Lock Number 

Another method of accounting for the unsteady 
induced flow is the use of an equivalent Lock num- 
ber y*, which can be derived from a single har- 
monic balance of the root moment equation: 



1 - 



1 + 8v/oa + 16K Waa 



(35) 



Although this approach is not a completely con- 
sistent treatment of the induced flow, since it 
does not give an exact harmonic balance of the 
blade flapping and thrust equations, it yields 
results which are nearly the same as those ob- 
tained from momentum theory. 



The practical use of Eq. (35) is somewhat 
limited because of the inaccuracies of momentum 
theory in forward flight, but a y* approach is 
nevertheless a valuable conceptual tool for under- 
standing the effects of induced flow. In particu- 
lar, Eq. (35) shows that one effect of induced 
flow perturbations is to decrease the effective 
Lock number (i. e. , decrease the aerodynamic ef- 
fectiveness) . This decrease is most pronounced at 
low values of v (i.e., low u and 8 ) and low 
values of w. For example, rotor roll moment is 
plotted in Figure 1 for two values of and com- 
pared with the value from elementary theory 
(steady induced flow only, induced flow perturba- 
tions neglected, equivalent to lim 8 •*■ °° )'. The 
curves for 8 " 0, 0.05 result in values of roll 
moment well below the elementary value. 



.020 


NO INDUCED FLOW 
PERTURBATIONS \^ 


8=oo 


.05 





.015 










.010 


/ / 








.005 


&£^ X 1 




I 


1 



.2 .3 

Sst rod 



.4 



Figure 1. 



Effect of induced flow on steady rotor 
response in hover, u = 0, m = 0, 
a ■ 0.1, a = 2ir, p » <*>. 



The effect of induced flow is most pro- 
nounced in the response derivative (the slope of 
the response curve at 8 S • 0) . For p = °°, the 
derivative is given by 



sfcj/oaj/i 



86 



8 =0 
1 s 



_1_ 
16 



y 



(l + 3/2 u 2 ) (36) 



indicating that y*/y < 1 results in a reduction 
of the roll moment response (or control power) 
from the elementary value. When the rotor is in 
hover with no lift (v » 0) , a quasi-steady per- 
turbation of 8 S (u » 0) results in no response 
because of the zero slope of the curve in Figure 
1. The mathematical justification for the van- 
ishing response derivative can be seen in Eqs. 
(35) and (36). With u - v - 0, y*ly and the 
response derivative must equal zero. As 9" in- 
creases, however, v and v increase so that 
Y*/y approaches unity and the derivative ap- 
proaches -1/16, as illustrated in Figure 2. With- 
in the practical range of thrust coefficients, 
however, the response derivative never recovers 
more than about 80 percent of the elementary 
value. Eq. (35) also implies that increasing ad- 
vance ratio (which increases v) will result in a 
partial recovery of y*/y (and of the response 
derivative). This recovery is evident in 



-.08 






NO INDUCED FLOW 
•PERTURBATIONS^,^ -_„ 


-.06 




-.04 


' ^^-~~~~~ 


-.02 


i i — i 1 



.2 



.4 





1 


.05 


V 


.10 
I 


.12 

i 



Figure 2. 



C T Ar 

Eff ect of induced flow on steady rotor 
response derivatives in hover, u = 0," 
to = 0, = 0.1, a = 2ir, p = ». 



Figure 3, where the roll response is given versus 
u; but no more than 90 percent of the elementary 
value is reached in the practical range of thrust 
and advance ratio. 



-.10 




NO INDUCED FLOW -^ 
PERTURBATIONS \^"^ 


-.08 








F = co — 




£ -.06 


■ 15 — 




£-.04 
*> 


- .05^^ 

0/ 




-.02 


i 


i i i i i 



.1 .2 .3 .4 .5 .6 

Figure 3. Effect of induced flow on steady rotor 
response derivatives in forward 
flight, w • 0, a = 0.1, a « 2ir, p » ». 

The unsteady terms (apparent inertia K_) 
also bring y*/y closer to unity, as seen by the 
role of Kx in Eq. (35). This recovery with fre- 
quency is illustrated in Figure 4, where, as us 
becomes large, the response derivative approaches 
the elementary value of -1/16. The rate at 
which the response approaches -1/16 is dependent 
upon the magnitude of the apparent inertia Kj. 
Large values of Kx result in a rapid return to 
the elementary value, and small values of Kx re- 
sult in a slow return. For Kx " 0.1132 and 
u < 0.3, the unsteady terms provide only small 
contributions to the response. Thus, the quasi- 



4 o 

Sf 200 



.ua 


NO INDUCED FLOW 
K I = O / PERTURBATIONS 


.06 
.04 


/ .M32_^^ ^ . — ' 


.02 


\. QUASI-STEADY 
INDUCED FLOW 

1 I i i i i 



.2 



.8 



1.0 



1.2 




132 



-£. 



0, oo 



.2 



.4 



.8 



1.0 



1.2 



Figure 4. 



Effect of induced flow time constant on 
rotor frequency response derivatives, 
u = 0, a = 0,1, a = 2ir, p = =°, 
v = X = 0.05. 



steady theory (with Ki = 0) would be adequate in 
this range. In the frequency range 0.3 < u> < 
1.0, the unsteady terms have a more significant 
effect. Above m = 1.2, the total effect of in- 
duced flow diminishes so that the elementary 
theory and the unsteady theory give similar re- 
sults; but the quasi-steady theory (with Ki = 0) 
is in considerable error in this region. 

The frequency range in which unsteady in- 
duced flow is important_is also dependent upon the 
thrust or mean inflow v as shown in Figure 5. 
For low values of v, the unsteady effects domi- 
nate at low frequencies; and for large values of 
v, the unsteady effects are delayed into the 
higher frequencies. This effect is implicit in 
Eq. (35) and is a direct result of the inverse 
dependence of time constant upon v, Eq. (31) . 
Thus, a low v implies a slow induced flow re- 
sponse; and a high v implies a rapid induced 
flow response. Equation (35) shows that advance 
ratio (which also increases v) • has a similar ef- 
fect on the induced flow behavior. It follows 
that the relative importance of the unsteady and 
quasi-steady nonuniform induced flow terms de- 
pends upon both the rotor operating conditions 
and the frequency range of interest. 



.08 



.06 



-, .04 



.02 - 







f = 00 


^N0 INDUCED FLOW PERTURBATIONS 


^=r^^^^^ 


' .05 y 




7 


iii., 




Figure 5. Effect of unsteady induced flow on 

rotor frequency response derivatives, 
u = 0, a = 0.1, a •* 2ir, p = ■», 
Kj. = 0.1132. 



NO INDUCED FLOW 

OR 

QUASI -STEADY INDUCED FLOW 

OR 

UNSTEADY INDUCED FLOW 



HIGH /J. 



MODERATE p. 
LOW/Lt i 

OR HIGH ~T~ 
C T /<r , 

VERY LOW jJl 
AND LOW _!_ 
C T /o- 




O .5 

ROTOR /FUSELAGE DYNAMICS 



NO 

INDUCED 

FLOW 

OR 

UNSTEADY 

INDUCED FLOW 

1.5 2.0 

BLADE DYNAMICS 



In Figure 6, the relative importance of 
these terms is presented qualitatively through a 
chart of the operating regimes in which (for no 
induced flow or quasi-steady induced flow) |y*j 
differs by less than 10 percent from the unsteady 
value. This is a subjective criterion and is 
merely intended to illustrate the trends with 
thrust, advance ratio, and frequency. Four 
regions are defined: (i) at high u and v, 
induced flow effects are small and either the 
elementary or quasi-steady approximation is ade- 
quate; (ii) at high u and low v, although in- 
duced flow effects are small (no induced flow 
being a good approximation) , the quasi-steady 



Figure 6. Regions of validity for steady (no 
induced flow perturbations) quasi- 
steady (Kj = K = 0) , and unsteady 
(K x = 0.1132, \ = 0.8488) induced 
flow models based on y*» Eq. (35), 
a = 0.1, a = 2ir. 

theory alone will be in error; (iii) at low id 
and high v, the opposite is true (i.e., the 
quasi-steady nonuniform theory is required, where- 
as neglecting induced flow results in error) ; and 
(iv) for low a) and v, complete unsteady theory 
is required. 



Comparison of Theory and Experiment 

the experimental data used in the following 
correlations were obtained with a 7 . 5-f t-diameter 
hingeless rotor model tested in the USAAMRDL-Ames 
wind tunnel. •"• The model configuration and test 
conditions covered a wide range of parameters. 
The results included here are for p = 1.15 and 
advance ratios from 0.0 to 0.6. 

Elastic Blade Bending 

In Fig. 7, experimental values of roll and 
pitch moments due to 6 S are compared with theo- 
retical results which are calculated neglecting 
induced flow perturbations. Two sets of theory 
are presented. The first theory employs a rigid 
centrally-hinged blade with root spring to model 
the elastic blade bending, and the second theory 
[uses a similar model, except that hinge offset is 
llowed. The largest differences between the two 
heories occur near resonant frequencies, i.e., 

0.15, 1.15. (The primary effect of mode 
ihape is aerodynamic, Eq. (8); it causes domi- 
nce at resonance.) A surprising element in 
figure 7 is that the centrally hinged model gives 
closer agreement with the high frequency response 
than does the hinge offset model. This reversal, 
however, is not a consistent trend in the data 
and may be somewhat coincidental. 



'[cmAoJ/W, 




Figure 7. Comparison of experimental data with 
rigid blade approximations without 
induced flow, p •> 1.15, y = 4.25, 
B = 0.97, e pc = 0.25, u - 0.60. 

Similar frequency response comparisons have 
been made when the blade is modeled by one or two 
of the rotating elastic mode shapes. When 
u < 0.8 and <d is at least once-per-revolution 
below the second flap frequency, the one- and two- 
mode calculations are within a few percent of the 
hinge-offset results. At higher advance ratios 
and frequencies, the effects of second-mode bend- 
ing can become significant; but in the range of 
operating conditions considered here, a single ro- 
tating mode is sufficient to model the blade. 

Three major types of discrepancies between 
theory and experiment which are found in Figure 7 



cannot be explained in terms of flapping mode shape 
effects. The first is the difference encountered 
at frequencies near one and two per revolution. 
This difference may be explained by the fact that 
the lead-lag frequency of this configuration is 
near two per revolution, causing resonance at these 
frequencies. The second discrepancy is the irregu- 
larity in the pitch response at id = 0.6. Here, a 
natural frequency of the rotor support stand is 
being excited and contaminates the data.-'-- 1 - The 
third discrepancy is found at u < 0.6, and will be 
shown to result from unsteady inflow perturbations. 

Effect of Induced Flow In Hover 

The low-frequency hover data provide some in- 
sight into the effects of unsteady induced flow. 
In Figure 8, rotor roll and pitch moments versus 
9 g are presented. The experimental results are 
for 5 = A", v = 0.03. The theoretical results are 
calculated using the actual blade rotating mode 
shape as a generalized coordinate and using three 
different representations of the induced flow. The 
first representation is the elementary model, which 
completely neglects induced flow perturbations. 
The second representation is quasi-steady momentum 
theory, which neglects the apparent inertia 
(Kj «■ 0) , assuming that nonuniform induced flow 
perturbations instantaneously follow the blade dy- 
namics. The third representation is unsteady mo- 
mentum theory, which gives a time lag on the in- 
duced flow perturbations. (The empirical model is 
not applicable in hover.) 

A comparison of theory and experiment reveals 
that the elementary theory is unsatisfactory below 
w = 0.6, failing to reproduce even the qualitative 
character of the data. On the other hand, the 
theories which include induced flow perturbations 
account for most of the important features of the 
response. The loss of aerodynamic effectiveness, 
which is a result of induced flow perturbations, 
causes a decrease in the excitation forces and an 
overall decrease in the response. But the loss of 
aerodynamic effectiveness also lowers the blade 
damping, causing a resonant peak effect near the 
blade natural frequency (with p = 1.15, w = 0.15). 

The effect of the unsteady induced flow terms 
is also evidenced in Figure 8. The major contri- 
bution of Kx is the determination of how rapidly 
with on the aerodynamic effectiveness returns to 
the elementary value. Above w = 0.6, the theo- 
retical value of Kj gives the proper amplitude 
and phase for the hub moments, while the quasi- 
steady theory (Kj * 0) fails to return to the con- 
ventional value and does not agree with the data. 
Below to = 0.6 the comparison is less clear. In 
the roll-moment phase and amplitude, a Kj less 
than 0.1132 would give better correlation than 
does this theoretical value. In the pitch-moment 
response, however, a smaller Kj would give worse 
correlation than does Kj = 0.1132. Further work 
would be necessary to determine if this effect is 
due to experimental difficulties (such as recir- 
culation) or to an actual deficiency in the in- 
duced flow model. 



.06 



.04 



_i 



3.02 



360 



v 

■a 

i 



180 



d[c L /a-a]/38 s 



.06 



O DATA REF II 

NO INDUCED FLOW 

QUASI-STEADY INDUCED FLOW .04 

UNSTEADY INDUCED FLOW 

(Kl- 0.1132, K„. 0.8488) 



a[c„/<ro]/as, 




Figure 8. Rotor response to cyclic pitch in 

hover, p =1.15, y = 4.25, B - 0.97, 
e pc = 0.25, u = 0, era « 0.7294, 
v «■ X = 0.03, momentum theory, single 
rotating mode. 

In Figure 9, rotor roll and pitch moments 
versus a are presented for the same test condi- 
tions as in Figure 8. Data are presented for 
shaft excitations in both roll and pitch, since 
in hover the response to these controls is 
ideally symmetric. A comparison of the two sets 
of data gives an indication of the experimental 
error due to test stand dynamics (and possibly 
recirculation) . Although the data are question- 
able for w > 0.3, the lower frequency data sub- 
stantiate three of the observations made from 
Figure 8. First, the elementary theory is quali- 
tatively inaccurate for amplitude and phase re- 
sponse. Second, a major effect of induced flow 
is a resonant peak effect near os » 0.15. Third, 
Ki < 0.1132 would give better correlation than 
the theoretical value at low w. Figure 9 also 



.06 



i[c,_/<ro]/ia 
O DATA REF II 
D DATA REF II (ROLL I 
NO INDUCED FLOW 



.06 



ra]/da 




Figure 9. Rotor response to hub motions in hover, 
p - 1.15, y - 4.25, B - 0.97, 
e pc ■ 0.25, u » 0, aa « 0.7294, 
v * X ■ 0.03, momentum theory, single 
rotating mode. 



shows that although induced flow decreases the 
blade damping, it can actually increase the rotor 
pitch/rate damping = -Re[3(C M /aa)/3ot] and also 
increase the rotor pitch/roll coupling 
= -Re[3(CL/aa)/3o]. The damping and coupling can 
be found by dividing the plotted curves by 
-iuj, A = iwa, which is approximately equivalent to 
taking the slope of the plotted curves with a 90- 
degree shift in phase angle. For this particular 
configuration, the damping and coupling are in- 
creased by induced flow effects, indicating that 
induced flow perturbations can be important in 
coupled rotor /fuselage dynamics. 

Effect of Induced Flow in Forward Flight 

In the next three figures, experimental data 
at high-advance ratio (u= 0.51) and very low lift 
(6 = 0.5°) are compared with theory using three 
induced flow descriptions. The first description/ 
is an analysis which neglects Induced flow perturf 
bations, the second description is the empirical 
model of Reference 2 with no time lag (quasi- 
steady, Kj_ ■ Kji = 0), and the third description j 
is the empirical model of Reference 2 adapted to : 
the unsteady case according to Eq. (34) (with the 



»[c L /o-o]/a8, 



afCM/o-aJ/se,, 




Figure 10. Rotor response to collective pitch in 
forward flight, p - 1.15, y - 4.25, 
B - 0.97,, e pc - 0^25, p - 0.51, 
era - 0.7294, v - X ■ 0, single 
rotating mode. 

theoretical values of Kj_ and K m ). The first 
comparison of theory and experiment is shown in 
Figure 10 for the roll- and pitch-moment response 
due to 6 . The elementary theory predicts a 
roll moment of 0.017 at w *• and a near-zero 
crossing (amplitude ■ 0, phase angle discontinu- 
ous) at co ■= 0.4. The data, however, displays a 
much lower steady value and completely avoids the 
zero crossing. The unsteady and quasi-steady 
empirical models provide a fairly accurate 
description of this behavior, showing quantita- 
tive agreement with phase and magnitude for 
a) < 0.6. For the pitch moment derivative, the 
empirical models predict the qualitative (but not 
the quantitative) aspects of the reduction in 
moment (from the conventional value) due to 
induced flow. 



10 



Another comparison of theory and experiment 
is shown in Figure 11 for the roll- and pitch- 
moment response due to S . The empirical models 
predict a roll-moment derivative which is less 
than the elementary value, exhibiting a near-zero 
crossing at a = 0.26. This characteristic is 
clearly evident in the magnitude and phase of the 
data, but it does not appear in the theory 




Figure 11. Rotor response to longitudinal cyclic 
pitch in forward flight, p = 1.15, 
Y = 4.25, B = 0.97, e pc = 0.25, 
u = 0.51, aa = 0.7294, v = X = 0, 
single rotating mode. 

without induced flow. For the pitch-moment deriva- 
tive, the elementary theory agrees with the data 
only for u > 1.2; the quasi-steady theory shows 
good correlation for < u < 0.6, and the unsteady 
theory gives quantitative correlation at all fre- 
quencies. 

The third comparison is shown in Figure 12 for 
the roll- and pitch-moment response due to 8 C . 
The data show that the roll-moment derivative is 
less than the elementary value at • u «■ 0, display- 
ing a resonant peak (hear oj = 0.15) which is 
greater than the elementary value and which is ac- 
companied by a 10-rdegree phase shift. The empiri- 
cal models predict the qualitative character of the 
resonant peak and quantitative character of the 
phase shift. The empirical models also correlate 
well with the pitch-moment response, for which the 
experiment shows the derivative to be greater than 
the elementary value for « < 0.3 and less than 
the elementary value for u > 0.3. 

In general, the empirical inflow models show 
this same degree of correlation at all advance 
ratios considered (u » 0.27, 0.36, 0.51, 0.60). 
This substantiates one of the qualitative 
conclusions of Figure 6. For moderate advance 
ratios and w < 1.0, an appropriate unsteady or 
quasi-steady induced flow theory is adequate, but 
the theory without induced flow is in consider- 
able error. Of course, Figure 6 only implies in 
which regions quasi-steady or unsteady terms may 
be significant. It does not Imply that any par- 
ticular quasi-steady or unsteady model will be 



.06 
g.04 

| 

3.02 



360 



j[c L /«i]/ae c 

O DATA REF II 

* NO INDUCED FLOW 

QUASI-STEADY EMPIRICAL MODEL 

UNSTEADY EMPIRICAL MODEL 

(Kj-0.1132, K m »0.84B8) 



.06 


*fc M /<™]/*>C 


04 




.02 






^V-^ 




. . . . *r - Bv r= -?"7 . °. . , 



,-180 



1.2 
360 

180 



1.2 



.4 .8 
u 



1.2 



.4 .8 



\2 



Figure 12. Rotor response to lateral cyclic 

pitch in forward flight, p = 1.15, 
Y = 4.25, B - 0.97, ep C = 0.25, 
U = 0.51, aa » 0.7294, v = A = 0, 
single rotating mode, 
adequate. For example, in Figure 13, pitch mo- 
ment derivates (as calculated using the theory 
without induced flow, unsteady momentum theory, 
and unsteady empirical theory) are compared with 
the experimental data. The comparison shows that 
unsteady momentum theory can be in qualitative 
disagreement with the data even though empirical 
theory shows good correlation. Even the empiri- 
cal model, however, does not show complete quan- 
titative correlation; and further refinements in 
the induced flow model may be necessary. 

Conclusions 

1. On the basis of an equivalent Lock number 
relation and p = °°, quasi-steady nonuniform in- 
duced flow perturbations can have a significant 
effect on rotor response throughout the entire 
thrust /advance ratio range; but the time lag of 
the induced flow is only important at low lift and 
low advance ratio. 

2. In hover, unsteady momentum theory with appar- 
ent mass terms from potential flow provides a 
significant improvement in data correlation over 
the theory without induced flow perturbations; 

but further work is required to refine the induced 
flow model. 

3. In forward flight and near-zero lift, the 
empirical inflow model of Reference 2, whether 
used with the unsteady time-lag effect or with- 
out the time-lag effect (quasi-steady) , corre- 
lates well with most qualitative and some quanti- 
tative aspects of the data, while unsteady momen- 
tum theory and the theory without induced flow 
provide little agreement with the data. 

4. A single rotating mode is sufficient for 
flapping response calculations when u < 0.8 and 
when the major excitation frequency is at least 
once-per-revolution below the second flapping 
frequency. 



11 



.03 



kj .02 
o 

t 

_J 

s 

« .01 



a[c L /o-a]/ae 



.03 



a[c L /o-a]/ae 5 



O DATA REF 10 

NO INDUCED FLOW 

MOMENTUM THEORY, g - 02 

UNSTEADY =J 

■EMPIRICAL MODEL, g 

UNSTEADY S 

Ov\ < -o 






.03 



uj .02 



.01 



a[c L /o-a]/ae c 




.2 .4 



360 




en 




XJ 




lj" 180 


■ CU3 =^et =: 


< 


^^ == ^^ = ~ :=:: =Q~. 


I 




0. 


1 1 1 1 1 1 



Figure 13. Effect of induced flow model on low frequency, roll response, p = 1.15, Y = 4.25, B = 0.97, 



e = 0.25, \i = 0.51, oa = 0.7294, V = X - 0, K T 
pc I 



0.1132, K. = 0.8488, single rotating mode. 



References 

1. Shupe, N. K. , "A Study of the Dynamic Motions 
of Hingeless Rotored Helicopter," PhD. Thesis, 
Princeton Univ. 

2. Ormiston, R. A. and Peters, D. A., "Hingeless 
Rotor Response with Nonuniform Inflow and 
Elastic Blade Bending," Journal of Aircraft , 
Vol. 9, No. 10, October 1972, pp. 730-736. 

3. Hohenemser, K. H. and Yin, Sheng-Kwang, "On 
the Question of Adequate Hingeless Rotor 
Modeling in Flight Dynamics," 29th Annual 
National Forum of the American Helicopter 
Society , Preprint No. 732, May 1973. 

4. Crews, S. T., Hohenemser, K. H. , and Ormiston, 
R. A., "An Unsteady Wake Model for a Hinge- 
less Rotor," Journal of Aircraft , Vol. 10, 
No. 12, December 1973. 

5. Potthast, A. J., "Lockheed Hingeless Rotor 
Technology Summary - Flight Dynamics", Lock- 
heed Report LR 259871, June 1973, p. 43. 



9. 



Bisplinghoff , R. L., Ashley, H. , and Halfman, 
R. L. , Aeroelasticity , Addison-Wesley, Read- 
ing, Mass., c. 1955. 

Peters, D. A. and Hohenemser, K. H. , "Appli- 
cation of the Floquet Transition Matrix to 
Problems of Lifting Rotor Stability," Journal 
of the American Helicopter Society , Vol. 16, 
No. 2, April 1971, pp. 25-33, 

Carpenter, P. J. and Fridovich, B., "Effect of 
Rapid Blade Pitch Increase on the Thrust and 
Induced Velocity Response of a Full Scale 
Helicopter Rotor," NACA TN 3044, Nov. 1953. 

Tuckerman, L. B. , "Inertia Factors of Ellip- 
soids for Use in Airship Design," NACA Report 
No. 210, 1925. 



10. Ormiston, R. A. , "An Actuator Disc Theory for 
Rotor Wake Induced Velocities," presented at 
AGARD Specialists' Meeting on the Aerodynam- 
ics of Rotary Wings, September 1972. 

11. Kuczynski, W. A. , "Experimental Hingeless 
Rotor Characteristics at Full Scale First 
Flap Mode Frequencies," NASA CR 114519, 
October 1972. 



12 



DYNAMIC STALL MODELING AND CORRELATION WITH 
EXPERIMENTAL DATA ON AIRFOILS AND ROTORS 

R. G. Carlson, Supervisor 

R. H; Blaekwell, Dynamics Engineer 

Rotor Dynamics Section 

Sikorsky Aircraft Division of United Aircraft Corporation 

Stratford, Connecticut 

G. L. Commerford, Research Engineer 

Aeroelastics Group, Fluid Dynamics Laboratory 

United Aircraft Research Laboratories 

East Hartford, Connecticut 

P. H. Mirick, Aerospace Engineer 

U. S. Army Air Mobility Research and Development Laboratory 

Fort Eustis, Virginia 



Abstract 

Two methods for modeling dynamic stall have 
been developed at United Aircraft. The a, A, B 
Method generates lift and pitching moments as 
functions of angle of attack and its first two 
time derivatives . The coefficients are derived 
from experimental data for oscillating airfoils. 
The Time Delay Method generates the coefficients 
from steady state airfoil characteristics and an 
associated time delay in stall beyond the steady 
state stall angle. Correlation with three types 
of test data shows that the a, A, B Method is 
somewhat better for use in predicting helicopter 
rotor response in forward flight . Correlation 
with lift and moment hysteresis loops generated 
for oscillating airfoils was good for both models. 
Correlation with test data in which flexibly 
mounted two-dimensional airfoils were oscillated 
to simulate the IP pitch variation of a helicopter 
rotor blade showed that both methods overpredicted 
the response, and neither gave a clear advantage. 
The <*, A, B Method gave better correlation of 
torsional response of full scale rotors and re- 
mains the method in general use. The Time Delay 
Method has the potential to be applied more easily 
and probably can be improved by consideration of 
spanwise propagation of stall effects . 



Stall-related phenomena limit the operation- 
al capabilities of the helicopter. Power, blade 
stress , and control system loads can all increase 
substantially due to blade stall. To predict 
such phenomena unsteady aerodynamics in stall must 
be modeled in blade aeroelastlc analyses. A num- 
ber of unsteady aerodynamic models have been 
developed. These include methods described in 
References 1 and 2. Reference 3 is a recent 
general survey article of rotor dynamic stall. 
The a, A, B Method and the Time Delay Method are 
two methods developed by United Aircraft. The 
a, A, B Method was developed to use airfoil test 
data obtained for a sinusoidally oscillating 

Presented at the AHS /NASA-Ames Specialists' 
Meeting on Rotorcraft Dynamics, February 13-15 » 
191b. 



Based on work performed under U. S. Army Air Mobil- 
ity Research and Development Laboratory Contract 
No. DAAJ02-72-C-0105 . 



two-dimensional model airfoil. The Time Delay 
Method was developed to provide an empirical method 
that would agree with the lift and pitching moment 
hysteresis characteristics measured in oscillating 
airfoil tests for a number of airfoils and test 
conditions . 

Evaluation of unsteady aerodynamic modeling 
techniques generally proceeds from correlation with 
data obtained in two-dimensional oscillating air- 
foil tests to correlation of full scale rotor blade 
torsional response. Two-dimensional rigid airfoil 
results are compared on the basis of aerodynamic 
pitch damping and lift and pitching moment hys- 
teresis loops , and full scale correlation is judged 
on the basis of agreement in blade torsional mo- 
ments or control rod loads . Evaluation of an un- 
steady model on the basis of full scale torsional 
response is made difficult by uncertainties in 
three-dimensional rotor inflow and blade bending 
and plunging motion. Correlation of the lift and 
pitching moment time histories of rigidly driven 
airfoils, on the other hand, is not the best method 
of comparison because it does not treat blade dy- 
namic response to stall. As an intermediate ap- 
proach, model test data were obtained for a flex- 
ibly mounted model airfoil which was dynamically 
scaled to simulate the dynamics of the first tor- 
sional mode of a rotor blade. This paper summa- 
rizes unsteady aerodynamic modeling techniques and 
includes comparisons based on two-dimensional aero- 
dynamic pitch damping, lift and pitching moment 
hysteresis loops , two-dimensional flexured airfoil 
response, and full scale rotor blade torsional 
moments . 

Description of the Unsteady Models 

a, A, B Method 

In the a, A, B method the aerodynamic moment 
is assumed to be a function of angle of attack and 
its first two time derivatives . Reference 1* demon- 
strated that unsteady normal force and moment data 
generated during sinusoidal airfoil tests and tabu- 
lated as functions of a, A = ba a nd B = b 2 *qj 

u u?r 

(where b is the airfoil semi-chord and U is the 
free stream velocity) could be used to predict the 
aerodynamic response of an airfoil executing 



13 



a nonsinusoidal motion. In a limited number of 
flexured airfoil tests described in Eeferenoe k, ' 
good correlation was achieved between measured 
and predicted airfoil dynamic response. The a, 
A, B lift and pitching moment data tabulations 
of Heference h were used in the calculation of 
torsional response for the dynamically scaled 
model airfoil. As applied in this investigation, 
two changes were made in the calculation. First, 
to consider the pitch axis of the model airfoil as 
a variable, provision was made to include pitching 
moment due to chordwise offset of the aerodynamic 
center from the pitch axis: 



c m (a,A,B)=c m (a,A,B)+(Xp A - 

7 cA 

X PA 



.25)c 1 (a,A,B) 



The second change involved scaling the un- 
steady data tables to account for differences in 
wind tunnel characteristics. The steady state 
lift and moment data for the present test program 
differed from the corresponding steady data 
obtained in Heference h because the tests were 
conducted in an open jet wind tunnel and because 
the airfoil effective aspect ratio was much 
higher. The method of scaling used for these 
analyses required a shift in the entire data tabu- 
lation by constant values of angle of attack, un- 
steady lift coefficient, and unsteady moment co- 
efficient according to the following relations : 



C l (a > A > B) open jet =c l (a+5( V A > B) TAB + 



6ci 



and 

c (a,A,B) . =c (a+6a ,A,BL.„+ 6e 
m '. 'open jet m m' ' TAB m 

The constants Sa^, 60^, 6c-|_ and S^ were es- 
tablished for each airfoil and were equal to the 
amount of shift necessary to make the open jet 
steady state stall points in lift and moment 
match the steady state stall points of the 
airfoil of Reference h. 

Time Delay Unsteady Model 

. Wind tunnel airfoil dynamic response was 
also calculated with the Sikorsky Time Delay un- 
steady aerodynamic method. This formulation was 
developed empirically by generalizing the re- 
sults of a set of oscillating airfoil test pro- 
grams. It is intended to predict the unsteady 
aerodynamic characteristics of arbitrary airfoils. 
Its aim is to provide the blade designer with un- 
steady lift and pitching moment characteristics 
of various airfoils without conducting extensive 
oscillating airfoil tests. This model, based on 
a hypothesis of the physical separation process , 
does not depend on an assumed harmonic variation 
of angle of attack. The basic assumption is that 
there exists a maximum quasi-static angle of 
attack at which the pressure distribution and the 
boundary layer are in equilibrium. During in- 
creases in angle of attack beyond this static 
stall angle, there are finite time delays before 
a redistribution of pressure causes first a moment 



break and then a loss of lift corresponding to flow- 
separation. The relative phasing of the moment and 
lift breaks with angle of attack produces either 
positive or negative damping of the motion. 

To test the Time Delay hypothesis , harmonic 

data from Reference 5 were examined. It was noted 

that the onset of stall can occur before, with, or 

after maximum amplitude of the oscillation. In 

accordance with the Time Delay hypothesis, the 

spread between the static moment stall angle and 

the dynamic lift break was evaluated in terms of 

elapsed time nondimensionalized by free. stream 

velocity and chord length, t* =At r ,__(U /c). 

olLr o 

Typical results show that separation generally 
occurs when t* exceeds about 6. 

Dynamic pitching moment stall has been 
handled similarly. Test data showed, in general, 
that the dynamic moment break occurred before the 
lift break. This has been noted in Reference 6 
and attributed to the shedding of a vortex at the 
airfoil leading edge at the beginning of the 
separation process . Rearward movement of the vor- 
tex over the surface of the airfoil tends to main- 
tain lift, but drastically alters the pitching 
moment . 

To apply the Time Delay Model to a given air- 
foil, only static aerodynamic data are required. 
First, the airfoil static lift and pitching moment 
data are used to define the approximate variation 
in center of pressure between the static moment 
stall angle ol and an angle of attack a_ above 
which the center of pressure is assumed fixed. 
Secondly, an approximation is made to the c. 
versus a curve for fully separated flow. The se- 
quence of events occurring during one stall-unstall 
cycle is detailed in Figure 1. Briefly stated, 
lift and pitching moment are determined from po- 
tential flow theory until the nondimensional time 
t_ (which begins counting when the angle of attack 

exceeds the static moment stall angle) reaches t . 
At this point the pressure distribution begins to 
change, leading to rearward movement of the center 
of pressure and loss of potential flow pitching 
moment. Later, when x_ = t*, the lift breaks from 

the static line and decreases gradually with time 
to the fully separated value, c. (a). For 

•^EP 
t_>t* the center of pressure coincides with 

C.P.a—pCa) . At the point where a = 0, the rates 

at which c. approaches c lqii , p (a) and C.P. approaches 

C.P. SEp (a) (if it does not already equal C.P. g _ p {ot)) 

are increased. When a falls back below the quasi- 
static stall angle a-, , the center of pressure 
returns to the quarter chord, potential flow 
pitching moment effects are reintroduced and a 
second time parameter x_ is recorded to govern the 
rate at which c 1 returns to c. 

1 -T?0T 



14 




® o^. 



"1 "1F0T 
O.P. .= 0.25 

°m = C mPOT = 



ANGLE OF ATTACK, o 



it oe. 

IT dT 



at a-a_ 



moment stall time constant t 



2.0 



. C.P. *begins to move rearward with time toward c, ^ , cep(~) 
. g -_- is eliminated 



® 



T s <T g <T« 



c. remains equal to c r 



C.P. continues to shift aft with time, T„ 
C.P. - 0.25 * Cz ' \ ) [c.P. SEp (tx) - O.25] 



e ffl = c l( C.P. 



0.25) 



T > 



© Tg » lift stall time constant t* = 6.0 

. c. begins to decay toward ci_j, F (oi) 



1 1P0T 
. moves to 
variation in a 



, C.P. moves to C.P. g __(a) independent- of subsequent 



© d - 



the exponential rate at which c^ approaches ci SEp (o) is 

increased by a factor of 3 

if t <t„<t* the rate at which C.P. approaches C.P. gEp (ci) 

is increased by doubling the time increment 

T 2n+1 = T 2n + 24t n(^°n) 



© oKa. 



t, counting begins t, » I At n (_£cJ 

n«o 
at cc«ai,a<o 



C.P. returns to 0.25 

potential flow moment is reintroduced c 



c l * C1 P0T 
C.P. » 0.25 
c_ « c_ 



T 3A 



m 



"mPOT 



— ORIGINAL LOOP 

o RECONSTRUCTED LOOP 
TIME DELAY PREDICTION 



«M- "' 







f = 75.25 cps 




1.80 




N 
O 


1.40 


A? 


»- 




J* / 


z 




JS J 


UJ 




1.00 


S'T 


u. 




yyy 


UJ 





0.60 


^S 


1- 
u. 


0,20 


£~~^ 


_l 







12 



16 20 






ui 
2 

O 

2 

o 



{£ -0.I6P 




4 8 12 16 20 
ANGLE OF ATTACK , a , DEG 



Figure 2. Correlation with NACA 0012 Lift and 
Pitching Moment Hysteresis Loops. ■ 



Figure 1. Time Delay Unsteady Aerodynamic Model. 



Although additional correlation studies must 
be made to identify the effects of airfoil type on 
the time delay constants and although refinements 
to the present model may he implemented, this 
rather simple model represents well the essential 
features of the dynamic stall process. Correlation, 
typical of that claimed for other empirical methods 
(References 2 and 7) has been found with data from 
References k, 5» and 8. Only the o, A, B Method 
has produced better correlation (Reference k), but 
it suffers from the requirement for extensive 
testing and data processing. Figure 2 compares 
the NACA 0012 unsteady lift and pitching moment 
hysteresis loops measured in Reference k with Time 
Delay results. This correlation was achieved by 
setting the lift break time constant t* equal to 
k.Q instead of 6.0. Three-dimensional effects 
encountered in this test apparently reduced the 
time interval between static stall and dynamic 
lift stall. Also shown are the hysteresis loops 



15 



FAIRED CURVE OF REFERENCE 3 

A 5 = 6° DATA 
Q — T iME DELAY PREDICTION FOR 5«6° 





M0.1IZ3 








£~- -S3 












'AIREO CU 


'VE-v^ ^0- 






i" 4 


ft 


=&a^ 


^— < 


^A 


STABLE 






' tar' 




^H 












UNSTABLE 























.0 




ks 0.3375 


























< s& 


yy 


j 


k A 


<*> 


A 






r 1 

1 








^< 


V\ f 




1 !/ 
















i 






-.2 






UNSTABLE 






i 
i / 












<V 


Ji 


7T 






-.6 


























A 









Figure 3. 



MEAN INCIDENCE ANGLE, a„,DEG 



Correlation of Two-Dimensional Aero- 
Dynwni c Pitch Damping. 



predicted using the a, A, B Method. The a, A, B 

Method correlation is with the data from which the 

a, A, B coefficients were derived. 
* 

In addition to predicting the exact form of 
lift and moment hysteresis loops j an unsteady 
model should represent faithfully aerodynamic 
pitch: damping. Accordingly, the Time Delay 
Model was used to calculate two-dimensional aero- 
dynamic damping for the reduced frequency/mean 
angles of attack test points of Reference 9. 
Sample results plotted versus airfoil mean in- 
cidence angle of attack are shown in Figure 3. 
Generally excellent correlation of measured and 
predicted damping is noted. 

Other correlation of the Time Delay Method 
with two-dimensional oscillating airfoil test 
data has been good. During development of the 
theory, correlation was carried out with forced 
oscillating airfoil data for a range of airfoils, 
frequencies of f breed oscillation, Mach numbers, 
and angles of attack. Typical examples of the 
correlation obtained are shown in Figure h where 
measured and calculated hysteresis loops are 
shown for the V13006-7 airfoil. These test data 
taken from Reference 1 show the correlation with 
the Boeing Theory of Reference 1 as well. Corre- 
lation included hysteresis loops for different 
airfoils and covered a Mach number range from 0.2 
to 0.6. In all cases, the general character and 
magnitude of the hysteresis loops were well match- 
ed. In particular, the method provides the sharp 
drop in pitching moment that is often found when 
stall occurs . The oscillation frequency in 



Figure It is constant for the two cases, but Mach 
number and mean angle of attack are changed. The 
lift break occurs before the angle of attack reach- 
es its maximum value. In terms of the non-dimen- 
sional time parameter t*, the x* value of 6 at 
which lift stall occurs is reached before the maxi- 
mum angle of attack is reached. The Time Delay and 
Boeing Methods show comparable correlation for 
lift. For the Mach number, 0.!* case (Figure kb) the 
return to potential flow occurs earlier for de- 
creasing angle of attack than the return given by 
the Time Delay Method. Pitching moment correlation 
is better for the Time Delay Method. The triple 
loop characteristic is well duplicated. Similar 
correlation obtained with the Time Delay Method for 
a wide range of conditions demonstrated its promise 
as a practical method for analyzing unsteady aero- 
dynamics . 

Dynamic Stall Tests 

In order to obtain data useful in evaluating 
the two unsteady aerodynamic methods dynamic stall 
wind tunnel tests were run using a two-dimensional 
airfoil model. The model was oscillated at a 
frequency simulating the cyclic pitch variation on 
a helicopter rotor blade. Torsional frequencies 
representative of helicopter blade frequencies were 
obtained by varying a torsional stiffness element 
between the drive system and the airfoil section. 
The airfoil models were made to be as stiff as 
possible along their span and light in weight to 
approximate scaled helicopter blade mass and iner- 
tia properties . Hence the non-dimensional coeffi- 
cients in the equation of motion of the' model air- 
foil were close to those of the helicopter blade 
torsional equation of motion based on the aero- 
dynamics of the three-quarter radius on the re- 
tracting blade. Two different airfoils were fab- 
ricated, an HA.CA 0012 and an SC 1095. 

The model airfoils and drive system were 
designed to permit investigation of the effects on 
torsional response of torsional natural frequency, 
chordwise pitch axis location and torsional inertia 
over a range of IP frequencies for an NACA 0012 
airfoil and a cambered SC 1095 airfoil. The 
oscillating mechanism provided an 8-degree ampli- 
tude of motion of the model with an adjustable 
mean angle of attack. The model has a span of 
1.75 feet and a chord of 0.5 feet. The wind tunnel 
velocity was 275 fps for all tests . Time histories 
of the model angular motion were recorded at 
nominal driving frequencies of 8.0, 10.0, and 12.5 
eps . Tests were run for the full range of angle of 
attack for all the combinations of pitch axis , 
torsional inertia, torsional natural frequency, and 
airfoil type. A typical set of time histories 
for a basic reference condition (HACA 0012 airfoil, 
25 percent pivot axis, nominal blade inertia, and 
5P natural frequency ratio ) is shown for four mean 
angles of attack in Figure 5. These time histories 
represent the time average of ten cycles . 

The elastic torsional deflection of the 
airfoil (difference between total angular motion 
and input angular motion) was obtained for each 
test condition by subtracting the input angular 



16 



2.4 



2.0 



Z 

UJ 



LlI 

o 
o 

UJ 

o 

O 



< 

o 



TEST DATA 

BOEING METHOD 

TIME DELAY METHOD. 

STATIC DATA 




1.6 



1.4 



1.2 



1.0 



0.8 



0.6 



0.4 





















/^ 










// 




\~~" N 




6 


'// — ■ 


. — 


\ 


1 


^ 

/> 


K 


x" \ 


^ 
^ 


Y" 




/' 




i 









0.1 



E 
o 



UJ 

o 



W -0.1 



o 

2 



-0.2 — 



5 -0.3 



-0.4 



! 
















i 

i 
i 


> 




"~~\ 




i 




\\ 






i 











4 8 12 16 20 24 

ANGLE OF ATTACK , a , DEGREES 
A)f= 11.92 Hz, K = 0.l(5, M = 0.2 




8 12 16 20 24 



ANGLE OF ATTACK , a , DEGREES 
8) f= 12,07 Hz, K = 0.057, M = 0.4 



Figure k. Correlation of Bynamie Loops for the V130C-6-7 Airfoil in Forced Pitch Oscillation. 



17 







































































































































s 


s s 




























8 










































O 




















4 \ 




















/ 


1- 














a 


f r 


























S>s a 
























































































































































v 










































\ 


s 



































































































Figure 5. 



NON-DIMENSIONAL TIME, Slt/Zt 



Averaged Time Histories of Angle of 
Attack for the Model Airfoil. 



position time history from the averaged airfoil 
angular position time history: 



S(t) ■ o(t) - (a + BBin2irr). 

where 0(t) is the difference between the measured 
non-dimensional angular time history response «(t) 
and the input driving motion. The non-dimensional 
time t is given by t/T, where the period, T, was 
established from the ten-cycle time-averaging 
process for that run. Some statistical variation 
in measured response was noted when stall flutter 
occurred, but in general the ten cycle time 
averaged response was representative of the 

I'Iref 

I,l.5xI REF 



(9 

W 

o 



1 1 

Li 
_l 
(9 

Z 
< 

z 



S 3 

UJ 







MAC A 0012 
5? PA =0.25 


















S« = ! 












\"7 


>* 






X 


tug 


./ 




ft 


V 









2 - 



< 







MACA 001 
*PA=0.22 


I 






















. 


i^^ 


■as 






/' 


f, 






-Og = 


1<" 


/ / 


f\ * 


.'" 


— •> 


'" 


' 4 


Ada =7 






3 

2h 







SC 1095 | 
Xp A = 0.25 
































F S« = 7 


> 






// 


s' 


s 

.' 








L' 


"~ 







6 







SC (095 | 
5! P a=0.22 
























,«■ 


^mm 




(SqsS 




//, 


s 








_J. 


A 


^; 


. W 






4 









10 14 18 ~6 10 14 

MEAN INCIDENCE AN~GLE,a. M , PEG. 
Figure 6. Model Airfoil Elastic Deflection. 



18 



measured data. Two measures of stall response 
amplitude were extracted from each of the Q(t) time 
histories . These were A0^ which is one-half of 
the initial response to stall and 9 JgpTp which is 
one-half of the overall peak-to-peak elastic de- 
flection. 

It was found that the initial stall response 
parameter A0, gave the most consistent indication 
of susceptibility to stall flutter. The possible 
reduction in flutter amplitude introduced by the 
time averaging procedure when there was cycle-to- 
cycle variation in phase made it somewhat difficult 
to assess the amplitude of flutter response. For- 
tunately, the initial stall deflection showed 
virtually no cycle-to-cycle variation. Figure 6 
compares measured initial deflection angles for an 
excitation frequency B of 10 cps for the two air- 
foils at all combinations of airfoil natural 
frequency ratio (Wq = us Q /&l torsional inertia, and 
pitch axis. Certain general trends of deflection 
angle can be identified in the test results. 

1. Elastic deflection increases with mean 
incidence angle. 

2. For the same torsional inertia, response 
is generally greater for the lower 
frequency airfoil section. 

3.. The amplitude of response is inversely 
related to torsional inertia. 

k. Forward movement of the pitch axis leads 
to a decrease in deflection. 

5. SC 1095 airfoil dynamic stall response 
begins to build up at a higher mean 
incidence angle than the 0012, but the 
two airfoils have comparable responses 
once stall is penetrated. 

Correlation Study of Two -Dimensional Results 

The two-dimensional flexured airfoil test 
data were compared with predictions based on 
various unsteady aerodynamic methods. The single 
torsional degree of freedom differential equation 
of motion for the flexibly mounted airfoil section 
oscillating in the wind tunnel test section is 
given by 

I-£.+ coi + KU-Ojh) = M(t) +Ko sin fit 

where c ■ equivalent mechanical damping per unit 
span 
I = airfoil torsional inertia per unit span 
K = torsional spring constant 
Mfc= applied aerodynamic moment 
t = '.time' 

-a~=-aix£oil angle of attack 

tt «*.. amplitude of angular oscillation 

om= mean angle of the oscillation 

Q » angular frequency of the applied" torque • 

This equation was solved numerically using 
the unsteady aerodynamic models to calculate the '- 
applied aerodynamic moment M(t). For the a, A, B 
Method unsteady data tables obtained from earlier 
oscillating airfoil tests, Reference l*,were scaled 
for both airfoils . The measured steady state lift 



18 



and pitching moment data served as inputs in the 
Time Delay calculations. Additionally, the air- 
foil mean incidence angle used in the Time Delay 
solution was two degrees less than that set in 
the wind tunnel. The open jet flow deflection 
experienced at high unsteady lift coefficients was 
sufficient to decrease actual peak angles of 
attack to a value somewhat lower than the geo- 
metrically impressed; pitch angle. The two-degree 
correction to ay, gave consistently better corre- 
lation of the initial stall time. 

Correlation between measured wind tunnel 
model response and response calculated with the 
unsteady models was examined for thirty-six test 
conditions. The set of cases studied was suffi- 
cient to evaluate the independent effects on air- 
foil stall response of mean incidence angle, 

TEST 

a,A,B 

TIME DELAY 



g. ° 










NACA 0012 AIRFOIL 
a M .14 e w s /ft"? 


! 








■; 




CHANGE IN BATUMI. 
FEEQUSHCY 








I=IrEF Xpa-0.25 






V 


9j 










l I 
























\\ 




4 


\ 




J- 


X. 


\S, 






f— 


^? J 




XZ 


^ 






""" 






— "S 


■4 / 


i 




^ 








































































































___ 




i I i 

SC 1095 AIRFOIL 


■ 




r 


. 




■ i 


'N 














CHAHGE 


II 


*■*» *»-«h 






i 


\ 






























i 


\ 








\ — 




/ 


^ 












— ^ 




> N 




i 






> N 






-*C; 




-7 




V 




/ 










■•> 


\V 




$ 






f- 






^ 


V. 




J> 


L> 















V 


l i 






4 






























\ 


J 





































































torsional natural frequency, chordwise pitch axis, 
torsional inertia, and airfoil type. Relative to 
a baseline case 1*aken to be the NACA 0012 airfoil 
at ^ = W 5 , ae/fl = 5, ^, A = 0.25, and I = I-^,, 
Figure 7 shows measured and predicted effects of 
mean angle, torsional natural frequency, pitch axis, 
and airfoil type on time histories of elastic de- 
flection. Comparison of the measured and predicted 
effects of airfoil mean angle of attack indicates 
that deeper penetration into stall results in 
sharper initial stall deflection and larger resid- 
ual stall flutter response. The two analyses pre- 
dict these effects qualitatively, but each - 
especially the Time Delay model - overprediets the 
amplitude of response. The main effects of an in- 
crease in torsional natural frequency are a shift 
in response frequency and a decrease in the am- 
plitude of elastic defleetion. Figure 7 shows 
good correlation of response amplitude, although 
both analyses predict an initial stall response 
earlier than that measured. Moving the airfoil 
pitch axis forward causes delay in initial stall 
time and reduction in amplitude of response. The 
analytical results do predict the reduction in 
response amplitude, but the Time Delay model still 
results in overpredicted response. Finally, a 
comparison between the NACA 0012 and the SC 1095 
airfoils shows a delay in the initial stall time 
for the SC 1095 airfoil, which had a static stall 
angle measured in this wind tunnel to be about 
three degrees higher than that of the NACA 0012. 
However , the SC 1095 stall flutter amplitude was 
comparable to that experienced by the NACA 0012 
at this condition. 

The time history correlation was good in 
that the initial response and the frequency of the 
subsequent oscillations were predicted. The trends 
observed in test were well matched by the analysis , 
although. the Time Delay model generally overpre- 
dicted stall flutter response. The basic effects 
of structural changes on blade response time 
histories were well predicted by either analysis . 

Although torsional elastic deflection is 
important in determining rotor stability and per- 
formance, the torsional moments resulting from 
stall flutter are the designer's primary concern. 
To measure the trends of torsional moment with 
parameter changes , the twisting moment experienced 
by the flexible connector in the model airfoil 
drive system was calculated for each test condition. 
The torsion moment Mq was calculated using the 
equivalent spring stiffness of the connector: 

(I 



M„ 



K 
eq 



airfoil 9 



2 
<a„ ) 



NON0IMENSIONAL TIME,Q</2r 

Figure 7, Effect of ParanBters~an rftirfoil Response. 



The torsional moments corresponding to the initial 
stall deflection angle A0 1 were used to show the 
effects of blade parameters on structural moments. 
Figure 8 presents typical results for three com- 
binations of airfoil type and mean angle of attack. 
It was generally found that decreasing torsional 
natural frequency reduced stall flutter moments. 
Although the stiffer system experienced lower 
response amplitudes, the corresponding structural 
moments were increased: 



19 



TEST 



40 
20 
L 
[ 60 
40 
20 

80 - 
60 - 
40- 
20 - 



TIME 
DELAY 



NACA 0012 AIRFOIL, ct M = I 



a,A,B 











N 


ACA C 


012 


AIRFO 


L, 


*M= ,4 ° 


- 























SC 1095 AIRFOIL, a M = 14° 



« H 



m O ix 



13 |x H 



Figure 8. 



Effect of Airfoil Parameters on Model 
Airfoil Vibratory Torsional Moments. 



05p u 05p 5p P 5p 



Forward placement of the airfoil pitch axis gener- 
ally decreased vibratory torsional moments. The 
two analyses predicted this trend with comparable 
accuracy. That forward movement .of the airfoil 
pitch axis relative to the aerodynamic center 
reduces stall flutter moments can he understood 
based on lift and pitching moment hysteresis 
loops . For an airfoil with pitch axis forward 
of the center of pressure, positive lift forces 
cause negative moments about the pitch axis. 
For positive lift, the lift hysteresis loop is 
usually traversed in the clockwise direction, 
which contributes a negative pitching moment 
loop in the counterclockwise (stabilizing) 
direction. A decrease in torsional moment am- 
plitude with decreasing torsional inertia was 
generally found throughout the testing. This 
trend, evident in two of the conditions shown in 
Figure 8, is predicted somewhat more correctly 
by the Time Delay Analysis. Finally the two 
airfoils are compared in Figure 9- For two 
different combinations of inertia and pitch axis, 
high stall flutter moments are delayed in mean 
angle with the SC 1095 airfoil. 



60 



40 



20 



I=I REF ,XpA=0.25,Og = 5 







TEST 

























,S H 




NACA 0012, 


i 

-SC 1 










/ / 

' 1 










1 









I«l.5xI REF ,5? 


»«0.22. 


"8=5 


- 




1 
TEST 














Ss'—~ 


-NACA 










/ 










' 'SC 1095 








T~ 









s . 



z 

Ul 




TIME 


3ELAY 




o 

S 40 

-J 
< 




..._. 










3 

§20 

1- 

to 








/ 








/ 








-1 

P n 






/ 









- 


TIME DELAY 









~/ 


/ 


• 










1 


/ 








1 
/ 















! 

a,/ 


*,B 






40 


■- 


---•• 




-- 






20 




/ 


s 


^ 






6 10 14 18 6 10 14 

MEAN INCIDENCE ANGLE, a M , DEGREES 

Figure 9. Effect of Airfoil Type on Structural 
Moments 

Flight Test Correlation 

Flight test data were correlated with the 
Normal Modes Blade Aeroelastic Analysis for both 
the CH-53A and CH-5ta aircraft. Both models of 
unsteady aerodynamics were used. Information on 
the blade analysis used can be found in Refer- 
ence 10. 

Correlation of CH-53 control system loads , 
blade stresses and required power was studied at 
a nominal aircraft gross weight of fe?,000 lb 
(Cm/a = 0.083), a tip speed of 710 ft/sec, and a 
3000 ft density altitude for airspeeds ranging from 
100 knots to 170 knots . Inclusion of variable in^ 
flow was found to be essential in calculating the 
proper levels of blade bending moments . It also 
provided some improvement in the correlation of 
blade torsional moments . 

The a, A, B and Time Delay aerodynamic models 
are compared at 137 knots in Figure 10. Figures 
10a and 10b shows that the computed blade stresses 
are comparable for the two methods. However, the 
push rod loads calculated with the Time Delay 
Model are much less than values calculated with the 
o , A, B Method and measured values . The Time 



20 



'(D) 



Cc) 



8000 



Q-_ 4000 
in 



-4000 



-8000 



4000 



r/R 


= 08 














• 




























n as 






















- TIME DELAt 


M 








































if 


\ 


















^^ — 


















11/ 
f 
















^ 


^ 


7j 


\>^ 






















z.^ 






J(V 


































V 




/ 


f\ 


/ 
















%tf\ 
















\ 


\ 


y 


















































60 



120 180 240 

BLADE AZIMUTH ANGLE , DEGREES 



300 



360 



60 



Figure 10. Correlation .of CH-53A Blade Stresses and Pushrod Loads. 



Delay* results generally do not give sufficiently 
large oscillations in stall. 

That better correlation of stall flutter 
moments was possible with the a, A, B Method is 
evident in Figure 11a which shows the buildup 
of vibratory pushrod load amplitude with airspeed. 
The a, A, B model predicts a buildup rate almost 
identical with the mean of the' test data. A 
discrepancy of no more than 10 knots in the knee 
of the control load curve is evident at this thrust 
coefficient. Figure lib shows the correlation of 
pushrod load amplitude achieved with the a, A, B 
Method at three thrust coefficients . 

Calculated C&-5^ control loads were also 
generally less than measured values . Figure 12 
indicates that a definite stall boundary is pre- 



dicted by the analysis . Relative to the CH-53A 
calculations , a decreased control load stall 
speed and an increased rate of buildup with air- 
speed are clearly predicted. Again, higher loads 
are computed based on the a, A, B Model. The 
comparison of measured and predicted push rod 
load time histories indicates that the a, A, B 
results reflect a buildup in the higher fre- 
quency loads much more accurately than do the 
Time Delay calculations. 

It is not entirely clear why, relative to 
the o, A, B method, the Time Delay model under- 
prediefcs helicopter control loads while over- 
predicting the stall flutter oscillations of the 
two-dimensional wind tunnel model. Examination 
of several blade section pitching moment /angle 
of attack hysteresis loops indicates not so much 



21 



(a) 
3200 



m 
«2400 

Q 

3 ' 



Q 

£1600 



O 

S 800 



... 

A 


1 r ■ 

— TEST DATA 

— a,A,B 


1 














— Tl 


ME Dt 


.LAY 
















































]// 


/ 
















/fi 


/ 
















Sv 


y 


s"- 












* 


<La 


s* 

































40 



80 120 

AIRSPEED, KNOTS 



160 



200 



(b) 



4000 



.3200 



2400 



1600 



< 

CD 

> 



800 



1 1 1 1 1 

C T /cr TEST ANALYSIS 










0.083 


& 










































PUSH 


ROD 


JOAD= 


±220 


LB 
















\ 


_ 




/ 1 


n/i 


/m, 


/ 

" 










1 f 


A/ 


r ■( 














i i 

1/ 


/ 


/ 














// 


', 


/ i 


3 












L^ 


f 


/ 


/ 
















Y^" 











































40 



80 



120 160 ' 

AIRSPEED, KNOTS 



200 



240 



Figure 11. 



2400 



Correlation of CH-53A Vibratory 
Pushrod Loads . 



1600 



800 



§ 




Figure 12. 



40 80 

AIRSPEED, KNOTS 

Correlation of CH-^B Vibratory 
Pushrod Loads. 



160 



that more negative pitch damping is present in the 
a, A, B results. Rather pitching moments along the 
blade are more in phase with each other, leading to 
larger modal excitation. In the o, A, B formu- 
lation, pitching moment coefficients are tabulated 
as functions of a, a and 8 values all along the 
blade. This formulation leads to similarly phased 
pitching moments. In the Time Delay Model, moments 
are calculated based on the angles of attack ex- 
ceeding the steady state stall angle for a certain 
interval of time and are not solely dependent on 
the instantaneous angle of attack characteristics . 
For small differences in calculated angles of 
attack, computed pitching moments for adjacent 
blade sections can be different in phase. Because 
the two-dimensional wind tunnel airfoil was modeled 
as a single panel for the calculation of aerodynamic 
forces, the effects of simultaneous spanwise stall- 
ing were not a factor in the correlation with that 
data. 

Comparison of Methods 

Because the a, A, B Method has demonstrated 
better correlation with flight test data, it con- 
tinues to be the method in use for blade design 
analysis. However , development of both methods 
continues. The ot, A, B Method provides a relative- 
ly direct and simple procedure for calculating un- 
steady aerodynamic loads . Correlation has been 
good with test data but its disadvantage centers 
largely on the apparent need for extensive tests 
to provide the body of tabulated data required for 
each airfoil. Some success has been obtained by 
scaling the KACA 0012 unsteady aerodynamic tables 
based on steady state differences between airfoils . 
Work is also being done on developing analytical 
expressions to replace the tabulated data. These 
may lead to the ability to synthesize the data 
required for a given airfoil , which would make the 
method more desirable for general applications . 

The Time Delay Method has the great advantage 
of requiring only .steady state airfoil data for its 
application. The correlation with forced oscil- 
lations of two-dimensional airfoils demonstrated 
its applicability over a wide range of conditions . 
Correlation with the tests described in this paper 
showed no clear advantage of the Time Delay Method 
over the a, A, B Method, and correlation with 
flight test data was definitely poorer with the 
Time Delay Method. Further work must be done to 
investigate the reasons for the poor flight test 
correlation. The problem may result from the 
assumption in the analysis that, on a blade, each 
radial section acts independently of its neighbor- 
ing sections . This causes a more random stalling 
along the span with time, which smoothes out the 
changes in blade loading. The propagation of stall 
along the span for the three-dimensional case of a 
helicopter blade must be added to the Time Delay 
Method. The a, A, B Method does provide spanwise 
correlation in loading by use of torsional mode 
acceleration to calculate the B parameter. This 
acceleration is in phase for each point along the 
blade span. Incorporation of a suitable radial 
propagation model in the Time Delay Method may make 
this a more versatile, more easily applicable, and 



22 



more accurate model of unsteady aerodynamics. 
Until this can be shown the a. A, B Method 
continues in use in blade design. 

Conclusions 

1. The a, A, B and Time Delay unsteady aerodynamic 
models predict with good accuracy the lift and 
pitching moment hysteresis loops and the aero- 
dynamic pitch damping of rigidly driven os- 
cillating airfoils. 

2. Two-dimensional stall flutter tests indicate 
that reducing blade torsional stiffness , re- 
ducing blade torsional inertia and moving 
blade pitch axis forward decrease stall flutter 
induced moments. Inception of stall flutter 
was delayed with the SC 1095 airfoil relative 
to the NACA 0012 airfoil; however, once 
initiated, stall flutter loads for the two 
airfoils were generally comparable. 

3. Stall flutter response of the two-dimensional 
model airfoils and the effects of airfoil 
structural design parameters on blade torsion- 
al moments can be calculated using both un- 
steady models. The Time Delay method gives a 
high prediction of response amplitude. 

k. Good correlation of CH-53A and CH-5^B blade 
stresses and control loads was obtained with 
a rotor aeroelastic analysis employing vari- 
able rotor inflow and unsteady aerodynamics. 
Best correlation was achieved using the a, 
A, B unsteady model. The Time Delay method 
generally underpredicted full scale rotor 
stall flutter response. 

5. The a, A, B model is in use for blade design 
analysis. Refinements to the Time Delay Method 
may make it a more versatile and more easily 
applied unsteady aerodynamic model. 



References 

1. Gormont, R. E., A MATHEMATICAL MODEL OF UN- 
STEADY AERODYNAMICS AND RADIAL SLOW KIR 
APPLICATION TO HELICOPTER ROTORS, USAAMRDL 
TR 72-67. U. S. Army Air Mobility Research 
and Development Laboratory, Fort Eustis, 
Virginia, May 1973. 

2. Ericsson, L. E. and Reding, J. P., UNSTEADY 
AIRFOIL STALL REVIEW AND EXTENSION, AIAA 
Journal of Aircraft, Vol. 8, No. 8, 
August 1971- 

3. McCroskey, W. J., RECENT DEVELOPMENTS IN 
ROTOR BLADE STALL, AGABD Conference Pre- 
print No. Ill on Aerodynamics of Rotary 
Wings, September 1972. 



h. Carta, P. 0., Commerford, G. L., 

Carlson, R. G. , and Blackwell, R. H., 
INVESTIGATION OF AIRFOIL DYNAMIC STALL AND 
ITS INFLUENCE ON HELICOPTER CONTROL LOADS, 
United Aircraft Research Laboratories; 
USAAMRDL TR 72-51, U. S. Army Air Mobility 
Research and Development Laboratory, 
Fort Eustis, Virginia, September 1972. 

5. Gray, L. and Liiva, J., WIND TUNNEL TESTS 
OF THIN AIRFOILS OSCILLATING NEAR STALL, 
The Boeing Company, Vertol Division; 
USAAMRDL TR 68-89A and 68-89B, U. S. Army 
Aviation Materiel Laboratories, Fort 
Eustis, Virginia, January 1969. 

6. Ham, N. D. and Garelick, M. S. , DYNAMIC 
STALL CONSIDERATIONS IN HELICOPTER ROTORS, 
Journal of the American Helicopter Society, 
Vol. 13, No. 2, April 1968. 

7. Tarzanin, F. J., PREDICTION OF CONTROL 
LOADS DUE TO BLADE STALL, American Heli- 
copter Society, 27th Annual National 
Forum, May 1971. 

8. Arcidiacono, P. J., Carta, F. 0., 
Caselini, L. M., and Elman, H. L., 
INVESTIGATION OF HELICOPTER CONTROL LOADS 
INDUCED BY STALL FLUTTER, United Aircraft 
Corporation, Sikorsky Aircraft Division; 
USAAVLABS TR 70-2, U. S. Army Aviation 
Material Laboratories, Fort Eustis 
Virginia, March 1970. 

9. Carta, F. 0., and Niebanck, C. F., 
PREDICTION OF ROTOR INSTABILITY AT HIGH 
FORWARD SPEEDS, Volume III, STALL FLUTTER, ■ 
USAAVLABS TR 68-18C, U. S. Army Aviation 
Materiel Laboratories, Fort Eustis, 
Virginia, February 1969. 

10. Arcidiacono, P.' J., PREDICTION OF ROTOR 
INSTABILITY AT HIGH FORWARD SPEEDS, Vol. I, 
Steady Flight Differential Equations of 
Motion for a Flexible Helicopter Blade with 
Chordwise Mass Unbalance, United Aircraft 
Corporation, Sikorsky Aircraft Division; 
USAAVLABS TR 68-18A, U. S. Army Aviation 
Materiel Laboratories, Fort Eustis, Virginia, 
February 1969. 



23 



COMPUTER EXPERIMENTS .ON PERIODIC SYSTEMS 

IDENTIFICATION USING ROTOR BLADE TRANSIENT 

FLAPPING-TORSION RESPONSES AT HIGH ADVANCE RATIO 

K. H. Hohenemser and D. A. Prelewicz 
Washington University, St. Louis, Missouri 63130 



Abstract 

Systems identification methods have 
recently been applied to rotorcraft to 
estimate stability derivatives from 
transient flight control response data. 
While these applications assumed a 
linear constant coefficient representa- 
tion of the rotorcraft, the computer 
experiments described in this paper used 
transient responses in flap-bending and 
torsion of a rotor blade at high advance 
ratio which is a rapidly time varying 
periodic system. It was found that a 
simple system identification method ap- 
plying a linear sequential estimator 
also called equation of motion estimator, 
is suitable for this periodic system and 
can be used directly if only the accel- 
eration data are noise polluted. In the 
case of noise being present also in the 
state variable data the direct applica- 
tion of the estimator gave poor results, 
however after pref iltering the data with 
a digital Graham filter having a cut-off 
frequency above the natural blade torsion 
frequency, the linear sequential estima- 
tor successfully recovered the parameters 
of the periodic coefficient analytical 
model. 

Notation 

B Blade tip loss factor 

F = (Ix/lSIfXc/R) 2 First blade tor- 
sional inertia 
number 

State matrix 

Process noise modulating 
matrix 

Fourier transform of 
weighting function 

Measurement matrix 

State matrix = measurement 
matrix 



F(x,t) 
G(t) 

H(u) 

H(x,t) 
H(5,a) 



Presented at the AHS/NASA-Ames Special- 
ists' Meeting on Rotorcraft Dynamics, 
February 13-15, 197t. This work was 
sponsored by AMRDL, Ames Directorate, 
under Contract No. NAS2-4151. 

"Now at the Westinghouse Bettis Atomic 
Power Lab. Westmifflin, Pennsylvania. 



*1 

If 

J 

P(t) or P 



Q= (I 1 /itI f )c/R 

R 

R 
a,b,c 

a 

c 

f 

t 
v 
w(jAt) 

w 

X 

z 



Blade flapping moment of 
inertia. 

Blade feathering moment 
of inertia. 

Quadratic cost function. 

Covariance matrix of 
conditional state vector 
probability distribution 
given measurements. 

Blade flapping natural 
frequency . 

Second blade torsional 
inertia number. 

Measurement noise co- 
variance matrix. 

Blade radius 

Unknown parameters to be 
estimated in flapping- 
torsion problem. 

Parameter vector. 

Blade chord. 

Blade torsional natural 
frequency. 

Non-dimensional time. 

Measurement noise vector 

Smoothing weights. 

Process noise vector 

State vector 

Measurement vector. 

Flapping angle. 

Blade Lock number. 

Blade torsion deflection 

Acceleration vector. 

Rate of displacement 
vector . 

Blade pitch angle. 

Rotor inflow ratio, con- 
stant over disk. 

Rotor advance ratio. 

Displacement vector. 



T In order to retain the conventional sym- 
bols in helicopter aerodynamics (Ref- 
erence 7) and in systems analysis (Ref- 
erence 9 ) some symbols are used in two 
different meanings. 



25 



Subscripts 

o 

c,t 



Superscripts 



Notation 
(conf) 

Standard deviation. 
Circular frequency. 



Initial or mean value. 

Beginning and end of fil- 
ter cut-off frequencies. 



Time differentiation. 

Smoothed data after fil- 
tering. 

Estimate 

Matrix transpose 



The question often arises, how to 
best select some parameters of a given 
analytical model of a dynamic system on 
the basis of transient responses to 
certain inputs either obtained analyti- 
cally with a more complete math model or 
obtained experimentally. In rotor craft 
flight dynamics one may want to use a 
linear constant coefficient math model 
and select the state matrix in an opti- 
mal way from the measured data ob- 
tained in a number of transient flight 
maneuvers. One also may have a more 
sophisticated non-linear analytical 
model of the rotorcraft. The problem 
then is how can the simpler linear math 
model be selected to best represent the 
responses of the more complete analyti- 
cal model; or one may have the dynamic 
equations of a rotorcraft without the 
effects of dynamic inflow and one de- 
sires to modify some of the parameters 
in such a way that dynamic inflow ef- 
fects are best approximated. It is 
known from theoretical studies, for 
example Reference 1, that a reduction in 
blade Lock number can approximately 
account for rotor inflow effects in 
steady conditions. The question then is 
whether changes in parameters can also 
account for inflow effects during 
transient conditions. 

The idea of using transient re- 
sponse data to determine parameters of 
an analytical model is certainly not 
new. Recently, however, considerable 
interest in this area has been de- 
veloped and a number of approaches have 
been studied which are unified under the 
title of "system identification". There 
is a considerable and rapidly growing 
literature in this field. System iden- 
tification methods generally fall into 



two classes: (1) deterministic methods - 
usually some variation of the classical 
least squares technique and (2) proba- 
bilistic methods which determine the 
parameters as maximum likelihood esti- 
mates of random variables. Some methods 
can also be interpreted either on a 
deterministic or on a probabilistic 
basis. References 2 and 3 are typical of 
recent work using deterministic methods. 
Both of these studies illustrate the 
feasibility of determining coefficients 
in time invariant linear systems from 
transient response data. Reference 4 de- 
scribes many of the probabilistic tech- 
niques. Reference 5 gives a detailed 
discussion of the various methods in 
their application to V/STOL aircraft and 
Reference 6 presents an identification 
method suitable for obtaining stability 
derivatives for a helicopter from flight 
test data in transient maneuvers. The 
studies of References 5 and 6 assume a 
linear constant coefficient representation 
of the system. A rotorcraft blade is, 
however, a dynamic system with rapidly 
changing periodic coefficients. It ap- 
peared, therefore, desirable to try out 
methods of system identification for a 
periodic dynamic system. 

Selection of Identification Method 

If one assumes that only the state 
variables have been measured but not the 
accelerations, one must use a non-linear 
estimator since the estimate of a system 
parameter and the estimate of a state 
variable appear as a product of two un- 
knowns. A non-linear sequential esti- 
mator was tried on the simplest linear 
periodic system described by the Mathieu 
Equation. It was found that the non- 
linear estimating process diverged in 
most cases, unless the initial estimate 
and its standard deviation were selected 
within rather narrow limits. Reference 6 
uses a sequential non-linear estimator 
but initializes the process by first ap- 
plying a least square estimator, which 
needs in addition to the state variable 
measurements also measurements of the ac- 
celerations. In the case of the problem 
of Reference 6 the least square estimator 
yielded a rather good set of derivatives 
and the improvement from the much more 
involved non-linear estimation was not 
very pronounced. From this experience 
it would appear that one needs to apply 
the least square or an equivalent linear 
estimator any way and that in some cases 
it is doubtful whether or not the sub- 
sequent application of a non-linear es- 
timator is worth the considerable effort. 

After conducting the rather unsatis- 
factory computer experiments to identify 
a simple periodic system with the 



26 



non-linear estimator, all subsequent work 
was done with a linear sequential esti- 
mator. This estimator is equivalent to 
least square estimation but has the ad- 
vantage of being usable for "on-line" 
system identification. The inversion of 
large matrices is avoided and replaced by 
numerical integration of a number of 
ordinary differential equations. The 
computer experiments were conducted with 
the system equations of Reference 7 for 
the flapping - torsion dynamics of a 
rotorblade operating at advance ratio 1.6. 
Reference 7 assumes a straight blade 
elastically hinged at the rotor center 
and stipulates linear elastic blade twist. 
The system used here for the computer ex- 
periments represents only a relatively 
crude approximation, since at 1.6 advance 
ratio blade bending flexibility is of im- 
portance, see for example Reference 8. 
The coefficients in the system equations 
are non-analytic periodic functions which 
include the effects of reversed flow. 

The identification algorithm used in 
this report is easily derived using the 
extended Kalman filter discussed in the 
next section. Although the algorithm 
does not provide for noise in the state 
variables, one can nevertheless use it 
also for noisy data if one interprets the 
estimate, which normally is a determinis- 
tic variable, as a sample of a random 
variable. The effects on system identi- 
fication of computer generated noise in 
both the acceleration data and in the 
state variable data were studied. However, 
no errors in modeling were introduced 
since their effects can only be evaluated 
on a case by case basis. 

Extended Kalman Filter 

The extended Kalman filter is an 
algorithm for obtaining an estimate x of a 
state vector x satisfying 



x = F(x,t) + G(t)w 



Process Equation(l) 



given noisy measurements z related to x 
via 



3F 
3x 



P + P 



(s) 1 * <«* - -(s)v^ 



3H 
3x 



x(o) 



Covariance Equation (H) 

P(o) = P 
Initial Conditions (5) 



x and P can be interpreted as vector mean 
and covariance matrix of a conditional 
probability distribution of the state 
vector x, given the measurement vector z. 

However, since the extended Kalman 
filter is a biased estimator (see Refer- 
ence 5) and since the correct value of 
P is not known, P cannot be used as a 
measure of the quality of the estimate. 
Rather, the rate of decrease of P is an 
indication of the amount of information 
being obtained from the data. When P 
approaches a constant value then no 
further information is being obtained. 

The extended Kalman filter may also 
be interpreted as an algorithm for ob- 
taining a least square estimate recur- 
sively. The estimate is such as to 
minimize the following quadratic oost 
function 



J = 1/2 { (x - V^ 1 (x - x ) 



J o wV" 1 * + |z-H(x,t)) T 

,t)W 



z - H(x, 



Cost Function (6) 

where now P , R and Q are arbitrary 
weighting matrices, which may be se- 
lected for good convergence of the 
algorithm. Since 1.) numerical methods 
for solving ordinary differential equa- 
tions are well developed and 2.) R is 
usually a diagonal matrix so that R -1 is 
easy to obtain, this algorithm is compu- 
tationally very efficient. 



H(x,t) + v 



Measurement Equation (2) Estimation of Unknown Parameters 



In the above equations w represents zero 
mean white Gaussian process noise with 
covariance matrix Q, v represents zero 
mean white Gaussian measurement noise 
with covariance matrix R. An optimum es- 
timate x of x can be obtained by solving 
the extended Kalman filter equations 
(see Reference 9) 



x = F(x,t) + Pj 



mVr 



i,t)J 



I - H(x, 
Filter Equation (3) 



If we wish to estimate the vector a 
of unknown parameters we substitute a for 
x in the Kalman filter Eq. 3. For con- 
stant parameters we have 

a = o Process Equation (7) 

so that F(x,t) = w = o. The system equa- 
tion is then used as the measurement 
equation 



H(c,a) + v 



Measurement Equation 
System Equation (8) 



27 



C is the vector of measured accelerations, 
5 is the measured state vector and v can 
be interpreted as acceleration measure- 
ment noise or as system noise (including 
modeling errors). The Kalman filter 
equations are then 



R _1 U - H(5,a)] 



/3H\ T -1 



Filter Equation (9) 



P = -Pf-rr 



3H 
3a 

Covariance Equation (10) 



For P ■*■ o the measurements lose influence 
on the estimate and one obtains 

a = o Asymptotic Filter Equation (11) 

which agrees with the process equation. 
Again P and R may be selected for good 
convergence. A convenient choice for the 
initial estimate is a(o) = o. The ele- 
ments of R should be large enough to pre- 
vent the elements of P from becoming 
negative due to computation errors in the 
numerical integration. 

Note that £, the state vector, is 
also a measured quantity. If measurement 
errors are present then this estimation 
algorithm is biased by an amount approxi- 
mately proportional to the noise to sig- 
nal ratio in the state variable measure- 
ments, see Reference 5. It is therefore 
advantageous to reduce the noise ratio 
before using the estimator. Methods for 
doing this are discussed in a later sec- 
tion on filtering of the response data. 

In practice, one can almost always 
choose the parameters to be identified in 
such a way that H(£,a) is a linear func- 
tion of a. The estimator (9), (10) is 
then linear and problems of nonuniqueness 
and filter divergence are easily avoided. 
For this case, we call the algorithm the 
linear sequential estimator. 

The extended Kalman filter assumes 
that the noise processes w and v are 
white and Gaussian. This will never be 
the case in practice especially if w must 
account for the effects of modeling er- 
rors. Because the extended Kalman fil- 
ter may also be interpreted as yielding a 
least squares estimate for a given sample 
of the state £ and acceleration c, we can 
regard the resulting estimate as a sample 
from a random variable. Determination of 
this random variable would necessitate a 
complete simulation, i.e., mean and 
variance determined by averaging over 
many runs. Since this approach is 
expensive of computing time, efforts here 



have been directed toward recovering para- 
meters from a single run of computer gen- 
erated response data. 

The above approach to parameter es- 
timation allows the use of high order of 
accuracy numerical integration (i.e., pre- 
dictor corrector) schemes to solve the 
system of ordinary differential equations 
provided that the response data are suf- 
ficiently smooth. The parameter estima- 
tion is rapid and requires little com- 
puter time. R and P can be freely se- 
lected to obtain good convergence. The 
reason for this benign behavior of the 
estimation method is the linearity of the 
filter equations in the unknowns. If the 
accelerations of the system are not 
measured, one must estimate state vari- 
ables and parameters simultaneously from 
a nonlinear filter equation. This non- 
linear estimation requires an order of 
magnitude more computer effort and it is 
very sensitive to the initializations and 
to the correct assumptions of process 
noise and measurement noise. As 
mentioned before, we began by applying 
the nonlinear estimator to the identi- 
fication of parameters in Mathieu's 
equation for a periodic system. The re- 
sults were unsatisfactory since filter 
divergence occured for many choices of 
P and R. However, for the linear 
sequential estimator divergence could be 
avoided by following simple rules in 
selecting x(o), P Q and R. 

Identif iability of System Parameters 

It is obvious ^from the filter 
equation (9) that a will asymptotically 
approach a constant value provided that 
P -*• o. The covariance equation (10) can 
be solved explicitly (see Appendix A) to 
yield 



fl& 



W 



3H 
3a 



dt 



(12) 



If the integral is replaced by a sum, this 
is the error equation for the standard 
least square method. If P Q / o, then 
P(t) -»• o whenever the integrand in the 
above equation is positive definite for 
all t. This is then a sufficient condi- 
tion for identif iability. Note that 
3H 

3a is a function of the system response 
and hence also of the excitation, so that 
the identifiability depends not only upon 
the system but also upon the type of ex- 
citation. From the measurement equation 
(8) we see that the matrix 3H is a 

3a 
measure of the sensitivity of acceleration 



28 



measurements to changes in the parameters. 
For estimating parameters, a well designed 
excitation is obviously one which causes 
the elements of the P matrix to decrease 
rapidly. If any elements of P are de- 
creasing slowly or not at all , then a dif- 
ferent type of excitation is needed. A 
look at which elements of P are causing 
the trouble will give a clue as to which 
modes of the system are not being properly 
excited. 

Filtering the Response Data 

In practice, we usually have some 
knowledge of the character of the re- 
sponse data. For example, because of the 
damping present in physical systems, the 
true response will not contain much 
energy at high frequencies. We also know 
that the acceleration is the derivative 
of the velocity which is in turn the de:- 
rivative of the displacement, etc. so that 
these responses are not independent. 

To remove high frequency noise with- 
out effecting the signal a zero phase 
shift low band pass digital filter was 
used. This filter completely removes all 
of the signal and noise above a certain 
termination frequency ut without phase or 
amplitude distortion below a cutoff fre- 
quency 6) c . The digital filter used, due 
to Graham, Reference 10, generates the 
smoothed data as a numerical convolution 
of the raw data and a set of numerical 
smoothing weights, i.e., 



N 



f(t„ + iAt) 



= y~] w(jAt)f(t + 

j=-N 

(i + j)At) 



(13) 



where f(t Q + (i + j )AtJ are the sampled 
values of the signal, f(t + iAt) are the 
smoothed sampled values and where the 
smoothing weights are given by 



w(jAt) = 



sin u-tJAt + sin <» c jAt 



2jAt 



(<fl t -w c ; 3 



) n At 



j = -N,...,+ N 
J * o 



w(o) 



c((«3 t + w c ) 
2i 



u c < (o t 



(1<4) 



where the constant c is chosen to satisfy 
the constraint 



+N 



w(jAt) = 1 



(15) 



j=-N 



The continuous weighting function 
w(t) , of which w(jAt) is a discretization, 
has the Fourier transform, i.e., fre- 
quency domain representation, shown in 
Figure 1. Convolution of this function 
with an arbitrary signal will obviously 
result in a smoothed signal which has all 
frequencies above m-\- completely sup- 
pressed and all signal components below 
w c undistorted. If w c and m-^ are pro- 
perly selected then response data with 
low frequency signal and high frequency 
noise can be improved via digital fil- 
tering, that is, signal to noise ratio 
can be significantly increased. 

In using the digital filter, it is 
tempting to achieve a "sharp" filter by 
taking o) c i w-f Graham, Reference 10, 
has determined empirically that the 
number of points N needed to achieve a 
given accuracy is approximately inversely 
proportional to |wt~ u cl at least over a 
limited frequency range. Since N = 40 
points were used to filter the data, we 
selected | oj^-a)-. |>_ 1 which according to 
Graham is sufficient to yield 2% accuracy. 

In this study, the numerical convo- 
lution was accomplished by using a moving 
average, i.e., f(t '+ iAt) was computed 
separately for each i using Eq. (13). For 
long data records it is possible to 
achieve considerable savings in computer 
time by using the Fast Fourier transform 
algorithm to do this convolution, see 
Reference 11. 

Improvements in the response data 
can also be obtained by making use of 
relationships among the various response 
signals. For the coupled flapping-torsion 
system considered in the next section the 
displacements £, velocities n and accel- 
erations c are related by 



5 = n 

n = c + v 



(16) 



We can use these equations as process 
equations in a. Kalman filter along with 
measurement equations 



s + 

W n v 2 



(17) 



29 



where 5 and IT denote smoothed measured 
values. In the process equation (16) 
replace ? by its smoothed measured value 
X and let R, the process noise covariance 
matrix account for remaining errors . 
Then the Kalman Filter is given by 



+ PR 



-1 



5-5 

ff - n 



(18) 



Note that n is available when solving the 
above equations and can be used as an 
improved estimate of X- Although this 
technique has not been used in this 
study, a similar procedure has been used 
successfully in Reference 6 for heli- 
copter derivative identification. 

Computer Experiments 

Coupled flapping-torsion vibrations 
of a rotor blade at high advance ratio 
are governed by the equations 

8 + P 2 B = \ CM 9l (t)6 + M x (t)X + 



M e (t)9< 



C(t)B - K(t)g] 



I + f 2 « = 4 e 



3 Y F [C e (t)9 + 



C 6 (t)6] 
3YQU r 8(t)B + * r6 (t)B + fc rX (t)A + 



*re (t)e o + K 6 (t)5] 



(20) 



where the periodic coefficients are 
defined in Reference 7 . Responses to the 
gust excitations shown in Figure 2 were 
generated by solving Eq. (20) numerically 
using a fourth order Adams Moulton method 
with a time step of .05 and the following 
parameter values: 



p 2 = 1.69 
f 2 = 64. 
B = .97 



4.0 



U = 1.6 



F = .24 



Q = 15. 



6 = 0. 



(21) 



Simulated noisy measurements were obtained 
by adding samples from zero mean computer 
generated Gaussian random sequences to 
the computer generated responses. First 
the noise was added only to accelerations 
using the standard deviations 



1.0 



°S = 10 



(22) 



The following three parameters with the 
values 



a = y/2 = 2.0 

b = -3yF = -2.88 (23) 

c = _3 y q s -180 f 

were assumed to be unknown. 

They represent blade flapping and 
torsional inertia numbers. Unsteady 
aerodynamic inflow effects may possibly 
be considered by modifications of these 
inertia numbers from transient rotor 
model wind tunnel tests. The linear 
sequential estimator was started with 
the initial values of the estimates and 
errors of the estimates 



40 
P(o) 55 
4000. 



"a(o)" 


b(o) 


_c(o)_ 



(24) 



The linear sequential estimator is, as 
mentioned before, quite insensitive to 
the initial standard deviations which 
could have been selected still much 
larger. The values for R used are the 
following 



R = 



10 
10 



(25) 



The method allows wide variations in the 
assumptions of the noise covariance 
matrix R. The integration scheme for 
solving filter and covariance Eqs. (9) 
and (10) was again a 4th order Adams 
Moulton method with a time step^of ^. 05. 
Fig. 3, shows the estimates a, b, c 
normalized with the true values and the 
3 diagonal terms of the error covariance 
matrix P normalized with the initial 
values vs. non-dimensional time t. The 
excitation for this case was a unit step 
gust at time t = o, as indicated in Fig. 
2 by the dash line. In about one 
revolution (t = 2tt) the diagonal compo- 
nents of the covariance matrix P a P5 P c 
are approximately zero and further 
improvements of the parameter estimates 
a 6 c are not obtained. There is a 
small bias error (deviation from the 
value 1) in two of the parameters , which 
have been recovered within about 5% error. 

The next case assumes that not only 
the accelerations but also the state 
variables are noisy. The following 
standard deviations were used 



30 



a B = .2 
■"l = - 6 

<S-A = 1.0 



.5 
3.0 



(22a) 



qj = 10 



The linear sequential estimator was first 
applied to the raw data. In this case 
the responses are far from smooth so' that 
the use of a high order numerical inte- 
gration scheme was unjustified. A first 
order Euler's method was used for the 
integration of the estimator equations. 
The initial values were 



|"a<o> 
Mo) 
c(o) 



P(o) 



30 











35 
1000 



The values for the R used in the 
estimator were 



R = 



16 





22S 



(24a) 



(2 5a) 



The excitation consisted of a upward unit 
step gust at t = 2.0 followed by a down 
step gust to X h -1 at t = 6.0, as 
indicated in Fig. 2. The second gust 
was added in order to provide to the 
system another transient useful for the 
estimator process. Fig. 4 shows that 
though two of the diagonal covariance 
terms go to zero after the second gust, 
the associated parameter estimates 
remain quite erroneous. The linear 
sequential estimator cannot be used if 
noise is present in accelerations as well 
as in the state variables. 

Next the same data were passed 
through a digital filter with cut-off 
frequencies w c = 12, u-t = 13, see Fig. 1. 
These cut-off frequencies are about 50% 
higher than the torsional frequency of 
f = 8 . Applying now the linear 
sequential estimator to the filtered data, 
the initial values were the same as 
before, Eq. (24a) , however R was reduced: 



R = 



1 
9 



(25b) 



The results of the estimation are shown 
in Fig. 5. All diagonal terms of the 
covariance matrix go to zero soon after 
the second gust, the estimates stabilize 
in less than 2 rotor revolutions and have 
only a small bias error of about 5% ; same 
as for the case with zero noise in the 
state variables. Digitally filtering the 



data to remove high frequency noise has 
thus appreciably extended the range of 
applicability of the linear sequential 
estimator. It might be argued that the 
success of the digital filter is due to 
the "white" character of the computer 
generated noise whereas real data will 
contain energy only at finite frequencies. 
It should be noted that the digital 
filter removes all of the signal above 
the truncation frequency and hence 
would be equally successful for any 
other distribution of the energy above 

In selecting the parameters for the 
digital filter it is important to keep 
w c large enough so that the responses 
are not significantly distorted. 
Initially, the noisy data was pro- 
cessed using different digital filters 
for the torsion and flapping responses. 
A digital filter with high cut-off 
frequency i.e., io„ = 12. and oa t = 13. 
was used for torsion responses while a 
lower bandpass filter with ui_ = 2. and 
uk = 3. was used to filter flapping 
responses. This resulted in poor identi- 
fication of the parameter a in the 
flapping equation. When the same 
filter with high cut-off frequency was 
used for all of the data, adequate 
identification of all parameters was 
obtained. Although w c = 2. is above the 
natural frequency of flapping vibration, 
the flapping response obviously contains 
higher frequency components because of 
the coupling with torsion. This can 
easily be seen by inspection of the 
flapping response in Figure 6. For a 
good identification it is necessary that 
these higher frequency components not 
be removed from the signal. Fig. 6 
compares the response without noise to 
the response with noise but after 
filtering. Also indicated are the 
standard deviations for flapping and 
torsion before filtering. It is seen 
that the filter was very effective in 
removing the noise corruption from the 
data. 

Conclusions 



2. 



The linear sequential estimator, also 
called equation of motion estimator, 
has been successfully applied to 
recover the system parameters of a 
periodic system representing rotor 
blade flapping-torsion dynamics at 
high rotor advance ratio with noise 
contaminated accelerations . 
Filtering of the noisy acceleration 
data was found to be not necessary. 

If noise is present in the state 
variables as well as in the acceler- 
ations, the linear sequential 



31 



estimator performed very poorly. 

3. Filtering both state variables and 
accelerations with a Graham digital 
filter with a low cut-off frequency 
for flapping and a high cut-off for 
torsion before estimation lead to a 
poor estimate for the flapping 
parameter. 

t. Filtering both flapping and torsion 
response with a high cut-off fre- 
quency digital filter before esti- 
mation resulted in an adequate para- 
meter recovery both in flapping and 
in torsion. 

5. As compared to non-linear estimation 
methods which are applicable also if 
acceleration information is not 
available, the linear sequential 
estimator has the great advantage of 
being insensitive to the assumption 
of initial values for the estimate 
and for the error of the estimate. 

No matter what the actual measurement 
noise is, the assumed noise covar- 
iance matrix should be over-rather 
than underestimated. 

6. As compared to the usual form of the 
least square estimation the linear 
sequential estimator does not re- 
quire the inversion of large matrices 
but merely the numerical solution of 
a system of ordinary differential 
equations, thus allowing on-line 
application. The digital filter 
smoothes the data sufficiently so 
that high order of accuracy predictor 
corrector methods can be used for 
the integration. 

7. The computer studies were performed 
assuming rather large measuring errors 
with standard deviations for the 
deflections of about 10% of the maxi- 
mum measured values. The foregoing 
conclusions assume the absence of 
modeling errors , which would require 
special investigations. 

References 

1. Curtis, H.C. Jr., COMPLEX COORDINATES 
IN NEAR HOVERING ROTOR DYNAMICS, 
Journal of Aircraft Vol. 10 No. 5 ,May 
1973, pp. 289-296. 

2. Berman, A. and Flannelly, W.G., THEORY 
OF INCOMPLETE MODELS OF DYNAMIC 
STRUCTURES, AIAA Journal, Vol. 9 No. H, 
August 1971, pp. 1481-87. 

3. Dales, O.B. and Cohen, R. , MULTI- 
PARAMETER IDENTIFICATION IN LINEAR 
CONTINUOUS VIBRATING SYSTEMS, Journal 
of Dynamic Systems, Measurement and 



10. 



Control, Vol. 93, No. 1, Ser. G. 
March 1971, pp. 45-52. 

Sage, A. P. and Melsa, J.L., SYSTEM 
IDENTIFICATION, Academic Press, 
New York 1971. 

Chen, R.T.N. , Eulrich, B.J. and 
Lebacqz, J.V. , DEVELOPMENT OF 
ADVANCED TECHNIQUES FOR THE IDENTIFI- 
CATION OF V/STOL AIRCRAFT STABILITY 
AND CONTROL PARAMETERS, Cornell Aero- 
nautical Laboratory Report, No. 
BM-2820-F-1, August 1971. 

Molusis, J. A., HELICOPTER STABILITY 
DERIVATIVE EXTRACTION FROM FLIGHT DATA 
USING THE BAYESIAN APPROACH TO ESTI- 
MATION, Journal of the American 
Helicopter Society, Vol. 18, No. 2, 
April 1973, pp. 12-2 3. 

Sissingh, G.J. and Kuczynski, W.A. , 
INVESTIGATIONS ON THE EFFECT OF BLADE 
TORSION ON THE DYNAMICS OF THE 
FLAPPING MOTION, Journal of the 
American Helicopter Society, Vol. 15, 
No. 2, April 1970, pp. 2-9. 

Hohenemser, K.H. and Yin, S.K. , ON THE 
QUESTION OF ADEQUATE HINGELESS ROTOR 
MODELING IN FLIGHT DYNAMICS, Pro- 
ceedings 2 9th Annual National Forum 
of the American Helicopter Society, 
Washington D.C. , May 1973, Preprint 
No. 7 32. 

Bryson, A.E. and Ho, Y.C., APPLIED 
OPTIMAL CONTROL, Ginn S Co., Waltham, 
Mass. , 1969, p. 376. 

Graham, R. J. , DETERMINATION AND 
ANALYSIS OF NUMERICAL SMOOTHING 
WEIGHTS, NASA TR R-179, December 196 3. 



11. Gold, B. and Rader, C. , DIGITAL PRO- 
CESSING OF SIGNALS, McGraw-Hill, 
New York, 1969. 

12. Ried, W.T., RICATTI DIFFERENTIAL 
EQUATIONS, Academic Press, New York, 
1971. 

Appendix A 

Solution of the Covariance Equation 

The covariance equation of the linear 
sequential estimator 

is a matrix Ricatti differential equation. 
It is well known that the general matrix 
Ricatti Equation with all matrices being 
time functions 

P = -PA - DP - PBP + C (A-2) 



32 



of which (A-l) is a special case, has the 
solution 



p = VU x 
where U and V satisfy 
V = CU - DV 
U = AU + BV 



(A- 3) 



(A-4) 



This and other aspects of matrix Ricatti 
equations are discussed in Reference 12. 



By comparing Eqs. (A-l) and (A-2) we 
see that Eq. (A-l) is of the form of Eqi 
(A-2) with A=C=D=0 and B = /3H\ T R _1 3H 

laa) 3a 

Therefore, from Eq. (A-t) V = V Q , a con- 
stant matrix and 



U = BV Q 
Integrating yields 



U = U + / 
Jo 



Now since from (A- 3) 



V Q Uo 



B dtV, 



-1 



(A- 5) 



(A- 6) 



(A- 7) 



we can satisfy the initial, condition by 
taking V = I and U 



Po" 1 - 



Hence 



U = ? r 






B dt 



(A- 8) 



and 



['.- 1 * /(-S) 1 *- 1 



-1 



15 

3a 



dt 



(A- 9) 



Minimizing the cost function Eq. (6) 
with w = o, x = a and z = c, one obtains 
the least square estimate 

* ■ Cp o _1 + /({if*" 1 42 dtrl[p o _1 ao < 



!5V 

3a/ 



R _1 Cdt] 



(A-10) 



where the first factor is the covariance P 
from Eq. (A-9). Eq. (A-10) is the equiv- 
alent of solving Eqs. (9) and (10) and 
has been used in Ref. 6 with P _1 = o 
after replacing the integrals by sums. 
In this case the result is independent of 
R which cancels out. 

Even in the general case of finite 



P(o) the error covariance matrix R need 
not be considered as a separate input. 
If R is a diagonal constant matrix it is 
evident that Eqs. (9) and (10) can be 
written in the form 



i ■ '-Off 1 



C- H(5,a)l 



P r = 



3H p 
3a ** 



(A-10) 



(A-ll) 



,-1 



where P r = P R J ". This was pointed out 
to the authors by John A. Molusis. 



H 



L 



■« t - w c 



Fig. 1. Fourier Transform of 
Weighting Function 



CdL 



(U, 





Fig. 3. Estimates 6 Covariance s vs. 

Time, Acceleration Noise Only 



33 



A 
10 



-1.0 



^ — *- 



-£ ^ ^ t 



Fig. 2. Gust Excitations 






Fig. 5. Estimates S Covariances, 
Filtered Data 





Fig. i». Estimates S Covariances vs. Time, 
Acceleration and State Variable 
Noise 




Fig. 6. Exact and Filtered Noisy Responses 
(Solid S dash line respectively) 



34 



DYNAMIC ANALYSIS OF MJlTI-DfiGHEE-OF-FEEEaX)M SYSTEMS 
USING HffiSIKG MATSICES 

Kiehard L. Bielawa* 

United Aircraft Research laboratories 

East Hartford, Connecticut 



Abstract 



A mathematical technique is presented for 
improved analysis of a wide class of dynamic and 
aeroelastic systems characterized by several 
degrees-of-freedcm. The technique enables greater 
utilization of the usual eigensolution obtained 
from the system dynamic equations by systematizing 
the identification of destabilizing and/or 
stiffening forces. Included, as illustrative 
examples of the use of the technique, are analyses 
of a helicopter rotor blade for bending- torsion 
divergence and flutter and for pitch-lag/flap 
instability. 

Notation 



[A], [B], Inertia, damping and. stiffness matrices, respec- 
[0] tively, Eq.. (1) 

A,v Elton-lag coupling for k'th edgewise mode 
(=A9/A4 Vi ) 

A,, Pitch-flap coupling for m'th flatwise Bode 

a Section lift curve slope, /rad 

a i1' *11* Elemen ' , ' s of tllfi CA], [b] and [0] matrices 



"% 



Viscous equivalent structural damping of k'th 
edgewise mode 

Blade chord, in. 



EL., E^ Flatwise and edgewise bending stiffness, respec- 
tively, lb-in. 2 

Jp(t)} Dynamic excitation force vector, Eg.. (1) 

f n Eesultant driving force for n'th degree of freedom, 
E<1. (5) 

[g(Xj)] ljynamic matrix for i'th eigenvalue, Eq. (3) 

2 



GJ 



"no' 

k "10 



"o 



Torsional stiffness, lb-in. 

Boot feathering spring, in.-lb/rad 

Polar radius of gyration of spar about its center, 
in. 

Section thickness-wise and chordwise mass radii of 
gyration, respectively, about spar center, in. 

o 2 
Section mass distribution, lb-sec /in. 

Reference mass distribution, (= 0,000776 lb-sec 2 /in. ) 



*Senior Besearch Engineer, Botary Wing Technology 
Group. 



!>Ai3> 
t?Ai]» 

Iwi 

4fi4 



E 
r 
1 
t 

yio cg , 

y 10 c /4» 
y l°3cA 



8, 



'Vfc 
'wi 

\ 
*±a 

V 
9 

»e 
X 

*1 
P 



{*<*>} 



35 



"Stability" Force Phasing Matrices for i'th eigen- 
value, Eq.s. (6) through (8) 



"Stiffness" fore* Phasing Mvtsrtces for i'th 
eigenvalue, Eqs. (10) through (13) 



k'th edgewise modal response variable 

i'th flatwise modal response variable 

j'th torsional modal response variable, (J = 1, 
for rigid feathering) 

Hotor radius, in. 

Blade spanwise location, in. 

Tension at r, lb 

Time, sec 

Vector of degrees of freedom 

Chordwise positions forward of spar center of mass 
center, quarter chord, and three-quarter chord, 
in. 

Spanwise variable section angle of attack about 
which perturbations occur, rad 

Blade pre-coning angle, rad 

Angle defined in Fig. 1 (= arg X ± ) 

k'th assumed edgewise mode shape 

i'th assumed flatwise mode shape 

j'th assumed torsion mode shape 

Coefficients describing quartie variation of profile 
drag coefficient with angle of attack 

Kronecker delta 

Dumber defined in Eq. (9) 

Geometric (collective) pitch angle at r, rad 

Elastic torsion deflection at r, rad 

(Uniform) rotor inflow 

i'th eigenvalue, /sec 

Air density, lb-sec e /in. 

Blade solidity 

Real part of i'th eigenvalue, /sec 

i'th eigenvector of dynamic matrix equation 

Hotor rotational speed, rad/sec 

Imaginary part of i'th eigenvalue, /sec 



<*) 

( )• 
(~) 

E 3 



Differentiation with respect to (fit) 

Differentiation with respect to T 

Indicates quantity is nondimensionalized using 
combinations of R, iHq and ft, as appropriate 

Diagonal matrix 



I. Introduction 

dynamic and aeroelastie analyses of aerospace 
structures typically involve deriving and solving 
sets of linear differential equations of motion 
generally written in matrix form: 



[A]{x} + Cb]{x} + [C]{xj- = |p(t)} (1) 



In general, the A, B and C matrices are square 
and real- valued. A recognized hallmark of rotary 
wing and turbomachinery dynamics is an abundance 
of nonconservative forces (usually involving rotor 
rotation speed). Consequently, the resulting 
analyses produce matrix equations of motion of 
the above type which are highly nonsymmetrical, 
and often of large orders. 

Although a large part of the dynamic analyst's 
job involves the calculation of dynamic loads and 
stresses due to explicit excitations, the scope 
of this paper will be limited to the equally 
important eigenproblem (F(t) = 0): 



This paper presents an easily implemented 
technique for the improved analysis of dynamic 
systems of the type described above. The technique 
requires a reliable eigensolution and involves 
manipulations of the given dynamic equations, 
their eigenvalues and eigenvectors. Specifically, 
the technique systematizes the identification of 
destabilizing and/or stiffening forces by the 
calculation of "force phasing matrices". Applica- 
tions of the technique to analyses of bending- 
torsion divergence and flutter and of pitch- lag/ 
flap instability of a helicopter rotor blade are 
presented. Furthermore, this paper essentially 
represents an expansion of a portion of an earlier 
paper ^ . ' 

II. Mathematical Development 

The principal function of the force-phasing 
matrix technique is to identify those force terms 
in the equations of motion which, for an unstable 
mode, are so phased by the mode shape as to be 
drivers of the motion. The technique is perhaps 
nothing more than a formalization of the intuitive 
use an experienced dynamicist would make of the 
eigenvector information. The basis of the tech- 
nique can be seen by writing any single equation 
of the set represented by Eq. (3) as the sum of 
the mass, damper and spring forces of the diagonal 
degree-of-freedom and the remaining forces acting 
as a combined exciting force. 



W -sK'}- 



a X?tpi 15 +t> ^•<i i) + c tPn 15 
nn i T n nn iti nn T n 

/£> nji nj i nj' Y j 



(5) 



X.t 



(2) 



[cA]xf + [B]X 1+ [c]l{cp (i >} = [GCXi)]^} - {0} 

(3) 

The eigenvalues X (= cr+io)), which give stability 
and natural frequency information are obtained 
from the familiar characteristic determinant: 



|[A]X 2 + [B]X + [C]| . 



(*0 



*y 



(1) 



by any of various well-established methods v ', 
(2), w) # jije "flutter" mode shapes, <ffc\ are 
obtained from Eq. (3) once the eigenvalues are 
known. 



For the usual case a^, bj^ and c nn are all 
positive numbers; that is, each mass when un- 
coupled from the others is a stable spring-mass- 
damper system. Since the root, k±, is generally 
complex, Eq. (5) can then be interpretted as the 
sum of four complex quantities or vectors in the 
complex plane which must, furthermore, be in 
equilibrium. Assuming that the root with 
positive imaginary part is used throughout, the 
argument of the root, 7^, is the angle by which 
the inertia force vector is rotated relative to 
the damper force vector and the damper force 
vector is rotated relative to the spring force 
vector. For an unstable root this angle will 
be less than 90 degrees. If a purely imaginary 
value is assigned to the spring force vector, 
unstable motion is assumed and it is recalled 
that the four vectors are in equilibrium, then 



36 



the real parts of the damper and inertia force 
vectors will be negative and the driving force 
must always have a positive real part. Figure 1, 
which demonstrates this argument, shows the four 
force vectors in the complex plane for an 
unstable oscillatory mode (Re(A^) = O" i >0) and 
for unit imaginary displacement: 



vwv 



(SPRING FORCE) 




7j '(mMXi) < 90deg. 



(INERTIA FORCE) 



Figure 1. Force- Vector Diagram for n'th Degree- 
Of-Freedom, i'th Mode (Oscillatory 
Instability) 

A secondary function of the technique is to 
identify those terms in the equations which, for 
any coupled mode, act as stiffness so as to 
increase the coupled frequency of the mode. 
Reference to Figure 1 shows that driving forces 
with positive imaginary parts will tend to rein- 
force the diagonal spring term and, hence, raise 
the frequency of the coupled mode. An interesting 
observation that can be made from Figure 1 is that, 
for unstable motion, the diagonal damper force 
also has a positive imaginary part. Hence, it 
tends to stiffen the (unstable) coupled mode in 
contrast to the frequency lowering effect of 
damping for stable motion. 

Figure 2 shows the same forces as vectors 
for an unstable aperiodic mode (divergence) for 
negative unit real displacement: 






SPRING, DAMPER AND 
INERTIA FORCES 



DRIVING FORCE 



Figure 2. Force-Vector Diagram for n'th Degree-of- 
Freedom, Divergence Instability. 

Again, the driving force is always a positive 
real number. Furthermore, for divergences, stif- 
fening forces are by definition stabilizing; hence, 
those components of the driving force which are 
negative are also those that stiffen the coupled 
mode. 

These interpretations of unstable motion can 
be quantitatively implemented first, by multiplying 
each of the dynamic equations (i.e., each row of 
the equation (l)) by a quantity which makes the 
diagonal stiffness force (stiffness matrix element 
x displacement) become pure imaginary and second, 
by representing the modal vector as a diagonal 
(square) matrix. This latter operation has the 
effect of evaluating the magnitudes of the com- 
ponent dynamic equation forces without numerically 
adding them together. The resulting "stability" 
force phasing matrices are then readily written as: 



^ 



ffleCn/q^axftAtfq* 1 ^ 



Cp b J --fleCTi/qf^CBH 9 (1) 3 



Cp_ ] = (Ret v/y a> Kelt <p c1) 
c i 



(6) 



(7) 



(8) 



where 



( i j for oscillatory instabilities 
(-1. , for divergences 



(9) 



and where the eigenvalue in the upper half plane 
is used. 



37 



la all cases, the real parts of the above 
indicated matrix expressions give instability 
driving force information. Forces defined by- 
elements of the A, B and G dynamic equation 
matrices which are phased by the mode shape so 
as to be drivers of the motion then cause the 
corresponding elements of the Pj^, Pg^ and P(Ji 
"stability" force phasing matrices, respectively, 
to be positive and proportional to their strength 
as drivers. 

Stiffening driving force information is 
obtained differently for oscillatory motion and 
for aperiodic motion. Those elements of the 
dynamic equation matrices which are phased so as 
to be stiffeners of the coupled mode will cause 
the corresponding elements of the matrix expres- 
sions to be either positive imaginary for oscil- 
latory motion, or negative real for aperiodic 
motion. The resulting "stiffness" force phasing 
matrices are then expressed as: 



[P Ai ] = JmF V<p (i) 3^WC <P (i) 3 

[P B .] = JmC iy<P (i) ^i[B]£ cp( 1 )^ 



cp c .] = jmf uv^rae v (x) i 



for oscillatory motion, and: 



(10) 
(11) 
(12) 



[*( )J - CP( )J 



for aperiodic motion. 



(13) 



It should be stressed that these force 
phasing matrices are no more than a more system- 
atic and efficient interpretation of the all too 
often voluminous eigensolution information. The 
following sections illustrate the usage of the 
force phasing matrix technique in substantiating 
what is generally known of some rather fundamental, 
classical helicopter rotor blade instabilities. 



III. 



Description of Illustrative 
Rotor Blade Example 



For illustrative purposes, relatively simple 
linear equations of motion were formulated for a 
generalized untwisted helicopter rotor blade and 
then applied to a realistic nonarticulated rotor 
configuration. The blade is assumed to be oper- 
ating in an unstalled hover condition at some 
collective angle and with a built-in coning angle. 
Perturbative elastic flatwise, edgewise and torsion 



motions are assumed to occur about the preconed 
position. The resulting linear aeroelastic 
equations are fairly standard *■''> ^ '; quasi- 
static aerodynamics (uniform inflow) is assumed 
and a normal mode description of the blade 
elasticity is employed. Thus, for the chosen 
configuration, two flatwise bending modes, one 
edgewise bending mode, and the rigid feathering 
degree-of- freedom are assumed. The resulting 
response vector, |xj- , consists of the quantities 
Iwij 4w2> 4vi> and- qg^ whose detailed dynamic 
equations are given in the Appendix. The dynamic 
equations then comprise a set of four differential 
equations written as a k x k matrix equation of 
the Eq. (l) type. The aeroelastic degrees-of- 
f reedom together with the general parameters are 
shown in Figure k: 



TOTAL 

PITCH 

ANGLE, 




Figure k. 



PRE-CONING- 



Schematic of Nonarticulated Eotor 
Configuration and Aeroelastic Degrees- 
of -Freedom. 



The basic configuration incorporates a 
counterweight over the outer 70 percent of the 
blade, pitch- flap coupling (determined from the 
geometry of the pushrod attachment and flatwise 
modal deflection) and pitch- lag coupling of 
arbitrary magnitude . The chordwise position of 
the counterweight and the magnitude of pitch- lag 
coupling are purposely varied in the following 
analysis to establish known blade instabilities 
in order to illustrate the phasing matrix 
analysis technique. Table I below summarizes 
the pertinent geometric and aeroelastic data 
for the rotor blade configuration: 



38 



TABIE I - BLADE CHARACTERISTICS 



17. 



Hadlua 
" Chord (0.1H to tip) 
lip Speed 

Siton-Flap coupling, A v ta |n = | 
Hoof feathering spring rate 
Blade coning 

Airfoil: (HACA 0012; Mach Ho. - 0): 
a 
So 
*U 

(Uniform) maBS distribution 
(Uncoupled) blade natural frequencies: 

first flatwise mode 

second flatwise mode 

first edgewise mode 

rigid feathering 



210 in. 
13.5 in. 

650 fps 

0. 
0.188 

3.55 x 10 6 in.-lb/rad 
2 deg. 

6.0/rad 

0.01 

3.30/^10^ 

O.OOO776 lb-sec 2 /in. 2 

1.09e/rev 
2.68l/rev 
1.390/rev 
3.820/rev 



(edgewise mode structural damping) 0.01 
(critical damping) 



Flight condition (hovering) 
collective angle, 8 
inflow ratio, X 



sea level, standard 

10 deg. 

-.0601 



The normal flatwise and edgewise mode shapes 
used are shown In Figure 5. 



1.0 r 



0.6- 




-0.6*- 



Figure 5. 



3pamd.se Variation of Normal Mode 
Shapes. 



Application of Analysis to 
Illustrative Example 



Basic Configuration 

For purposes of comparison, the data in 
Table I was used together with a collective angle 
of 10 deg, inflow ratio of -0.0601, and with zero 
counterweight chordwise offset (from the quarter 
chord), and zero pitch- lag coupling. This basic 
case is stahle in all modes as is shown hy the 
following list of resulting eigenvalues: 



^1,2 = 

J 5,6 = 

X 7,8 = 



-0.5CA- ± iO.960 
-O.lHl ± 12.610 
-0.027 ± il.398 
-1.14-72 ± i3-506 



While all the aeroelastic modes represented by 
the various eigenvalues comprise responses in all 
the four degrees of freedom, they could be 
characterized as follows: mode 1 ( X^ 2) is first 
flatwise bending, mode 2 ( X3 14.) is second 
flatwise bending, mode 3 ( \<j } 6) is first edgewise 
bending, and mode k ( X7 g) is rigid feathering. 

Configuration With Rearward Chordwise 
Counterweight 

If the 70 percent outer span counterweight^ is 
artificially shifted aft so as to place the chord- 
wise section mass center at the 32 percent chord 
point, reference to the dynamic equations (A-l, 
A-2 and A-3) then yields the following A, B and 
C matrices (given in E format): 



A (Biertia) Matrix: 

.2909-00 -.0000 
-,0000 ,2006-00 

*.oooo -.0050. „ 

-.1884-02 ,3162-03 



B (Damping) Matrix: 



•,0000 

',0000 
,25*2-08 
•tOOOO 



«, 1884-02 
,3482*09 

•.0000 
,1133-03 



.2B87-00 
.7992-01 
.3099*01 
,1536-07 



,7757-01 
,1665-00 
,1943-01 
,6326-01* 



-, 1*500-02 -, 1256-01 

•,6166-02 -.$309-03 

,7901-02 -.2020-02 

•,1068-03 ,3370-03 



C (Stiffness) Matrix: 



39 



.3382-00 -.5763-01 ,4633-01 -.3067-00 
-.8232-03 , 1^22+01 ,7464-02 -.7080-01 

.4633-01 -,2974-02 ,4968-00 -.1396-01 
-, 2300-02 -,5423-02 -,2919-03 ,1646-02 



The first two rows of the matrix equation 
are the equations for the first and second 
flatwise bending modes, respectively. The third 



row is the equation for the first, edgewise mode, 
while the fourth is for rigid feathering. 
Correspondingly, elements in the first two columns 
of the matrices are terms multiplying the flatwise 
■bending responses and their derivatives. Simi- 
larly, third and fourth column elements are terms 
multiplying edgewise bending and rigid feathering, 
respectively, and their- derivatives . 

The eigensolution for these matrices (see 
Eq. (k)) reveals the configuration to he unstable 
in "both divergence and flutter: 



*1,2 

*3,4 

X 7,8 



0.1(08, -4.466 

0.300 +1I.789 

-0.0088 ± il.402 

-0.578 ± i3.099 



Since the equations are in nondimensional form, 
the units of these eigenvalues are per rotor 
revolution frequency or "p". Using Eqs. (6) 
through (9) the following "stability" force 
phasing matrices are written for the unstable 
divergence mode: 

A Phasing Matrix, Paj_ 

-.1*840-01 -.0000 -.0000 ,5061-03 
-,0000 -.3338-01 -.0000 -,1726-02 
-.0000 -.0000 -.4229-01 -.0000 
.1941-03 -.1944-05 -.0000 -.1885-04 



B Phasing Matrix, Pg^ 

-.1178+00 -.1715-02 .4408-04 ,3272-02 
-.6016-00 -.6791-01 .1476-02 .4022-02 
-.5265-00 -.1789-01 -.3223-02 .5540-01 
-.3681-08 -.8660-06 .6418-06 -.1375-03 



C Phasing Matrix, P^, 

-.3382-00 ,3123-02 -.1113-02 ,4953-00 

,1519-01 -.1422+01 -.3316-02 .2110+01 

-.1929+01 .6710-02 -.4988-00 ,2956+01 

.1424-02 .1620-03 ,4341-05 -.1646-02 



The larger of the destabilizing driving 
forces, which show up as positive terms, have been 
underlined for clarity. A reasonable yard-stick 
for measuring the size of the destabilizing forces 
is to compare them to the size of the stabilizing 
element in each matrix equation row. For 
oscillatory instabilities that element would be 
the diagonal damping force; for divergences, it 
would be the diagonal stiffness force. As would 
be expected of a divergence instability, the 
major destabilizing forces are displacement de- 
pendent (i.e., appear in the C matrix). By making 
additional reference to Eqs. (A-l) and (A-2), and 



(A-3) and to their evaluation given above, the 
following interpretation can be drawn from these 
results : 

1. The unstable mechanism involves a 
coupling mainly between first flatwise bendirig 
and rigid feathering. The mode shape, ^w = 
(O.619, O.O336, 0.0149, and 1.0), confirms this 
result . 

2. The position of the chordwise mass 
center behind the elastic axis, as indicated by 
negative dynamic equation elements a^_ jj. and cij 1 
and reference to the explicit statements of the 
equations in the Appendix, is a major link in 
the unstable coupling chain of events . This 
result confirms well-known results concerning the 
divergence of rotor blades ^J. specifically, 
that the torsion modes drive the flatwise modes 
aerodynamically (elements Cj ^ and c 2 ^) while the 
flatwise modes drive the torsion modes with 
centrifugal inertial forces through the rearward 
mass center position (elements c^ -j. and c k ?)• 

3. The first edgewise bending mode is being 
driven by the rigid feathering through aerodynamic 
and inertia terms (element bo ij.) but is not 
actively participating in the unstable mechanism. 

In a similar manner the following force 
phasing matrices are written for the unstable 
oscillatory (flutter) mode, \~ 



x 3" 



A Phasing Matrix, P^_ 



•, 3125-00 -.0000 -.0000 -.1904-01 

■,0000 -.2155-00 -.0000 -.1101-01 

■.0000 -.0000 -.2731-00 -.0000 

.1421-02 -.1114-04 -.0000 -.1217-03 



B Phasing Matrix, R 



••5163-00 
,6947-00 



B3 

.7587-02 

•, 2973-00 

.1659-02 



,2749-02 
-.1561-01 
-.1413-01 



.1968-01 
,1282-01 
,4186-05 



-.1493-00 
,5201-09 -.5155-05 -.6870-05 -.6027-03 

C Phasing Matrix, Pq_ 

-.1260-08 .6779-02 -.4336-02 .8548-00 

-.4997-02 -.1325-08 .1751-01 -.1802-00 

.3814-01 .1173-02 -.0000 .3956-00 

-.6696-03 .2794-04 -.3155-04 -.0000 

Again the major destabilizing terms have 
been underlined for clarity. With few exceptions 
the same interpretations can be made of the 
flutter force phasing matrices as were made for 
the divergence ones. While the feathering 



40 



\ 



degree-of-freedom again drives the first flatwise 
mode aerodynamically (element c-^ j,.), the flatwise 
mode now drives the feathering degree-of-freedom 
with vibratory inertia forces (element a^ jj • 
Again these results confirm well-known findings. 

Configuration With Pitch-Lag Coupling 

Using the reconfirmed knowledge that an aft 
chordwise center of mass is destabilizing, the 
configuration is altered back to the original 
quarter chord balanced configuration. In 
addition, unit pitch-lag coupling (Ay... = 1.) is 
introduced into the configuration. Hie resulting 
dynamic equations are as follows: 

A (inertia) Matrix 

.2909-00 -.0000 -.0000 -.0000 

-.0000 ,2006-00 -.0000 -.0000 

-.0000 -.0000 .25ft2-00 -.0000 

-.0000 -.0000 -.0000 .1133-03 



B (Damping) Matrix 



.28B7-00 .7757-01 -.1622-01 -.1256-01 

.7992-01 ,1665-00 -.3596-02 -.3309-03 

.3015-01 .1953-01 .1003-01 -.2169-02 

,1536-07 ,6326-04 ,3801-03 ,3370-03 



B Phasing Matrix, Pg 

-.3822-00 ,5003-02 -.2424-01 .8736-02 
.2114+01 -.2204-00 .2079-00 -.4936-02 
.1546-01 -.3974-03 -.1327-01 -.2946-03 

-.3235-07 .6967-05 .4427-03 -.4461-03 



C Phasing Matrix, Pq 



.1260-08 
,2728-02 
,2015-01 
■,0000 



•.4643-03 
.5299-08 
.7210-04 

-.0000 



.3315-00 .7077-01 

-,1886>0i -.2069-00 

-.0000 -.1373-01 

-.0000 -.0000 



C (Stiffness) Matrix 



By referring to the explicit dynamic 
equations given in the Appendix, the following 
observations can be made: 

1. The instability appears very similar to 
classic pitch-lag instability ' '' and is mainly a 
three-way coupling between first flatwise and 
edgewise bending modes and the rigid feathering 
degree-of-freedom. The resulting coupled mode, 

0(3) = (-O.383 - 10.435, 0.015 + i0.024, 1., 

-0.100 - i.0317). 

2. The edgewise bending mode is being driven 
by inertia forces generated by flatwise bending 
motion: coriolis forces proportional to precone 
and flatwise bending rate and forces proportional 
to pitch angle and flatwise bending deflection. 



,3382-00 -.5763-01 -.2563-00 -.30*9-00 
-.8232-03 ,11*22+01 -.5358-01 -.7113-01 

.4633-01 -.2974-02 .4431-00 -.4337-01 
-.0000 -.0000 -.0000 ,1646-0? 

The eigenvalues for this configuration reveal the 
configuration to be unstable in the edgewise 
bending mode: 



-0.573 
-0.408 
Xjjg = 0.0119 
X 7j8 = -1.449 



L l,2 
L 3,4 



iO.977 
i2.609 
il.324 
13-505 



Again, the stability force-phasing matrices 
are formed for the unstable mode ( ka) and the 
larger positive terms are underlined; 

A Phasing Matrix, P& 

-.9137-02 -.0000 -.0000 -.0000 

-.0000 -.6302-02 -.0000 -.0000 

-.0000 -.0000 -.7985-02 .-,0000 

-.0000 -.0000 -.0000 -.3558-05 



3. The flatwise bending mode is being 
driven by aerodynamic forces generated chiefly 
by pitch- lag coupled edgewise bending and to a 
lesser extent rigid feathering deflection. 

4. The rigid feathering degree-of-freedom is 
being driven principally by a centrifugal force 
moment involving chordwise mass radii of inertia, 
pitch angle, and edgewise bending rate. 

The stiffness force-phasing matrices for this 
mode are formed and the significant terms for the 
edgewise bending equation are underlined: 

A Phasing Matrix, P A 

-.5097-00 ,0000 .0000 .0000 

,0000 -.3516-00 ,0000 .0000 

,0000 ,0000 -.4455-00 .0000 

,0000 ,0000 ,0000 -.1985-03 



41 



B Phasing Matrix, Pg_ 



.3*25-02 

-,3696-00 

.1722-01 

-.1*69-07 



.7*27-03 .2799-01 

.1975-02 -.339*-00 

-.6235-03 .1190-03 

-.1870-05 -.1*50-02 



C Phasing Matrix, Pq_ 



,3362-00 

,16*7-01 

-.1777-01 

.0060 



.280^-02 
,1*22+01 
■,*505-0* 
,0000 



.2923-00 
■. 1179*01 
.**M-00 
,0000 



,3780-02 
-.1230-02 
-,'906*-n3 

.3996-05 



■, 1596-00 
.7995-00 
,*327-02 
.16*6-02 



It can be seen that the principal stiffening 
terms are, not unexpectedly, the diagonal mass and 
stiffness terms. Mae only other significant 
stiffening terms are those involving flatwise 
tending rate and deflection which are also the 
drivers of the unstable edgewise motion. 

That the flatwise tending deflection term is 
negative and numerically greater than the rate 
dependent term can be appreciated by noting that 
the unstable coupled edgewise mode frequency, 
1.32*, is lower than the original corresponding 
stable mode frequency, 1,398- 

V. Concluding Remarks 

The "force-phasing" matrices technique 
provides yet another tool for understanding 
dynamic/aerodynamic phenomena. While it does 
not, by itself, indicate stability levels such 
as are provided by the eigensolution, it does 
complement the eigensolution by giving insight 
into the details of the dynamic configuration 
which are not director available from the 
eigenvalues and eigenvectors alone-. Moreover, 
the technique requires, in particular, eigenvector 
information as a starting point. Hence, it is 
inherently incapable of answering the more 
fundamental question of why, for any one mode, 
the eigenvector elements are indeed phased as 
they are. It should also be stressed that the 
technique is a tool to be used with, and in 
support of, engineer/analyst judgement; the 
results have to be Interpreted properly, generally 
in the context of the specific application. 
Finally, the relative simplicity of the formula- 
tion makes the incorporation of the technique in 
any aeroelastic eigensolution program a straight- 
forward and easily implemented task. 



Appendix - Details of Dynamic Equations 

The linear dynamic equations used to 
represent the aeroelastics of the rotor blade in 
hover are formulated using an assumed modal 
approach; the derivation is standard and uses 
the nomenclature of Reference 6. The lineariza- 
tion and subsequent simplifications are based 
upon the following assumptions: 

1. quasi-static, incompressible, nonstalled 
airloads , / 

2. coincident spar center, shear center 
and tension center. 

3. zero twist. 

*. two flatwise bending modes, one edgewise 
bending mode and the rigid feathering 
degree-of- freedom . 

5. normal uncoupled bending mode shapes 
(zero twist and pitch angle). 

The flatwise bending equations are then written 
as: 



J {(^wiYwm) *4 m + (» yi0 0g YwiYe 3 ) *<& 3 
+[2mY Wi (PY Vlc + y 10og sin9Y^.)] l^ 

-(a I i^cose yq^) qe d (a-i) 

•K^VV* 116 oose ) ^ + csy Wl (yio og ooB2e 

-r9 S ine)Y 9;J ] *8j + *gj V^fre^j 



42 



The edgewise equations are written as: 



where: 



-[2 MYwmCPYvfc + yiOcgSine Yvfc)3 fw^ 

-[2 s(Pyio cg Yv k + % s ^ e y^) y 9; .3 3^ 

+(3y Y w sine cos6) q w (A-2) 

v k m m w m 

+ CEi z Y^ n + ry^ n - 5 cos 2 e Yvk Y Tn ] q^ 

+[mY V:k .(?P cose + yio cg sin2e)Y 9; .] q^ 

+ ^ Y^[-r2(2 « - U Jarg)(Y ej qej+ *^% 



+ 2 *<? - ^M 



to + v^ 



The torsion equations are written as: 

,1 



* 



(» yio cg Y 9;j Yw m ) X + M^ 10 + k z 2 10 )Y ejYe k ] *4 k 

+ (2 m k^cose y^J q^ 

+[2 m(py 10cg Y Tn -+ kf^sine Y^) Y e ^ *v n 

+» yiO^C^Yw " sin2 9 Y w )y 9 .] fc 



J cg ™m 



«m" B j J 



+[«w + i&r) Ye* vi + m (^ 10 -J§ 10 ) cos2 9 y 9 ,y 9 ^ 



3 j' a k N z 10 Tio 
+ ( 5 yiO Qg sln e cos 9 Y 9;J Y Vq ) q Vn 
pacR r _o_ 



J'«k J ^k 
(A-3) 



+ ~2^ Ye/-^yiQ cA (Ye k qe k + ****** + VW 

+ * yiO c /u(Y W Jw m - *oY v Jv n + yio3 C /^(Y ek ^ k 
+ %%+ %%)) - iS * yio 3cA (Ye k qe k 
+ A Wa | Wm +A Vn | Vn )]J<a? + ^ 3 ^qe 1 = o 



* = + \/r 



(A-^) 



Beferences 

Wilkinson, J. H. : Ihe Algebraic Eigenvalue 
Problem. Clarendon Press, Oxford, 1965. 

Programmer's Manual: Subroutines ATEIG and 
HSBG. IBM System/360 Scientific Subroutine 
Package, Version III, GH20-0205- 1 *, August 1970. 

Leppert, E. 1., Jr.: A Fraction Series 
Solution for Characteristic Values Useful in 
Some Problems of Airplane Dynamics . Journal 
of the Aeronautical Sciences, Vol. 22, No. 5, 
May 1955. 

Bielawa, S. L. : Techniques for Stability 
Analysis and Design Optimization with Dynamic 
Constraints of Nonconservative Linear Systems. 
AIAA/ASME 12th Structures, Sturctural Dynamics 
and Materials Conference Paper No. 71-388, 
Anaheim, California, April 1971. 

Miller, R. H. and C. W. Ellis: Blade 
Vibration and Flutter. Journal of the 
American Helicopter Society, Vol. 1, No. 3, 
July 1956. 

Arcidiacono, P. J,: Prediction of Rotor 
Instability at High Forward Speeds; Vol. I, 
Differential Equations of Motion for a 
Flexible Helicopter Rotor Blade in Steady 
Flight Including Chordwise Mass Unbalance 
Effects. USAAVIABS Technical Report 68-18A, 
U. S. Army, February 1969. 

Chou, P. C: Pitch- lag Instability of 
Helicopter Rotors. Journal of the American 
Helicopter Society, Vol. 3, No. 3, July 1958. 



43 



SOME APPROXIMATIONS TO THE FLAPPING STABILITY OF HELICOPTER ROTORS 

James C. Bigger s 

Research Scientist 

Ames Research Center, NASA 

Moffett Field, California 94035 



Abstract 
The flapping equation for a helicopter in for- 
ward flight has coefficients which are periodic in 
time, and this effect complicates the calculation 
of stability. This paper presents a constant 
coefficient approximation which will allow the use 
of all the well known methods for analyzing constant 
coefficient equations. The flapping equation is 
first transformed into the nonrotating coordinate 
frame, where some of the periodic coefficients are 
transformed into constant terms. The constant 
coefficient approximation is then made by using 
time averaged coefficients in the nonrotating frame. 
Stability calculations based on the approximation 
are compared to results from a theory which cor- 
rectly includes all of the periodicity. The com- 
parison indicates that the approximation is reason- 
ably accurate at advance ratios up to 0.5. 

Notation 

a blade lift curve slope 

B tip loss factor 

c blade chord 

I blade flapping inertia 

i /T 

kg flapping spring stiffness 

N number of blades 

R rotor radius 

t time, sec 

V forward velocity 

a angle of attack of hub plane 

J3^ flapping of ith blade relative to hub plane 

B vector of rotor degrees of freedom in non- 
rotating coordinates 

B rotor coning angle 

6 lc rotor tilt forward (longitudinal flapping) 

B ls rotor tilt to left (lateral flapping) 

B 2 rotor differential flapping 

V blade lock number, pacRVl ■ 

X eigenvalue or root, nondimensionalized by fi, 

a 3 a 

V rotor advance ratio, V (cos a) /OR 

v flapping natural frequency of rotating blade 

a real part of eigenvalue 

p air density 

iji azimuth angle, Bt 

S2 rotor rotational speed 

to imaginary part of eigenvalue 

Ug flapping n atural frequency of stationary 

blade, ]/Icg7l" 

( ) derivative, d( )/di/i 

(") derivative, d 2 ( )/di|/ 2 

For helicopter stability and control studies, 
it is desirable to use as simple a math model as 
possible while retaining reasonable accuracy, both 



Presented at the AHS/NASA-Ames Specialists 1 Meeting 
on Rotorcraft Dynamics, February 13-15, 1974. 



to reduce computation effort and to gain insight 
into system behavior. However, for a helicopter in 
forward flight, the rotor flapping motion is 
described by a differential equation having coeffi- 
cients which are periodic in time (azimuth) . This 
fact complicates the solution of the equation, 
requiring methods which use considerable numerical 
computation and which give little insight. Thus it 
is desirable to find a differential equation with 
constant coefficients (hence an approximation) 
which adequately represents the forward flight 
flapping dynamics of a helicopter rotor. If such 
an equation is found, all of the well known tech- 
niques for analyzing constant coefficient equations 
may be used. 

The flapping equation may be transformed into 
the nonrotating coordinate frame, as done in 
References 1 and 2, where some of the periodicity 
is transformed into constant terms. This result 
suggests that the use of constant coefficients in 
the nonrotating frame will retain some of the 
periodic system behavior. The constant coefficient 
approximation examined herein is made by using time 
averaged coefficients in the nonrotating frame. A 
comparison is made between the eigenvalues (sta- 
bility) obtained from the approximation and the 
results from a theory which correctly includes all 
of the periodicity. The comparison indicates that 
the approximation is a useful representation of 
helicopter flapping dynamics for both hingeless and 
articulated rotors. This approximation was briefly 
discussed in Reference 1 for one set of rotor 
parameters. The present paper discusses the 
approximation in a more general manner and gives 
more insight into its features, limits, and 
applicability. 

The rotor math model used here is for fixed 
shaft operation and includes only first mode 
(rigid blade) flapping, with spring-restrained 
flapping hinges at the hub center. Flapping 
natural frequency may be matched by selecting the 
spring rate. Thus the only approximations are in 
the use of the aerodynamic terms for rigid blade 
motion. Uniform inflow is used, and for the 
advance ratios considered here (u < 0.5), reverse 
flow effects are not included. 

Equations of Motion 

In this section, the single blade homogeneous 
flapping equation is presented for a rigid, spring- 
restrained, centrally hinged blade. This equation 
is then transformed to a nonrotating coordinate 
frame, using a coordinate transformation which is 
briefly discussed. Insight into the fundamental 
behavior of the rotor is gained by examining the 
hovering (u = 0) eigenvalues of the equation in 
nonrotating coordinates. 



45 



For the single blade, the homogeneous equation 
of motion is 



where 



iL + Mgij + (v 2 + M g )B i = 



M. = I B" + y I B 3 sin i^ 



CD 



= u X B 3 



2l R 2 



HtB 3 cos 4. + p* 4- B z sin 2$ 

D 1 O 



1 + 



"a 2 " in 2 



Note that reverse flow has not been included here. 
Although it could be included, it would not, signifi- 
cantly affect the results for u < 0.5, since the 
additional terms are fourth order in u. 

By a coordinate transformation of the Fourier 
type, the single blade equation may be written in 
terms of nonrotating coordinates. The transforma- 
tion accounts for the motion of all blades, and the 
number of degrees of freedom is equal to the number 
of blades. For example, with a three-bladed rotor, 
the degrees of freedom are coning (all blades flap- 
ping together), rotor pitching (cosine i)» flap- 
ping) , and rotor rolling (sine i|j flapping) . 
Adding a fourth blade adds a differential flapping 
degree of freedom, where blades 1 and 3 flap in one 
direction while blades 2 and 4 flap in the other 
direction. This type of differential motion is a 
degree of freedom with rotors having any even num- 
ber of blades. Adding more blades adds degrees of 
freedom which, in the nonrotating frame, warp the 
plane described by the sine iji and cosine ty 
flapping motion. 

The coordinate for the single blade is p.. 
For a three-bladed rotor, the corresponding 
nonrotating coordinates are 




where B , g, , and g ls are rotor coning, pitch- 
ing, and rolling motions. For a four-bladed rotor, 




where 2 is the differential flapping motion 
discussed above. 

In general, the blade degrees of freedom in 
the transformation are 










i=l 












N 








B nc = 


2 

N 


E 

i=l 

N 


8 i 


cos 


nipj 


B ns = 


2 

N 


i=l 

N 


h 


sin 


rof^ 


B N = 


1 

N 


E 


h 


(-D 


, N eve 


2 




i=l 








Then the motion of the ith blade is 


K 










'■ 



B i ■ e o + £ (e nc cos »*i + e ns sin n *i) + s n ( - 1)i; 



n=i 



K =< 



j (N - 1), N odd 
j (N - 2), N even 



The equations of motion (that is, eq. (1)) must 
also be converted from a rotating to a nonrotating 
frame by a similar procedure. This process is 
accomplished by operating on the equations with the 
summation operators 

IJ(...) > |S(...)cosn+. ) 
i i 



This is virtually the same procedure used in 
Reference 1. 

It may be seen that the transformation 
involves multiplication by sin ty, cos ty, sin 2$, 
cos 2ip, etc. This changes some of the periodic 
terms of the equations in the rotating reference 
frame into constants (plus higher harmonics) due to 
products of periodic terms, and vice versa. 

Performing the indicated operations for N = 3 
yields the following equations for a three-bladed 
rotor. 



J B» . u ^ B 3 sin 3* 2 - u j 

» J B 3 -I.iiiB'eoslt t B" - i 



r B 3 sin 3* 



v 2 u 2 ^ b ! sin 3* -u 2 ^ B 2 cos » 

■J «>♦»'$ B 2 sinW v 2 _ 1 . „ X B 3 cos 3* | (B'.iu 2 B 2 ).|jjB 3 sin3* 

-v 2 jB ! cos3» - J(B»-ju 2 B 2 ).ujB 3 sin3» » 2 - 1 - M J B ! cos 3* 



(2) 



Similarly, operating as above with N = 4, the 
equations of a four-bladed rotor are obtained. 



46 



a -^ B 3 sin 2iJ> 



W z £B z sin2j. 



v z - 1 * y 2 ^r B z sin 4$ 
U Xb 3 cos 2* 



i(B" + iy z B2-|BVcos4*3 

v 2 - 1 - y 2 ^ B z sin 4i^ 
u 2- B 3 sin 2* 



u 2 £ B 2 sin 2* 
u I- B 3 cos 2* 

u J B 3 sin 2* 



(3) 

The thrtee- and four-bladed rotors have similar 
behavior except for the terms which are periodic in 
<fi. The periodic terms are 3/rev for the three- 
bladed rotor, but are 2 and 4/rev for the four- 
bladed rotor. 

The main advantage of the transformed equations 
is that it is easier to express the combined rotor 
and airframe motions because the rotor equations 
are now in a nonrotating reference frame and 
include the motions of all blades. Furthermore, 
rotor motions are more intuitively understood, 
since the degrees of freedom are those seen by an 
observer in or beside the helicopter. 

In the nonrotating coordinates of equa- 
tions (2) and (3), the equations are coupled by 
off-diagonal terms. Note however, these are actu- 
ally independent blades (unless some sort of feed- 
back is added) and the coupling is due to the 
coordinate transformation. 

To gain understanding of these degrees of 
freedom, the hovering (v = 0) behavior is examined 
next. The hover equations for four blades are 
given below. 



i* 



| B 4 2 



J 



B* 



IB* 



v 2 - 1 $ B* 
- J B* v 2 - 1 





8 ■ 



(4) 



For three blades, the hovering equations are iden- 
tical, except that the B 2 equation is then absent. 
The Bg and 6 2 equations at hover are completely 
uncoupled and are both identical to that of the 
single blade in rotating coordinates. 



(1), for u = 





B. + X B 4 B, + v 2 B. = 

3. o 1 X 


1-0 


K + 1- ^K * v2 b„ = o 






6 2 + % b^B,, + v 2 B 2 = 0, 



from equation (4) 



Eigenvalues of these equations are easily calcu- 
lated, and are shown on figure 1. These will be 




J3+-^B 4 j3 + I/ 2 /3 = 

Figure 1. Hover eigenvalues of coning and reaction- 
less modes (v ■ 1.2, y - 8). 

called coning and reactionless modes. The reaction- 
less mode is so named because at hover it produces 
no net reaction at the hub. The equations for 
rotor pitching and rolling are, 







J* 




x>* 



i B * 




and the characteristic equation is then, 

(x 2 + £ B"X + v 2 - l) 2 ♦ (2X + } B 1 *) 2 - 

The eigenvalues for this equation are shown in 
figure 2. By analogy to a gyro, these modes will 
be called precession (the lower frequency mode) and 
nutation (the higher frequency mode) . The damping, 
-•y/16, is the same as for the single blade of equa- 
tion (1) and for the &„ and B 2 modes discussed 
above. However, the coordinate transformation has 
resulted in the precession mode frequency being ft 
lower than the single blade mode frequency in the 



47 




Figure 2. Hover eigenvalues of precession and nuta- 
tion modes (v = 1.2, y = 8) . 

rotating frame. Similarly, the nutation mode fre- 
quency is G higher than that of the single blade in 
the rotating frame. 

The coning mode (fig. 1) will excite vertical 
motions of the vehicle, while the precession mode 
will excite pitch and roll motions. Thus vehicle 
responses are more intuitively understood by use 
of the nonrotating coordinates. Also, these equa- 
tions may be used to study feedback control systems 
such as rotor tilting or rotor coning feedback, 
which were discussed in Reference 1. Note (from 
eqs. (1) and (2)) that the performance of such sys- 
tems will depend on the number of blades used, 
since the blade motions become coupled by the feed- 
back terms and the coupling will vary with the 

number of blades. 

i 

To compare the various modes with each other 
and with other theories, it is necessary to trans- 
form all eigenvalues into the same reference frame. 
The obvious choice is the rotating coordinates of 
equation' (1), since most other theories are appli- 
cable to this frame. As may be seen by comparing 
figures 1 and 2, the precession and nutation modes 
may be transformed back into the rotating frame by 
adding and subtracting Q respectively. This 
process results in four identical eigenvalues, as 
expected, since the rotor is composed of four iden- 
tical blades, each described in the rotating frame 
by equation (1). As noted above, the frequencies 
of the 0. and 3, modes do not change 

The equations for u = have been easily 
solved and the nonrotating coordinate system has 
been presented and discussed. In nonhovering 
flight, however, the equations have periodic 
coefficients, which makes the equations more diffi- 
cult to solve, as well as giving the solutions some 
special characteristics. These will be discussed 
in the next section. 



Periodic Coefficient Solutions 

Floquet Theory 

Eigenvalues of equations such as (1) may be 
found with Floquet theory, as for example in 
References 3 and 4. The equation is integrated for 
one period Op = 0,..,, 2ir) for each independent 
initial condition to obtain the state transition 
matrix. The frequency and damping of the system 
modes are then obtained by taking the logarithms of 
the state transition matrix eigenvalues. 

This technique has been applied here to three 
cases, and the results are shown on figure 3 for 



-1.5 




a. v-\.\ , y = 6.0 

b. !/=l.O, y = 6.0 

c. v-\.0, y=l2.0 



a 



.5 



-- -.5 



-1.0 



.2 .3 .4 .5 



x -1.5 



Figure 3. Floquet theory root loci for varying y; 
single blade in rotating coordinates. 

varying u. Note that as u is increased, the 
frequency (w/fi) decreases, while, the damping (a /Si) 
remains constant at -y/16 until the frequency 
reaches an integer multiple of 1/2 /rev. As. u is 
increased further, the frequency remains constant 
while the damping both decreases (the upper roots) 
and increases (the lower roots) as shown for cases 
a and c. This behavior may be surprising to those 
accustomed to constant coefficient equations, but 
is typical of periodic systems. The nonsymmetry 
about the real axis is analogous to the behavior of 
a constant coefficient equation root locus when the 
locus meets the real axis. At that point, the 
roots separate (no longer complex conjugates) , one 
becoming less stable and the other becoming more 
stable. With periodic coefficient equations, the 
separation can occur at any multiple of 1/2 the 
frequency of the periodicity. Actually, the con- 
stant coefficient equation is a special case of the 
periodic One, where the frequency in the coeffi- 
cients is zero. This behavior may be seen in more 
detail by plotting the eigenvalues versus y, as in 
figure 4, which again shows results from Floquet 
theory. 



48 



\:<l 








a 




1.0 


b 






.8 




a. i/ = i.i,y = 6 

b. !/ = l.0,y = 6 


.6 


^^^Nc 


c. i/ = l.o,y-l2 



t 



I 
(a) FREQUENCY 
1.0 




(b) DAMPING 

Figure 4. Floquet theory variation of frequency and 



damping with 
dinates. 



u; single blade in rotating coor- 



Figures 3 and 4 have shown the eigenvalues in 
the rotating coordinate system. These may be 
transposed into the nonrotating system to examine 
the behavior of the nonrotating modes. Choosing 
case c (v = 1.0, y = 12) as an example, the root 
locus is plotted on figure 5. The coning mode has 
the same eigenvalues shown in the two previous 
figures. The nutation and precession modes have 
the same damping, but as mentioned before, their 
frequencies are Q higher and lower, respectively, 
than the coning frequency. 

The regions where the frequency remains constant 
while the damping changes, called critical regions, 
may be illustrated by constructing the y - \i 
plane as in figure 6 (and discussed in References 3 
and 4). In the 0/rev region, the behavior is like 
that of a constant coefficient equation when the 
root locus meets the real axis; there are two real 
roots, with order u 2 changes in damping. In. the 
|/rev region, the frequency is exactly half of the 
rotational frequency (fig. 3, case b), and the 
damping changes somewhat more rapidly. In the 
1/rev region (fig. 3, case a) the frequency is the 
same as the rotational frequency (fi), and again 
the damping changes are order y 2 . As previously 
noted, damping is constant at -y/16 outside of the 
critical regions. Note that varying v has little 
effect on the boundaries of the 0/rev and ■j/rev 
regions, but as v is increased the 1/rev region 
moves upward. 

In this section, the characteristics of the 
periodic coefficient solutions have been discussed 




NUTATION 



-■--2.0 



Figure 5. Floquet theory root loci for varying u; 
three-bladed rotor in nonrotating coordinates 
(v = 1.0, y = 12 j case c). 



24 



20 





'CASE a 



-J_ 



.1 
(a)v = l.l 



.2 .3 



Figure 6. y - v plane for single blade in rotat- 
ing coordinates based on Floquet theory. 



49 




Note that the 6 2 equation is not coupled to the 
others and is the same as the 2 equation for 
hover; hence it yields only the y = roots. 
Therefore the 6 2 equation will not be discussed 
further or included in subsequent figures. The 
B equation has only one u-dependent term, 
coupling it to the 3 1S motion. The pitch and 
roll equations are coupled by both damping and 
aerodynamic spring terms. 

Comparison 

As noted earlier, eigenvalues may be compared 
by adding Q to the precession frequency and sub- 
tracting f2 from the nutation frequency. In 
examining the constant coefficient approximation, 
any differences in eigenvalues will be due to the 
dropped periodicity. That is, all of the roots 
should approximate those obtained by using Floquet 
theory to solve equation (1) . Using the comparison 
method mentioned above, the constant coefficient 
approximation is compared to Floquet theory results 
in figures 7, 8, and 9. The frequency scales have 
been expanded to exaggerate the effects of forward 
speed. Each of the three cases is discussed below. 



Case a. 



1.1, y = 6. (fig. 7) 



Figure 6. Concluded. 

for nonhovering flight. The next sections will 
discuss an approximation which has constant coeffi- 
cients, yet gives some of the behavior of the peri- 
odic coefficient system. 

Constant Coefficient Approximation 

In equation (1) the periodic coefficients are 
all of the speed (y) dependent terms, and a con- 
stant coefficient approximation yields only the 
hover solution. However, in the nonrotating frame 
of equations (2) and (3) , these periodic terms 
have been transformed into constants plus higher 
harmonic periodic terms. This result suggests that 
the primary effects of u may be determined by 
using the average values of the coefficients. The 
constant coefficient approximation thus obtained 
for a four-bladed rotor is given in equation (5) . 
The corresponding equation for three blades is 
identical, except that the B 2 motion is absent. 



^B* 2 



X B 3 



*!» 3 



x(b^,^) 



I (B" - I V W) v* - 1 





(5) 



This case corresponds to a hingeless rotor 
similar to the Lockheed XH-51. For this rotor the 
variations with y of frequency and damping are 
small but significant since the 1/rev critical 
region is encountered (see fig. 6). All three 
modes of the approximation agree well with Floquet 
theory at low advance ratios, where the influence 
of the periodic coefficients is small. As the 
advance ratio is further increased, the precession 
mode displays the same type of behavior as the 
Floquet theory results, but the other two modes do 
not. For the precession mode (and the Floquet 
theory), the frequency becomes constant at 1/rev, 
and the damping then has two values as previously 
discussed. It is useful to examine why the con- 
stant coefficient approximation displays periodic 



1.075 



1.050 



1.000 



O PRECESSION 
D NUTATION 
A CONING 
FLOQUET THEORY 




t 



.1 
(Q) FREQUENCY 



.3 



Figure 7. Comparison of constant coefficient 
approximation to Floquet theory (v = 1.1, y = 6). 



50 



-.5 



-.4 



O PRECESSION 
D NUTATION 
A CONING 
FLOQUET THEORY 



,-Q 



«=*#= 



CASED 
V'U 

y--e 




.1 

(b) DAMPING 



NUTATION 



CASE a 

y-e 

•FLOQUET THEORY 
■ APPROXIMATION 



CONING 



L 



/j. = 0.5 



4 .w 

'a 

2.05 
-- 2.00 



1.05 
1.00 



-- .05 



PRECESSION / 

1 — — «--K 

-.5 -.4\ 



■V 



a 



~1 



CONING 



NUTATION 



.05 



---1.00 
-1.05 

-J- -2.05 



(C) ROOT LOCI 



Case b. v = 1.0, y = 6. (fig. 8) 

This case corresponds to an articulated rotor 
having relatively heavy blades, such as might be 
used for a high speed helicopter. This case is 
well removed from critical regions, and there are 
no significant changes in the eigenvalues for the 
u range shown. The constant coefficient approxi- 
mation agrees well with results from Floquet theory. 



Figure 7. Concluded. 

system (critical region) behavior. In this case, 
the precession roots at hover (u = 0) are very near 
the real axis due to the coordinate transformation. 
As \i is increased, the precession roots move 
toward the real axis and then split when they reach 
the axis, as usual with constant coefficient sys- 
tems. Thus the damping both increases (the left 
branch) and decreases (the right branch) . 



1.0 



.7 



O PRECESSION 
D NUTATION 
A CONING 
FLOQUET THEORY 



CASE b 
I/=I.O 
y = 6 



1 



(Q) FREQUENCY 



-.2 



-.1 



r ° 


PRECESSION 


□ 


NUTATION 


A 


CONING 




FLOQUET THEORY 




Pi -.PI rS 9 - 


¥ 






CASE b 


- 


v=\.0 




y-6 




1 1 1 1 1 



.1 

(b) DAMPING 



Figure 8. Comparison of constant coefficient 
approximation to Floquet theory (v = 1.0, y = 6). 



51 



CASE b 

!/=I.O 

y-6 

-FLOQUET THEORY 



-« APPROXIMATION 

/x= 0.5 



'a 

-- 1.95 
1.90 

.95 
.90 



.10 
-- .05 



4- 



-*« 



-- -.05 
.10 

-.90 
-.95 

-- -1.90 
1.95 



(C ) ROOT LOCI 



Figure 8. Concluded. 

Case c. v = 1.0, y = 12. (fig. 9) 

This case corresponds to a typical articulated 
rotor with blades similar to many aircraft flying 
today. The Floquet theory indicates that the 4/rev 
region is encountered at u = 0.215. It is seen 
that the nutation mode is a poor approximation. 
Apparently, the constant coefficient approximation 
is not adequate for higher frequency modes if a 
critical region is encountered. The precession and 
coning modes (combined), however, do display the 
correct type of behavior: the frequency approaches 
■j/rev and the damping both increases (the precession 
mode) and decreases (the coning mode) . In this 
case, the correct behavior is obtained because two 
modes are involved. As may be seen in figure 9(c), 
the two sets of Floquet roots approach each other, 
meet at ^-/rev, and split (no longer complex con- 
jugates) . This behavior is approximated by the 
coning and precession modes, but in the approxima- 
tion, the roots remain complex conjugates as shown 
in figure 9(c). The frequency of the precession 
mode does not agree well with Floquet results, but 
its damping is increasing; hence it is of less 
interest. The coning mode agrees well with the 
Floquet results, predicting the reduced damping 
very accurately. 

Perturbation Theory 

Equation (1) has also been studied in 
Reference 5, using a perturbation technique known 
as the method of multiple time scales. Analytic 
expressions are derived for the eigenvalues, with 
expressions valid near and within each of the 
critical regions and ones which are valid away from 
the critical regions. These results are very 



.75 



.70 



r ° 


PRECESSION 


D 


NUTATION 


A 


CONING 




FLOQUET THEORY 


- 


CASEC 




f = 1.0 




y = l2 




t 



.1 
(a) FREQUENCY 



3. .3 



.5 



-1.0 



-.8 



a 



-.7 



-.6 



-.5 



O PRECESSION 
□ NUTATION 
A CONING 
FLOQUET THEORY 




(b) DAMPING 

Figure 9. Comparison of constant coefficient 
approximation to Floquet theory (v = 1.0, y = 12), 



52 









i 


'a 










-2.0 




CASE 
l/=I.O 
y = l2 


C 


NUTATION 


- 1.5 








FLOQUET THEORY 






-•— 


— 


APPROXIMATION 






= 0.5 





/coning 


- i.o 








/ 


■ ». 


- .5 






/ I * 








" 1 PRECESSION 




cr 

1 a . 


-1.5 




-1.0 -.5 




.5 






^ 




- -.5 










- — ■*. 










- -i.o'' 








~L 


- -1.5 












--2.0 





(C) ROOT LOCI 

Figure 9. Concluded. 

useful; they give additional insight into the behav- 
ior of periodic systems in general and equation (1) 
in particular. A comparison between the Floquet 
results of the present work and the analytic results 
from Reference 5 indicates that the latter are also 
useful quantitatively. An exception is near the 
i/rev region, where the perturbation solution was 



carried only to order 
extended to order u 2 
perturbation solution. 



y. It should evidently be 
as was the rest of the 



Discussion 

Based on the cases described above, it is 
apparent that the constant coefficient approximation 
may be used to calculate rotor eigenvalues at 
advance ratios up to 0.5. A range of rotor param- 
eters (y and v) have been studied which are repre- 
sentative of most conventional helicopters. The 
lower frequency modes agree well with Floquet 
results and display behavior approximating that of 
the Floquet theory critical regions. Therefore, 
there are many cases where the approximation may be 
used instead of more complicated methods. 

The higher frequency modes of the approxima- 
tion, however, do not display the correct behavior. 
Where these modes are important, for example, in 
using high gain feedback, the approximation should 
be used with caution. 

The perturbation theory of Reference S is very 
easy to use for rotor stability calculations. 



However, the solutions are for uncoupled blades in 
the rotating coordinate frame. To account for 
inter-blade coupling (as with certain feedback 
schemes) one must either use another technique 
such as that described herein or rederive the solu- 
tions with the coupling included. 

Conclusion 

Transforming the flapping equation of a heli- 
copter rotor in forward flight into the nonrotating 
coordinate frame results in a set of differential 
equations where some of the periodicity due to 
forward flight is transformed into constant terms. 
Using the time-averaged values of these, i.e., 
dropping the remaining periodicity, gives a con- 
stant coefficient approximation which retains some 
of the periodic effects. Comparison between results 
of the approximation and those of Floquet theory 
indicates that the approximation should be accept- 
ably accurate for calculating flapping stability of 
most helicopters for the advance ratios shown 
herein. Use of the nonrotating coordinates has 
given insight into rotor behavior and indicates how 
the vehicle motion would be affected by the rotor 
modes . 

The higher frequency modes of the approximation 
do not agree well with Floquet theory. Where these 
modes are important for example, in using high gain 
feedback control systems, the approximation should 
be used with caution. 

References 

1. Hohenemser, K. H. and Yin, S-K., SOME APPLICA- 
TIONS OF THE METHOD OF MULTIBLADE COORDINATES, 
Journal of the American Helicopter Society , 
Vol. 17, No. 3, July 1972, pp 3-12. 

2. NASA CR- 114290, RESEARCH PROGRAM TO DETERMINE 
ROTOR RESPONSE CHARACTERISTICS AT HIGH ADVANCE 
RATIOS, Kuczynski, W. A. and Sissingh, G. J., 
February 1971. 

3. Peters, D. A. and Hohenemser, K. H., APPLICATION 
OF THE FLOQUET TRANSITION MATRIX TO PROBLEMS OF 
LIFTING ROTOR STABILITY, Journal of the American 
Helicopter Society , Vol. 16, No. 2, April 1971, 
pp 25-33. 

4. Hall, W. Earl Jr., APPLICATION OF FLOQUET THEORY 
TO THE ANALYSIS OF ROTARY-WING VTOL STABILITY, 
SUDAAR No. 400, Stanford University, February 
1970. 

5. NASA TM X-62,165, A PERTURBATION SOLUTION OF 
ROTOR FLAPPING STABILITY, Johnson, W., July 1972. 



53 



FLAP-LAG DYNAMICS OF HINGELESS HELICOPTER BLADES AT MODERATE AND HIGH ADVANCE RATIOS 



P. Frledmann 
Assistant Professor 

and 

L.J. Silverthorn 

Research Assistant 

Mechanics and Structures Department 

School of Engineering and Applied Science 

University of California, Los Angeles 



Abstract 



Equations for large amplitude coupled flap- 
lag motion of a hingeless elastic helicopter blade 
in forward flight are derived. Only a torsionally 
rigid blade exicted by quasi-steady aerodynamic 
loads is considered. The effects of reversed flow- 
together with some new terms due to forward flight 
are included. Using Galerkin's method the spatial 
dependence is eliminated and the equations are 
linearized about a suitable equilibrium position. 

The resulting system of equations is solved 
using multivariable Floquet-Liapunov theory, and 
the transition matrix at the end of the period is 
evaluated by two separate methods. Results 
illustrating the effects of forward flight and 
various important blade parameters on the stability 
boundaries are presented. 

Notation 



a 

A 
k 

^i'^i 

^Fi'*Li 
b 



Two dimensional lift curve slope 
Tip loss coefficient 



Periodic matrix with elements A. 
defined in Appendix B 



*J' 



C T 
C 

c 



do 



Generalized aerodynamic force for 
i tn flap and lag mode respectively 

Same as above, in reverse and 
mixed flow regions. 

Semi-chord nondimensionalized with 
respect to R 

Tip loss coefficient 

Generalized masses defined in 
Appendix A 

Thrust coefficient 

Constant matrix 

Profile drag coefficient 



Presented at the AHS/NASA-Ames Specialists' 
Meeting on Rotorcraft Dynamics, February 13-15, 
1974. 

_ 
Presently, Dynamics Engineer, Hughes Helicopter 
Company, Culver City, California. 



C(k) 



FEE 6 

^Gi'^cr B ik' 

Is _cs „cs 



im' 

(EI) 
3 



A *k 
8 SF ,S SL 



\ 

A \ 
i- yCT 

h 
i 



L ,L 

y z 



im' ik 



Theodorsen's lift deficiency 
function 

Defined in Fig. 1 

Terms associated with elastic 
coupling defined in Appendix A 



Stiffness for flapwise bending 

Stiffness for inplane of rotation 
bending 

Flap coefficients defined in 
Appendix A 



Generalized coordinate, k 
normal flapping mode 



th 



Static value of 



Perturbation in 



\ 



in hovet 



about 



\ 



Viscous structural damping in 
flap and lag respectively 

Generalized coordinate, m 
normal inplane mode 

Static value of h, in hover 

, o 



Perturbation in 



\ 



about 



\ 



Unit vectors in x,y and z direc- 
tions (Fig. 1) 

Mass moment of inertia in flap, 
defined in Appendix A 

Unit matrix 

Length of blade capable of 
elastic deflection 

Aerodynamic load per unit length 
in the y and z directions 
respectively 

Lag coefficients, defined in 
Appendix A 

Mass of blade per unit length 



55 



M,N 
^i>\i 



( Vikr ( Vimr 



P »P »P 
r x r y r z 



ikm 
P(t) 

R 

R 

Q 
T 
u,v,w 



V ,v 
e eo 



w ,w 
e eo 



x,y,z 



V Y G 

a 



Number of modes in lag and flap 
respectively 

Generalized mass for the i*-" 
flap and lag mode respectively, 
defined in Appendix A 

Defined in Appendix A 

Resultant total loading per unit 
length in the x,y and z direc- 
tion respectively 

Defined in Appendix A 

Periodic matrix 

Blade radius 

Constant matrix used in Floquet- 
Liapunov theorem 

Constant matrix 

Common nondimensional period 

x,y and z displacement of a point 
on the elastic axis of the blade 

Component of air velocity w.r.t. 
the blade at station x perpendic- 
ular to x-y plane (hub plane) , 
positive down 

Same as above, in the x-y plane, 
tangent to a circle having a 
radius x 

Elastic part of the displacement 
of a point on the elastic axis of 
the blade parallel to hub plane, 
(see Fig. 1), subscript o de- 
notes the static equilibrium 
value 

Velocity of forward flight of the 
whole rotor 

Elastic part of the displacement 
of a point on the elastic axis of 
the blade, in the k direction, 
approximately, (Fig. 1) 

Rotating orthogonal coordinate 
system 

Running spanwise coordinate for 
part of the blade free to 
deflect elastically 

Defined in Appendix B 

Angle of reversed flow region 
(Fig. 2) 

Angle of attack of the whole 
rotor 



% 



T1 SF 1 ' n SL 1 






m,%) 



Droop, built in angle of the 
undeformed position of the blade 
measured from the feathering axis 
(Fig. 1) 

Preconing, inclination of the 
feathering axis w.r.t. the hub 
plane measured in a vertical 
plane 

Lock number (y=2p bR a/I. ) for 
normal flow 

m tn inplane bending mode 

Symbolic quantity having the same 
order of magnitude like the dis- 
placements v and w 

Real part of the k character- 
istic exponent 

k flapwise bending mode 

Viscous structural damping coef- 
ficients defined in Appendix A 

Collective pitch angle measured 
from x-y plane 

Critical value of collective 
pitch at which the linearized 
coupled flap-lag system becomes 
unstable in hover 

Inflow ratio, induced velocity 
over disk, positive down, non- 
dimensionalized w.r.t. Rfl 

Diagonal matrix, containing 
eigenvalues Aj^ of R 

Diagonal matrix containing eigen- 
values Aj^ of J,(T,0) 

Advance ratio 

Critical value of advance ratio 
at which flap-lag system becomes 
unstable 

Density of air 

Blade solidity ratio 

State transition matrix at ty , 
for initial conditions given at 

o 

Azimuth angle of blade #=fit) 
measured from straight aft 
position 



■"C 



Flutter frequency 



ith 



Imaginary part of k character- 
istic exponent 



56 



%l'\l 



Natural frequency of l" 1 flap or lag 
mode , rotating 

Speed of rotation 



Special Symbols 



( ) 



( )' 
(*) 



(.) 



() 



Nondimensionalized quantity, length 
for elastic properties nondimensional- 
ized w.r.t. A; all other w.r.t. R 
frequencies w.r.t. £2; mass properties 
w.r.t. ^ 

Differentiation w.r.t. x 

o 

Differentiation w.r.t. \|) 

Subscripts, denoting real and imagin- 
ary parts of the appropriate quantity 

The symbol beneath a quantity 
denotes a vector or a matrix 

Denotes the inverse of a matrix 



The dynamics of a helicopter blade in forward 
flight are usually described by a system of linear 
differential equations with periodic coefficients. 
A growing acceptance of hingeless helicopter 
blades for conventional helicopters flying at 
relatively high forward flight speeds has intensi- 
fied the need for fundamental research on the 
aeroelastic stability of such systems. 

Studies dealing with the effect of forward 
flight (or periodic coefficients) have been 
primarily devoted to the study of flapping insta- 
bility at high advance ratios. 1 " 8 A limited 
number of studies dealing with the effect of 
periodic coefficients on coupled flap-lag ' or 
coupled flap-lag-pitch 11 motion were also con- 
ducted. The case of coupled flap-lag motion has 
been, somewhat inconclusively, investigated by 
Hall using multivariable Floquet theory, the 
same problem was also considered by Friedmann and 
Tong 9 but the treatment was limited to low advance 
ratios (y<0.3). The coupled, linearized, flap- 
lag-torsion motion has been investigated by 
Crimi 11 using a modified Hill method. In both 
cases 10 ' 11 only a limited number of numerical 
results were obtained and the physical mechanism 
of the aeroelastic instabilities has not been 
clearly identified, in particular the degree of 
freedom which triggers the instability was not 
identified and the results for forward flight were 
not compared with those for hover. 

Recent investigation of the aeroelastic sta- 
bility of hingeless blades in, hover 12 indicated 
that the aeroelastic stability boundaries are 
quite sensitive to the number of degrees of free- 
dom employed in the analysis. Therefore it is 
important to determine how the flapping behavior 
of a blade at high advance ratios is modified by 
the lag degree of freedom. This important problem, 
which has not received adequate treatment before, 
is one of the main topics of the present study. 



The mathematical methods used in previous 
studies dealing with the effects of forward flight 
were: (a) The rectangular ripple method 1 , (b) Ana- 
log computer simulation, (c) Various forms of 
Hill's method, 2 ' 11 (d) Multivariable Floquet- 
Liapunov theory, 6 ' 7 ' 1 (e) Perturbation method in 
multiple time scales. 8 ' 9 The mathematical method 
employed in the present study is the Floquet- 
Liapunov theorem, and the transition matrix is 
evaluated by two separate methods. It is also 
shown that careful use of this method enables one 
to circumvent problems associated with identifying 
the results encountered in previous studies. 



10 



In addition, a new and convenient approxima- 
tion for the reversed flow region is developed, 
this approximation is believed to be adequate for 
most blade stability analyses. Finally, the effects 
of various important parameters such as collective 
pitch setting, structural damping, droop and pre- 
coning on the instability associated with forward 
flight is investigated. 

1. The Equations of Motion 

1.1 Basic Assumptions 

The present study is based upon a consis- 
tently derived system of equations of motion for 
the linearized coupled flap-lag motion of a 
cantilevered rotor blade at arbitrary advance 
ratios . 

The derivation itself is algebraically 
tedious, thus only a brief outline will be given 
in this paper, the complete details of the deriva- 
tion can be found elsewhere. 

The geometry of the problem is shown in Fig. 
1. The following basic assumptions were used in 
deriving the equations of motion: (a) The blade is 
cantilevered at the hub. It can have an angle of 
droop Bj) at the root. In addition, the feather- 
ing axis can be preconed by an angle 3p. The 
angles 3n and 0p are small, (b) The blade can 
bend in two directions normal to the elastic axis 
and is torsionally rigid, (c) The deflections of 
the blade are moderately small so that terms of 
0(ej}) can be neglected compared to one. (d) 
Moderately large deflections have only a small 
effect on the tension due to elastic effects on 
the blade since one of its ends is free, thus a 
linear treatment of the elastic restoring forces 
is adequate, (e) Two dimensional quasi-steady 
aerodynamic strip theory is used C(k)=l and 
apparent mass effects are neglected, (f) Reversed 
flow is included using an approximate model for 
reversed flow described in Appendix C. (g) Stall 
and compressibility effects are neglected. 

Using the assumptions given above a system 
of nonlinear partial differential equations for 
the coupled flap-lag motion of the blade is 
derived, with respect to an x,y, and z coordinate 
system rotating with the blade. The derivation 
follows essentially along the lines of Reference 
14, all the details can be found in Reference 13. 



57 



a 

i 

3h 



1.2 Brief Derivation of the Equations of Motion 

The differential equation for the dynamic 
stability of a cantilevered rotor blade can be 
written as ' 



3x' 
o 

3x 2 



( 3 2 w 



a 2 
3 v 



+ E 



C2 3x 2 
o 



3x u 3x J z 
o o 



[(HI) z - E cl ] 



3x' 



3x 2 j 
o / 



^L-L-r 



3 v 3 w^ , „ „ 

-/ +E C2 7^-3x-tix-^y 
o o ' 

(1) 



where the quantities E(^, E(j2 are given in 
Appendix A. 

The distributed loading terms in the x,y and 
z directions with terms up to 0(ei) in displace- 
ments can be written as 



p = _?_ = nin 2 [(x + e n ) + 2*] 

x x o 1 ' 

9 isis it 

P = L - mft [v - (e + v) + 2u] 

y y o 

9 ieit it 

P z = L z ~ m Q W " ^F 12 W 



8 SL n * 



(2) 



The boundary conditions for this kind of blade are 
well known. The displacement field of the blade 
with sin6s9 and cos8 = 1 can be written as 13 ' 11 * 

\Z] 



-wv-W-iCI®'*®] 



v=v - x 8_9 
e o D 

w-w e + x (B p + 3 D ) 



Hx x 
(3) 



and 



» B - E* vx^ct) - * w 

k=l 
M ' 

v — yi& Y (x )h (t) = -X, y h 
e *—! 'm v o' m ' 'm n 



(4) 



m=l 



where it is understood that repeated indices imply 
summation unless otherwise stated. 



by 



The aerodynamic loads L» and L^ are given 



L z " a P A tR V U T 9 ~ V c 
L y - ap A bR|u p (U T e - U p )+ -f- 



4 



where the velocities D_ and U_ are given by 

Up - fiw + ffitfx + u cos^) l^-j 

o ' 



* /- 3v \ 

U„ - S2v + fiRlx+p sinifi + u cos^) -^-j 



(5) 
(6) 

(7) 
(8) 



The last term in Equation (8) is due to the 
radial flow along the blade. This term has been 
neglected in some previous analyses. For arbi- 
trary advance ratios this is an important and non- 
negligible term. 



Combination of Equations (1) through (8) and 
application of Galerkin's method to eliminate the 
spatial variable reduces the problem to a system 
of ordinary differential equations. 

** — _ * _ s —2 cs 
*V±H+ 2 W Fi M Fi T1 SFi S i +E ik 8 k + "FAi 8 ! = E im h m 

+2 WkV ( VV 5 ^ 2 <VV 5 L h m + h± (9) 

— rfesfe _ _ 4> — —2 =5 =cs 

\l\ +2 \±\±\L± h ± + \i\± h ±- E im h m " E iA 

-2 B? k ( VVlX 1 ^D 9 * 2 [ § imr-< fi Y >imr ] \ h r 

" 2 ^mM V* 2 B LA> e \ + ^i (10 > 

where the various quantities Mp^, Pi^, Mli, S± m: , 

0*r\)lkSL are generalized mass integrals given in 

References 13 and 14, and also in Reference 12 

-1 -3 
(for i=k=£=m=r=l) . While the quantities B^, B im , 

E ik> 5 im. Km' E ik> Si 1 , I7 k , Bf amd If m etc. 
are given in Appendix A. The quantities Ap^, A L i 
are generalized aerodynamic forces defined by 



* 



(ID 



(12) 



Equations (9) and (10) are coupled nonlinear 
ordinary differential equations. In the present 
study these equations will be linearized about a 
suitable equilibrium position, which is taken to 
be the steady state equilibrium position of the 
blade in hover . Through this process of lineari- 
zation many nonlinear terms are transformed into 
coupling terms. At this stage one encounters a 
considerable number of terms which are small and 
therefore negligible. In order to neglect the 
appropriate terms a rational ordering scheme is 
used which enables one to neglect terms in a 
systematic manner. In this scheme all the impor- 
tant parameters of the problem are assigned orders 
of magnitudes in terms of a typical displacement 
quantity Ej) thus: 



v 
R 



0(e D );| 



0(e D ); x=0(l); y-O(l); 
X=0(e D ); 9=0 (e D ) 



fr=0( eD );|^=0( eD );3 D -e p -0( eD ); 
o o 



do 
a 



»0(e 2 ) 



(13) 



An order of magnitude analysis of the equa- 
tions indicates that in general terms up to and 
including O(ep) must be included in the linearized 
flap equations, while for lag equations some 0(e-h 
terms have to be retained. 

The process of the linearization consists of 
expressing the elastic part of the displacement 
field as 



58 



v = v + A v =-Y(h° + Ah) (W) 
e eo e 'mm m 

where the static equilibrium condition in hover 
is given by 

s iA +52 Fi M FA - ^X=-(^v 5 i-i<|) 2 ( 9F i- XF i> 



where i = 1,2,. 



11 IRk 



Q 

+ J (|) 2 [X(8lJ- XL 2 ) + -|°- I*] 1-1,2,... H (15) 

The various quantities F , L are defined in 
Appendix A. Next, for the sake of simplicity, 
the equations are specialized to the case of one 

elastic mode for each degree of freedom, i.e. one 
flapping and one lead-lag mode. 

Furthermore for mathematical convenience the 
equations of motion have to be transformed into a 
system of first order equations. This is achieved 
by using the following notation 



Ag l = *1 



Ah 1 -y A 



(16) 



For the stability analysis, only the homo- 
geneous part of the equations of motion is 
required, thus the equations of motion in their 
final form can be written as 

£ = A i*)y < 17 > 

where A is a 4x4 matrix defined in Appendix B. 

The equations of motion (17) will have a dif- 
ferent form for the normal flow region and for the 
reversed flow region. The representation of the 
reversed flow together with its effect on the form 
of Equations (17) is described in Appendix C. 

2. Method of Solution 

The stability investigation of the blade 
motions is based upon the Floquet-Liapunov 
theorem 15 which states the knowledge of the state 
transition matrix over one period is sufficient in 
order to determine the stability of a periodic 
system having a common period T. Based upon the 
Floquet-Liapunov theorem, the transition matrix 
for the periodic system can be written as 15 

SOMO 



*(*.*„) = £ 



OJOe 



£<*„> 



where 



PflrtT) = P(i|>) 



(18) 



(19) 



where R is a constant matrix and £(t) is a periodic 
matrix. Clearly the stability of the system is 
determined by the matrix R, where R is given by 
following relation 



*(T,0) 



*&-£. 



(20) 



A direct result of the Floquet-Liapunov theorem is 
that the knowledge of the transition matrix over 
one period determines the solution to the homo- 
geneous system everywhere through the relation 



RT V 



&0P+ST.0) = £0|»,-0)(e~ ) 
where < i|i <_ T, s any integer. 



(21) 



In general R is a fully populated (nxn) square 
matrix. If it has n independent eigenvalues, it 
is well known from elementary linear algebra 5 
that a similarity transformation can be found such 
that 



OR Q = A 



(22) 



where the columns of Q are the n-linearly inde- 
pendent eigenvectors of R and A, is a diagonal 
matrix whose elements are the eigenvalues of g,. 
Combining Equations (20) and (22) and using the 
definition of the matrix exponential 15 one has 



RT 



Q ^ Q -l 



J? . A = g~ l C £ 



jf^cr.ooij 



(23) 



where A, is a diagonal matrix containing the eigen- 
values of the transition matrix at the end of one 
period. The eigenvalues of £(T,0) or the char- 
acteristic multipliers are related to the eigen- 
values of R, denoted characteristic exponents, 
through the relation 

V 

e = A^ k=l,2,...r (24) 

Clearly Aj. and A^ are both complex quanti- 
ties in general, thus 



\ = \ + iW k 
\ = A kR +lA kI 



from which 



and 



<* - 2? fa[i 4R + 4* 



\ = T^ A^ 



(25) 



(26) 



(27) 



the quantity to^ can be determined according to the 
Floquet-Liapunov theory only within an integer 
multiple of the nondimensional period. 

The stability criteria for the system is 
related to the eigenvalues of g. or the real part 
of the characteristic exponents f^. The solu- 
tions of the Equation (17) approach zero as iJj •*■ °° 
if 



+ AJ^I < 1 or Zy.< k=l,2,.. 



. ,n 



59 



Finally a brief description of the numerical 
implementation of the scheme described above will 
be given. The transition matrix at the end of 
one period $£T,0) is evaluated using direct numer- 
ical integration. Equations (17) are integrated 
for the set of initial conditions corresponding 
to $£0,0) = £. The numerical integration is per- 
formed using a fourth order Runge Kutta method. 
The eigenvalues of the transition matrix are 
evaluated by a Jacobi type eigenvalue routine. 
For some of the cases the value of 4>XT,0) has 
been evaluated using Hsu's method. 13 ' 17 This was 
done in order to obtain results by two different 
numerical schemes and also because Hsu's method 
was found to be more efficient numerically. Both 
methods yield identical results, therefore it is 
not specified on the plots which scheme was used 
to evaluate $0,0). 

3. Results and Discussion 

3.1 Humerical Quantities Used in the Calculations 

In computing the numerical results the fol- 
lowing assumptions were made, 

Mass and stiffness distribution was assumed 
to be constant along the span of the blade. Two 
different kinds of mode, shapes were used: 

(a) For most of the cases for which essen- 
tially trend type studies were conducted an 
assumed mode shape in flap and lap was used as 
given by the appropriate expression in Reference 
12. When an assumed mode shape is used the elastic 
coupling effect 16 is neglected. 

(b) For a few cases an exact rotating mode 
shape in flap and lag was employed. These mode 
shapes were generated by using Galerkin's method 
based upon five nonrotating cantilever mode shapes 
for each flap or lag degree of freedom. For these 
cases the elastic coupling effect was included. 

The coefficients F x , L and B defined in 
Appendix A and in References 12 through 14 were 
evaluated using seven point Gaussian integration. 
For the region of reversed flow these coefficients 
were treated in a special manner as explained in 
Appendix C. 

For the cases computed the inflow was evalu- 
ated using an expression for constant inflow ratio 
in hover , given by 



, aO 
A = 16 



1 + 



248 
acr 



- 1 



(28) 



This inflow relation is equivalent to taking 
the induced velocity of 3/4 blade radius as repre- 
sentative of a constant induced velocity over the 
whole disk. It is clear that for forward flight 
one should use the expression 



X = u tano^ + C T /2 Yu 2 +X 2 



(29) 



Use of this relation would have required the 
use of a trim procedure in the calculations. It 



can be seen from Reference 14 that the require- 
ment of trimmed flight at a fixed C T results in 
an increase of 8 at advance ratios of p > .3 and 
it also requires continuous changes in B-± c and 
e^g. The experience gained when using this 
approach in Reference 14 indicates that when the 
trim requirement is included in the calculation, 
the value of y c at which instability will occur 
will be usually lower. Furthermore, when using 
this approach it was found that it is difficult to 
determine which part of the degradation in stabil- 
ity is related to the increase in 8, 8 ls and 8i c 
and which part is due to the periodic coefficients. 
This added complication is not warranted in a 
trend study such as the present one, and it is 
not consistent with the stated purpose of this 
paper, which is; a clear illustration of the 
effects of the periodic coefficients when the lag 
degree of freedom is included in the formulation 
of the problem. 

Finally, in all the computations the follow- 
ing values were used: 

C do = .01; a=2ir; 0= .05; e^O; A=0. ; I"=l 

Various other pertinent quantities are specified 
on the plots. 

3.2 Results 



The results obtained in the present study 
usually are given in form of plots representing 
the variation of the real part of the character- 
istic exponent £fc with the advance ratio y. Most 
of the cases presented in this study were evalu- 
ated using an assumed mode shape, as described in 
the previous section, and neglecting the elastic 
coupling effect. 

For some cases an exact rotating mode shape 
in flap and lead-lag was used and the elastic 
coupling was included, when this approach was used 
a statement to this effect appears on the appro- 
priate plots. When no such statement appears it 
is to be understood that the assumed mode shape 
is used and the elastic coupling is neglected. 

A typical case is shown in Figure 3 for a 
collective pitch setting of 8= .15. Starting the 
computation at y=0, enables one to easily identify 
the instabilities encountered, by using results 
previously obtained for hover. As shown the lag 
degree becomes unstable and the frequency of the 
oscillation is <% = 1.28119. This result clearly 
indicates that by neglecting the lag degree of 
freedom one could obtain completely incorrect 
stability boundaries. 

The importance of the reversed flow region is 
illustrated by Figure 4. As shown with the 
reversed flow region the instability occurs at 
higher values of y than without the reversed flow 
region. Similar trends were observed in previous 
studies when only the flapping motion was con- 
sidered, 5 indicating that by neglecting the 
reversed flow region one could expect conservative 
results from a stability point of view. It also 



*8 7 .Sj cyclic pitch changes. 



60 



indicates that in this particular case the 
reversed flow region starts being important above 
advance ratios of p = 0.8. 

It is important to note that the frequency at 
which the lag degree of freedom becomes unstable 
is not 1/2 or 1 as is usual for the case of para- 
metric excitation. Thus it seemed important to 
identify the source of this destabilizing effect. 
The results of this study are presented in Figures 
5 and 6. The effect of neglecting the radial flow 
terms on the real part of the characteristic 
exponent, associated with the flap degree of free- 
dom, is shown in Figure 5. As shown, the radial 
flow terms have a stabilizing effect on the flap- 
ping motion with the radial flow forms suppressed 
the flap degree of freedom becomes unstable at 
p=1.33. The effect of the radial flow terms on 
the lead-lag degree of freedom is illustrated by 
Figure 6, as shown without the radial flow terms 
the instability in the lag degree of freedom is 
completely eliminated and the system becomes 
unstable in flap. When the radial flow terms are 
included, the lag degree of freedom is the crit- 
ical one and the system becomes unstable at u= . 754 . 
This matter was pursued further by identifying the 
actual destabilizing term in the equations of 
motion, which was found to be an aerodynamic 
coupling term. This term couples the flap motion 
with the lag motion in the flap equation, its form 
is 

2 2, 9w 8u 

v cos * Si" ST 

o o 

This term is due to the U T Up term in Equation (5) . 
The term shown above is the complete nonlinear 
one, clearly the one retained in the equations of 
motion is the coupling term obtained from linear- 
izing this expression. 

As mentioned in the previous section the 
results presented in Figures 3 through 6 were 
obtained by neglecting the elastic coupling effect. 
In order to asses the effect of this assumption 
the typical case has been also recomputed with the 
exact mode shape and the elastic coupling effect, 
the results are shown in Figure 11. From Figure 
11 it is clear that use of the exact rotating mode 
in flap and lag reduces the value of p c to 
p c = 0.653, when the elastic coupling is also 
included p c is further reduced to p c = .583. Thus, 
for this case, p c seems to be more sensitive to 
the type of mode shape used than to the inclusion 
of the elastic coupling effect. It is also 
interesting to note, that for this case the elastic 
coupling effect is destabilizing, while for hover 
its effect on 9 C is quite stabilizing. 

Previous studies 12 dealing with the effect of 
viscous type of structural damping on the stabil- 
ity boundaries for hover indicated that this para- 
meter has an important stabilizing. The effect 
of this parameter for forward flight is shown by 
Figures 7 and 8. The stabilizing effect of 
structural damping in the lag degree of freedom 
is evident from Figure 7, where the real part of 
the characteristic exponent associated with the 



lead-lag degree of freedom is plotted as a func- 
tion of the advance ratio p , again only the struc- 
tural damping in lag is important. A summary of 
these results is presented in Figure 8 showing the 
variation of p c as a function of the structural 
damping. It is interesting to note that this plot 
indicates that the greatest stabilizing effect 
due to structural damping is obtained in the 
range < risLl < -^ 2 &% o£ critical damping), 
after which, the gain in stability tends to level 
off. Similar trends were obtained from stability 
studies in hover. 12 

Again in order to illustrate the sensitivity 
of the results to the mode shape and elastic 
coupling, the results have been recomputed with 
these effects included; these results are also 

shown in Figure 8. As seen the use of the correct 
mode shape and the elastic coupling effect reduce 
the value of p c , at which instability occurs. 

The sensitivity of the results, to different 
collective pitch settings is illustrated by 
Figure 9. Comparison of Figures 3 and 9 indicates 
that by decreasing the collective pitch setting 
from 9 = .15 to 8 »' >05 eliminated, the instability 
associated with thelead-lag motion. The instabil- 
ity in this case occurs at p c • 1.88 with a 
frequency of or 1. This is a typical flapping 
instability due to the periodic coefficients. 
Comparison of Figures 3 and 9 seems to indicate 
that the assumption of nonlifting rotors used in 
some forward flight studies 7 can be 
nonconservative . 

Finally, Figure 10 shows the dependence of 
Pc on the angle of droop 6n- As shown p c is 
relatively insensitive to B D . On the other hand 
3j) has a very important effect on the value of 8 C 
at which the linearized system in hover becomes 
unstable. 

It should be also noted that a considerable 
number of additional numerical results, including 
the effects of elastic coupling can be found in 
Reference 13. 

4 . Conclusions 



The major conclusions obtained from the pre- 
sent study are summarized below. They should be 
considered indicative of trends and their appli- 
cation to the design of a helicopter blade should 
be limited by the assumptions used. 

(1) Flapping instability and response 
studies at high advance ratios can be inaccurate 
and misleading due to the neglection of the lag 
degree of freedom. The effect of the periodic 
coefficients on the coupled flap-lag system shows 
that in general instability can occur at lower 
values of advance ratios than when the flap 
degree of freedom is considered by itself. 

(2) In addition to the instabilities associ- 
ated with the periodic coefficients (i.e. with 
frequencies of 0, 1 or 1/2) the coupled flap-lag 



61 



system has the tendency to become unstable due to 
an aerodynamic coupling effect associated with the 
radial flow terms. This instability which has a 
frequency close to the rotating lag frequency of 
the system, occurs usually at values of p c much 
lower than those for which the flapping degree of 
freedom becomes unstable. 

(3) Viscous type of structural damping in the 
lead- lag degree of freedom has a stabilizing 
effect on the instability discussed in previous 
conclusion. 

(4) The value of the collective pitch setting 
has a considerable effect on the value of the 
advance ratio at which instabilities due to the 
periodic coefficients or the radial flow 
aerodynamic coupling terms occur. Increase in 
collective pitch is destabilizing, therefore high 
advance ratio studies which do not include this 
effect (nonlifting rotors) may yield nonconserv- 
ative results. 

(5) The numerical results obtained in the 
present study agree with the analytical results 
obtained previously 9 indicating that hingeless 
blades with a rotating lag stiffness of 1/2 or 1 
can easily become unstable due to the effect of 
periodic coefficients. 

(6) While droop has a very strong effect on 
the stability boundaries of hingeless blades in 
hover, it has a very minor effect on the stability 
boundary in forward flight. 

References 



2. 



3. 



5. 



Horvay, G. and Yuan, S.W., STATILITY OF ROTOR 
BLADE FLAPPING MOTION WHEN THE HINGES ARE 
TILTED. GENERALIZATION OF THE 'RECTANGULAR 
RIPPLE' METHOD OF SOLUTION, Journal of the 
Aeronautical Sciences , October 1947, pp. 583- 
593. 

Shulman, Y. , STABILITY OF A FLEXIBLE HELI- 
COPTER ROTOR BLADE IN FORWARD FLIGHT, Journal 
of the Aeronautical Sciences , July 1956, 
pp. 663-670, 693. 

Sissingh, G.J., DYNAMICS OF ROTOR OPERATING 
AT HIGH ADVANCE RATIOS, Journal of American 
Helicopter Society , July 1968, pp. 56-63. 

Sissingh, G.J., and Kuczynski, W.A., INVESTI- 
GATIONS ON THE EFFECT OF BLADE TORSION ON THE 
DYNAMICS OF THE FLAPPING MOTION, Journal of 
the American Helicopter Society , April 1970, 
pp. 2-9. 

R & M No. 3544, THE STABILITY OF ROTOR BLADE 
FLAPPING MOTION AT HIGH TIP SPEED RATIOS , 
Lowis, O.J., 1968. 

Peters, D.A. , and Hohenemser, K.H. , APPLICA- 
TION OF THE FLOQUET TRANSITION MATRIX TO 
PROBLEMS OF LIFTING ROTOR STABILITY, Journal 
of the American Helicopter Society, April 
1971, pp. 25-33. 



8. 



9. 



10. 



11. 



12. 



13. 



Hohenemser, K.H., and Yin, S.K., SOME APPLI- 
CATIONS OF THE METHOD OF MULTIBLADE COORDI- 
NATES, Journal of American Helicopter Society , 
July 1972, pp. 3-12. 

Johnson, W. , A PERTURBATION SOLUTION OF ROTOR 
FLAPPING STABILITY, AIAA Paper 72-955. 

Friedmann, P. and Tong, P., NONLINEAR FLAP- 
LAG DYNAMICS OF HINGELESS HELICOPTER BLADES 
IN HOVER AND IN FORWARD FLIGHT, Journal of 
Sound and Vibration , September 1973. 

SUDAAR No. 400, APPLICATION OF FLOQUET THEORY 
TO THE ANALYSIS OF ROTARY WING VTOL STABIL- 
ITY, HALL, W.E., Stanford University, 
February 1970. 

NASA CR-1332, A METHOD FOR ANALYZING THE 
AEROELASTIC STABILITY OF A HELICOPTER ROTOR 
IN FORWARD FLIGHT, Crimi, P., August 1969. 

Friedmann, P., AEROELASTIC INSTABILITIES OF 
HINGELESS HELICOPTER BLADES, AIAA Paper 73- 
193 January 1973, (also Journal of Aircraft, 
October 1973). 

UCLA School of Engineering and Applied 
Science Report, AEROELASTIC STABILITY OF 
COUPLED FLAP-LAG MOTION OF HELICOPTER BLADES 
AT ARBITRARY ADVANCE RATIOS, Friedmann, P., 
and Silverthorn, J.L., to be published 
January 1974. 

NASA-CR-114 485, DYNAMIC NONLINEAR ELASTIC 
STABILITY OF HELICOPTER ROTOR BLADES IN 
HOVER AND FORWARD FLIGHT, Friedmann, P. , 
and Tong, P., May 1972. 

Brockett, R.W. , FINITE DIMENSIONAL LINEAR 
SYSTEMS, John Wiley and Sons, 1970. 

Ormiston, R.A. , and Hodges, D.H. , LINEAR 
FLAP-LAG DYNAMICS OF HINGELESS HELICOPTER 
ROTOR BLADES IN HOVER, Journal of the 
American Helicopter Society , Vol. 17, No. 2, 
April 1972, pp. 2-14. 

Hsu, C.S., and Cheng, W.H. , APPLICATION OF 
THE THEORY OF IMPULSIVE PARAMETRIC EXCITA- 
TION AND NEW TREATMENTS OF GENERAL PARAMETRIC 
EXCITATION PROBLEMS, Journal of Applied 
Mechanics , March 1973, pp. 78-86. 



Appendix A. Definitions of the Generalized 
Masses, Aerodynamic Integtals and other Quantities 

The quantities, ^11,Mf1,M^ 1 ,'sX 11 , (M^) lu , 

(My) in are generalized masses given, in Appendix 
A of Reference 12, with the general i,m,k indices 
these quantities can be found in References 13 
and 14. 



14. 



15. 



16. 



17. 



62 



5 L =)i3 X Ti i[^ mY » d5 i] d V i b 
i x ° 

o x 

11 x ° 

3£r ** AU »<*i +i i> di ] d Vv 

o 1 



x dx 

o o 



Structural damping is represented by 

g SF = 2 n 5 Fi n SFi ; ^'28^^ 

The elastic coupling effect is represented by 
E C1 = [(EI) z -(EI) y ]sin 2 e; E C2 = [ (EI) a -(EI) 1 



1 




sin6cos9 


j(\i« d *o 


"i^ 1 




i 

/ E C2« d *o 
*^o 


" ^ V^* 




1 

/ Yi' Y" E_-dx 
y 'i 'n CI o 




' 5m V 1 ** 




1 
/« E C2 d *o 


- ^ V& 





When using these expressions in a one mode analysis 
for each degree of freedom the lower indices are 
deleted for these expressions and the expressions 
for the generalized aerodynamic integrals. The 
generalized aerodynamic integrals F*, iA can be 
found in References 12,13 and 14. For this study 
some additional expressions had to be defined, only 
these are given below. 



■J 1 

im 



F 23. 
ikm 



C 20 
im 



/ x ruY'dx ; F 22 - / r].y' d: 
I o I'm o im I I'm 

A A 



A 



A 
B 



Y'dx ; L' 
'm o 



21 , 

'ikm 



/vM^o 



t 22 - 
ikm 



L 24 = 



fa 



,Y'dx ; L 23 
k'm o ' im 



k 



Y'x dx 
'm o 



£\w 



dx 



Appendix B. Elements of the A ~ Matrix 

The elements of the A - matrix, which defines 
the equations of motion when written as first order 
differential equations, are given below: 

A 21 - 1; A 22 = A 23 = A 24 = 

A 43 - lj A 41 - A 42 - A 44 - 

A ll ""®D1 + 2 ( l } H<- F9sin, (' + F 24 h°cos\J) ) 

hi' ~(4 + f) + \ ¥\^«* *♦ + ^ 6cos *) 

2 "1 

+ | F 23 h° (1+cos 2M 

A 13 = y | (|) 3 I -2euF 11 siniiH-p(e p +e D )F 11 cosi(- 
+ F 1A g°u cosJ 

A 14 - ^ + | (|) ? Je(-2F 21 uco8i( ) -u 2 sin2tpF 22 ) 

2 

+ XuF 22 cosi|; +ij (l+cos2i)»)[F 23 g°+F 22 (ep+e D )] 

A 31 = Y G + \ (f) 3 [9L 8 u sinip-2uL 17 g°cos*-6yL 22 h°cosij) 
- 2vKe p +e D )L 8 cosij/] 



_£8 . 2 



+ i(i 



* 32 ~ *u ' 2 



L U sin 2ifH-L 10 y cosij;) 



2 

-2AL 1:L y costfi-e^ (1+cos 2^)L 21 h°- 

-y 2 (e p +3 D ) L 11 (1+cos 2i|>) -L 24 ]i 2 g° (l+cos2i|0j 

c 

A 33 = -ijj^ X (£) 3 [-9uL 16 g°cos*-2 -22- uL 13 siniJ ) 



0p(e p +6 D )L 13 coaiil 



^34 



(4" ^) + i <|) 2 '{-e^ 20 cos^ + f 2 (i 



"li' 

+cos2i|>)[-L 21 gi- (g p +B D )L 20 ]e+ -£■ (-)A 20 sin2.J. 
-2u L 23 cosi|))| 



63 



where 
21 



2P e° 
z lll g l 



V" 



f- (|) 3 [26F 10 -F 11 X]+ S- 2(e_+fi ) 
«F1 2M F1 V 

ZklH!l + _X_ (|) 3 [ L 7 9_2AL 8 ]- ll 2(g 1J +e_) 
2M L1 ^1 



"ta 



Appendix C. Approximate Reverse Flow Model and 
the Associated Aerodynamic Loads 

the circular region of reversed flow, which 
exists over the retreating blade, is quite well 
known. In past treatments of reversed flow it has 
been customary 3 to define three separate regions: 
(a) normal flow, (b) reversed flow, (c) mixed flow, 
and evaluate the appropriate aerodynamic expres- 
sions for each region. When this model is used 
together with a modal representation of the blade 
the evaluation of the generalized aerodynamic 
expressions &■,!?■ becomes quite cumbersome, and a 
more convenient procedure had to be devised. 

The approximate reverse flow model developed . 
for the present study consists of replacing the 
circular region, by an approximate region which is 
a circular sector as shown in Figure 1. The 
approximation is based on the assumption that the 
area contained in the circular sector must be equal 
to the area contained in the approximate region. 
Two separate cases must be considered: (1) y < 1, 
(2) u > 1. 



Case (1). For u < 1, the radius of the circular 

part is taken as y. Simple geometric consid- 
erations show that the angle a is always a 
constant . 

given by a = ir/2 

Case (2) . For y >^ 1 simple geometric considerations 
show that 



j. 2 -11 r 2 
a = it - 2 sin (— ) + y sin (— ) - vp -1 



v 



Thus, for y < 1 the generalized aerodynamic 
loads are calculated from 



Api 



hi 



a\ 



u\ 



" / Vk dx o + / Vi dx o 
- J k *y J 

- H. 5 n 

- / L Y, dx + / L Y. dx 
J y 1 o J y'i o 

La I, J 



while for y >_ 1.0 

hi " _A Fi and *Li = -^Li 

These expressions are based on the assumption that 
the lift curve slope in the reversed flow region 
is equal to the negative value of the lift curve 
slope in normal flow. 




« r — ' . » 

A v .j X, i 



TOP VIEW 

Figure 1. Displacement Field of a Torsionally Rigid Cantilevered 
Blade with Droop and Preconing. 



<a)(*<1 



k 




(b)n>1 



Figure 2. Geometry of Approximate and Exact Reverse Flow 
Regions. 



64 



1. ALL (cjj) TERMS IN LAG EQ. NEGLECTED 

2. ALL TERMS INCLUDED 

3. ONLY (eg) TERMS ASSOCIATED WITH 
DAMPING INCLUDED 




Effect of Third Order Terms in the Lag Equation on 
Characteristic Exponent for Lag. 



WITH REVERSE FLOW 




<3pj« 1.175 




u u - 1.28303 




y - 10.0 




o - 0.05 
Op - 0.0 


WITHOUT 


REVERSE FLOW 


e - .« 




"SF, " •<" 




"SL,- 01 





Figure 4. Effect of Reversed Flow on Characteristic Exponent 
for Lag. 



-1.6 
-1.2 
-0.8 


- 


"F 
U L 
7 
o 

h 
Po 


- 1.176 

- 1.28303 

- 10.0 

- .06 


^7- RADIAL FLOW TERf 
y / INCLUDED 


-0.4 


- 




— — u k CONTINUOUS — W 


' V RADIAL FLOW TER! 
/ NOT INCLUDED 







1 1 1 1 


1 T v ^ 1. 






0.2 0.4 0.6 0.8 


1.0 1.2 L4» 




Figure 5. Effect of Radial Flow Terms on Characteristic 
Exponent for Flap. 



Figure 6. Effect of Radial Flow Terms on Characteristic 
Exponent for Lag. 



5 F1 


- 1.175 


"L1 


- 1.28303 


T 


- 10.0 


a 


- .05 


"P 


- 0.0 


0„ 


- 0.0 


B 


- .16' 




Figure 7. Effect of Viscous Structural Damping on Characteristic 
Exponent for Lag. 




.005 



.010 



.015 



ELASTIC COUPLING, 
EXACT MODE SHAPE 
NO ELASTIC COUPLING. 
EXACT MODE SHAPE 
NO ELASTIC COUPLING. 
ASSUMED MODE SHAPE 
_1 
.025 



.020 



'SF^-'SL, 



Figure 8. Critical Advance Ratio /u c vs Structural Damping 
Coefficients jj sf , 1J SL ■ 



65 



,f c 



< 


-1.0 


cc 
o 

u. 


-0.8 


z 

UJ 

z 


-0.6 


£ 
X 

Ul 


-0.4 






UJ 

3 


-0.2 



< 
z 
u 


0.2 



r 




X 
X 


a F1" 


1.175 




- 




X 
X 
X 
X 


S L1" 


1.28303 

10.0 

.05 




- 




X 


h - 


0.0 








X 
>» 


%> - 


0.0 








w k CONTINUOUS 


' 


.05 
.20 






i 


cJi, - 0.0 OR 1.0 
IT 1 1 


N^ 1 


I B 




0,2 


0.4 0.6 0.8 1.0 


1.2 1.4 1.6 


1.8 2.6' 


-2.2 


_ 













1.0 



0.8 



0.6 



0.4 



Opi» 1.09 
GJ L1 = 1.00 
y = 10.0 
a - .05 
(Jp - 0.0 

e = .is 

I 



_L 



-L 



-2.0 



-1.5 



0.0 



Figure 9. Effect of Collective Pitch on Typical Case. 



-1.0 -0.5 

D (DEGREES) 

Figure 10. Effect of Droop on u 



0.5 



ASSUMED MODE SHAPE 
NO ELASTIC COUPLING 




WITHOUT ELASTIC 
COUPLING 



Figure 1 1 . Effect of Exact Mode Shape and Elastic Coupling 
on Characteristic Exponent for Lag. 



66 



CORRELATION OF FINITE-ELEMENT STRUCTURAL DYNAMIC 
ANALYSIS WITH MEASURED FREE VIBRATION CHARACTERISTICS 
FOR A FULL-SCALE HELICOPTER FUSELAGE 

Irwin J. Kenigsberg 

Supervisor - Airframe Dynamics 

Sikorsky Aircraft 

Stratford, Connecticut 

Michael W. Dean 

Dynamics Engineer 

Sikorsky Aircraft 

Stratford, Connecticut 

Ray Malatino 
Helicopter Loads and Dynamics Engineer 
Naval Air Systems Command 
Washington, D. C. 



Abstract 

Both the Sikorsky Finite -Element Airframe 
Vibration Analysis Program (FRAN /Vibration Analy- 
sis) and the NASA Structural Analysis Program 
(NASTRAN) have been correlated with data taken 
in full-scale vibration tests of a modified CH-53A 
helicopter. With these programs the frequencies 
of fundamental fuselage bending and transmission 
modes can be predicted to an average accuracy of 
three percent with corresponding accuracy in 
system mode shapes . 

The correlation achieved with each program 
provides the material for a discussion of modeling 
techniques developed for general application to 
finite-element dynamic analyses of helicopter 
airframes. Included are the selection of static 
and dynamic degrees of freedom, cockpit structural 
modeling, and the extent of flexible- frame model- 
ing in the transmission support region and in the 
vicinity of large cut-outs . The sensitivity of 
predicted results to these modeling assumptions 
is discussed. 



cut-outs and concentrated masses such as the trans- 
mission, main rotor, and tail rotor, which are 
unique to helicopters, play a major role in con- 
trolling vibrations . 

Although advanced analytical methods based on 
finite-element techniques have been developed for 
studying the vibration characteristics of complex 
structures , a detailed correlation of such methods 
with test data is not available in the general 
literature. Further, little information is avail- 
able on the accuracy of various modeling assump- 
tions that might be made to reduce the cost and 
time of applying these vibration analyses . 

As a result a research project was establish- 
by Naval Air Systems Command with Sikorsky Aircraft 
to: 

a) Determine the accuracy of the Sikorsky 
Finite-Element Airframe Vibration 
Analysis in predicting the vibration 
characteristics of complex helicopter 
"airframe structures. 



Introduction 

Helicopter vibration and resulting aircraft 
vibratory stress can lead to costly schedule 
slippages as well as to problems in field service 
maintenance and aircraft availability. At the 
core of vibration control technology is the require- 
ment to design the helicopter structure to minimize 
structural response to rotor excitations . Both the 
complexity of the structure and the increasingly 
stringent mission and vibration control specifica- 
tions demand development of airframe structural 
vibration analyses that can be used rapidly and 
economically to evaluate and eliminate vibration 
problems during the preliminary design phase of 
heli copters . 

The complex helicopter structure consists of 
sections that differ considerably in structural 
arrangement and load carrying requirements . These 
sections include the cockpit, cabin, tail cone, 
and tail rotor pylon. In addition, large fuselage 

Presented at the AHS/NASA-Ames Specialists ' 
Meeting on Rotorcraft Dynamics, February 13-15,197^ 



and 

b) Develop and evaluate general helicopter 
dynamic modeling techniques that could 
be used to provide accurate estimates 
of vehicle dynamic characteristics while 
at the same time minimizing the com- 
plexity and cost of the analysis . 

Due to the increased usage of NASTRAN 
throughout the industry as well as the efficiency 
resulting from employing a single analytical sys- 
tem for both static and dynamic analyses , a par- 
allel correlation study using NASTRAN has been 
performed. The results of these correlation 
studies are the subject of this paper. 

Phase I - Stripped Vehicle 

Test Vehicle 

At the initiation of this effort, the phi- 
losophy guiding the development of modeling tech- 
niques was based upon the concept of gradually in- 
creasing the complexity of the analytical repre- 
sentation. It was decided that the first 



67 



correlation study would be conducted on an air- 
craft stripped of all appendages. It was believed 
that the modeling techniques for representing air- 
frame response characteristics could be identified 
and developed most easily in this manner. Then, 
as various appendages were added to the basic 
vehicle, only the modeling techniques required for 
the structure or masses added need be developed. 

The vehicle used in this test and correla- 
tion study was the CH-53A Tie Down Aircraft, Ve- 
hicle designation number 613. A general arrange- 
ment of the structure is illustrated in Figure 1. 
For initial correlation, all appendages were re- 
moved. These included the nose gear, main landing 
gear, main landing gear sponsons, fuel sponsons, 
tail pylon aft of the fold hinge, tail rotor and 
associated gear boxes, engines, cargo ramp door, 
horizontal stabilizer, and all remaining electri- 
cal and hydraulic systems. The main rotor shaft 
and all gears were removed from the main transmis- 
sion housing and only the housing itself was re- 
tained for the test configuration. 

Testing 

The ground test facility employed to estab- 
lish the dynamic characteristics of the test vehi- 
cle was a bungee suspension system that simulates 
a free-free condition, a rotorhead-mounted uni- 
directional shaker, and the Sikorsky shake test 
instrumentation console. Instrumentation con- 
sisted of lit fixed and 10 roving accelerometers . 
A complete description of the test apparatus and 
the instrumentation is provided in Reference 1. 

All accelerometer signals and the reference 
shaker contactor signal were transmitted to the 
console. The signals were processed automatically 
by the console resulting in a calculation of the 
in-phase and quadrature components of the acceler- 
ations. The accelerations were then normalized to 
the magnitude of the shaker force at the particu- 
lar frequency. As frequency was varied, the re- 
sulting response of each accelerometer was record- 
ed on a XYY' plotter, Figure 2, as g's/1000 lbs. 
versus frequency. 

Ideally, a fuselage mode can be identified 
by a peak in the quadrature response and a simul- 
taneous zero crossing of the in-phase response. 
Once a mode is located, all quadrature responses 
at this frequency can be recorded to define the 
mode shape. The modes defined in this manner from 
the shake tests are listed in the left-hand column 
of Table I. It should be noted that this tech- 
nique is applied more easily at lower frequencies, 
where sufficient modal separation exists so that 
the forced response in the vicinity of a resonance 
is dominated by a single mode. As shown in Figure 
2, the mode shapes at higher frequencies must be 
extracted from the coupled response of many modes. 

Analysis and Correlation 

The shake test data indicated that the 
natural modes of vibration of a helicopter can be 
categorized as beam-like modes controlled by 



overall fuselage characteristics (e.g., length, 
depth, overall bending stiffness, mass distribu- 
tion, etc.) and those controlled by the transmis- 
sion support structure. Therefore, the overall 
helicopter structure was modeled utilizing three 
modules : 



1) 



and 



3) 



center section including the transmission 
support region 

forward fuselage and cockpit 



aft fuselage and tail. 



The center section was modeled in greatest detail 
by applying finite-element techniques . The struc- 
tural characteristics of the forward and aft fuse- 
lage were derived from beam theory. These equiva- 
lent beams were located at the neutral axis of the 
airframe section and were assigned the bending and 
torsional properties of the total section. The 
beam models of the forward and aft fuselage were 
cantilevered from rigid frames at the respective 
forward and aft ends of the center section, Figure 
3. The influence coefficients of these beams with 
respect to their cantilevered ends were then com- 
bined with the influence coefficient matrix of the 
remaining structure. 

The Phase I correlation was performed using 
the Sikorsky Finite-Element Airframe Vibration 
Analysis (FRAN /Vibration Analysis) . This analysis 
consists of two programs: PPFRAH and a 200 dynamic- 
degree-of-freedom eigenvalue/eigenvector extraction 
procedure. PPFRAH is derived from the IBM/MIT 
Frame Structural Analysis Program, FRAN (Reference 
2), a stiffness method, finite-element analysis 
limited to two types of elements , namely bending 
elements (bars) and axial elements (rods). This 
limitation necessitated further development of FRAN 
for application to stressed skin structures. This 
development consists of the addition of pre- and 
post-operative procedures linked to FRAN. In the 
pre-operative procedure (Pre-FRAN), the fuselage 
skin is transformed into equivalent rod elements. 
This transformation is developed by satisfying the 
criterion that the internal energy of the skin 
structure under an arbitrary set of loads be the 
same as that of the transformed structure under the 
same set of loads. The post-operative procedure 
(Post-.PRAN) extracts the influence coefficient 
matrix corresponding to the selected dynamic 
degrees of freedom. A detailed description of the 
FRAN /Vibration Analysis is provided in Reference 1. 

The elements used to represent the airframe 
structure are: 

1) bending (bar) elements for fuselage frames 
and for the nose and tail beams 

2) axial (rod) elements for. the stringers 
and 



3) equivalent, diagonal rod elements for skin 
panels . 



68 



For dynamic analysis, the structure is as- 
sumed to be unbuckled, so that all skin panels are 
considered fully effective in resisting axial 
loads. Thus, the total axial area of each skin 
panel is lumped with the areas of adjacent string- 
ers. 

During Phase I correlation, three modeling 
parameters were varied: the number of bays over 
which the finite-element (flexible -frame) model 
extends (Figure k) , the number of nodes per frame 
(number of stringers), and the number of dynamic 
degrees of freedom assigned to each frame (Figure 
5). The results of the correlation are presented 
in Table I, which shows the sensitivity of the 
analysis to each of the above parameters and the 
accuracy of the predicted frequencies and mode 
shapes. The criteria for establishing the level 
of mode shape correlation are: 

E (Excellent) - Correct number of nodes, nodes 

less than 2.5 percent of fuselage 
length from measured location, 
local modal amplitudes within 20 
percent of test values. 

G (Good) - Correct number of nodes, nodes 

less than 2.5 percent of fuselage 
length from measured location, 
difference between actual and pre- 
dicted local modal amplitudes ex- 
ceeds ±20 percent of test values . 

F (Fair) - Correct number of nodes, nodes 

more than 2.5 percent of fuselage 
length from measured location, 
difference between actual and pre- 
dicted local modal amplitudes ex- 
ceeds ±20 percent of test values. 

P (Poor) - Incorrect number of nodes, nodes 

located improperly, difference be- 
tween actual and predicted local 
modal amplitudes exceeds ±20 per- 
cent of test values. 

A comparison of the 30- and 60-stringer anal- 
yses indicates that there is no change in the re- 
sults when modeling the structure with half the 
number of actual stringers. In addition a compari- 
son of results obtained with the basic and reduced 
dynamic degree of freedom allocation indicates that 
no more than 16 dynamic degrees of freedom per 
frame are required for dynamic modeling. 

Although mode shape correlation resulting 
from the analysis in the frequency range of inter- 
est (below 1500 cpm) is encouraging, see Table I, 
the absence of a representative mass distribution 
made the analysis overly sensitive to certain mod- 
eling assumptions. This sensitivity appears to 
account for the less than satisfactory frequency 
correlation. For example, the frequency of the 
transmission pitch mode is normally controlled by 
the mass of the fully assembled transmission and 
the properties of the structure in the transmission 
support region. In the absence of a mass distribu- 
tion representative of a fully assembled vehicle, 
however, any element of the structure and any 



lumped mass can contribute significantly to the 
control of the dynamic characteristics. In this 
case, the analytical representation appears to be 
too stiff because of the beam model used for the 
fuselage forward of F.S. 262, which constrains the 
upper and lower decks to deform equally. This 
constraint is not imposed by the actual structure. 
A comparison of the results of the 9- and l8-bay 
analyses indicates that due to the local nature of 
the transmission pitch mode, extension of the 
flexible-frame model aft beyond the limit of the 
9-bay model has no significant effect on the pre- 
diction of this mode. 

The poor frequency correlation for the first 
lateral bending mode persisted throughout this 
phase of correlation. This mode was characterized 
by differential shearing of the upper and lower 
decks of the aft cabin, Figure 7. The 6-bay and 
9-bay flexible-frame model represented most of this 
structure experienceing the differential shearing 
as a beam capable of only bending and torsion. 
This overly constrained model resulted in predicted 
frequencies substantially higher than test values. 
Extending the flexible-frame representation to 18- 
bays appears to be the solution. However, size 
limitations in PPFRAN required that the l8-bay 
flexible-frame model be generated in two 9-bay sub- 
structures , married at a rigid intermediate frame 
at F.S. kk2, Figure 3. Although the extended model 
improved the correlation of the first lateral bend- 
ing mode, absence of a representative mass distribu- 
tion again appears to make the model overly* sensi- 
tive to the presence of the rigid frame at F.S. hk2. 
This accounted for the remaining difference between 
test and analysis. 

Many of the higher frequency modes are con- 
trolled by the structure in the area of the rear 
cargo ramp. This accounts for the failure to pre- 
dict the Transmission Vertical mode until the flex- 
ible-frame model was extended into the ramp area, 
see Table I. Although this extension of the model 
improved correlation, the high frequency modes 
above 1500 cpm are difficult to identify analytical- 
ly due to the coupling of overall fuselage modes 
with local frame modes . This difficulty is com- 
pounded in this investigation, because the fre- 
quencies of the basic fuselage modes are raised due 
to the stripped condition of the vehicle, while 
frequencies of the local frame modes are lowered 
due to the lumped-mass modeling used to represent 
each frame. Tests of a more representatively load- 
ed fuselage can be expected to minimize the problem 
of mode identification. 

From the results of this phase of the corre- 
lation, it is concluded: 

1) The selection of static degrees of freedom in 
the flexible frame model can be based on a 
structural model that contains stringers num- 
bering one half the number of actual 
stringers . 

2) No more than sixteen dynamic degrees of free- 
dom on each flexible frame are required for 
dynamic analysis. The typical location of 



69 



these degrees of freedom is Illustrated in 
Figure 5- 

Transmission modes can be predicted by a 
flexible-frame representation of the trans- 
mission support region extending about 1.5 
transmission lengths forward and aft of the 
corresponding transmission supports, about 
9 bays. If the vehicle contains large cut- 
outs, such as the cargo ramp of the test 
vehicle, the flexible-frame model should ex- 
tend through this region as well. 



PHASE II - BALLASTED VEHICLE 



Testing 



Shake tests were performed after adding bal- 
last to provide a more realistic representation of 
a helicopter mass distribution, Figure 6. At the 
transmission mounting plate, two lead blocks hav- 
ing a total weight of 1*570 pounds were mounted so 
that the mass and pitching moment of inertia of 
the simulated transmission and rotor head approx- 
imated that of the actual CH-53A. At the tail, a 
1500-pound block was mounted to simulate the re- 
moved tail pylon, stabilizer, and tail rotor. At 
the nose, a 3000-pound block was mounted on the 
nose gear trunnion fitting to balance the vehicle. 

The natural modes of vibration identified by 
shake tests are listed in Table II along with the 
frequencies measured during Phase I. Not only did 
the ballast succeed in lowering the fuselage modes 
into a frequency range more representative of that 
encountered on a fully assembled aircraft, but ad- 
ditional modes were also identified that are 
strongly controlled by the ballast. In fact, 
these modes were identified as local modes of the 
ballast blocks themselves. Due to the complex 
structural nature of the ballast, Figure 6, these 
appendages did not lend themselves to simple ana- 
lytical representations. Therefore, the flexi- 
bility of each ballast block was measured by in- 
strumenting both the block and the adjacent air- 
frame structure and then measuring the accelera- 
tions occurring at both locations near the modal 
frequencies of interest. The mass of each ballast 
block and its absolute acceleration resulted in a 
force which produced the relative motion between 
the two instrumented parts. The empirically de- 
fined flexibilities of the ballast were then used 
in the dynamic model. 

Analysis and Correlation 

The modeling techniques developed in Phase I 
of this study were applied to both the FRAN/Vibra- 
tion Analysis and NASTRAH. 

The finite-element model analyzed in Phase 
II was identical to the 18-bay model analyzed in 
Phase I, except for adding the mass and structural 
characteristics of the ballast blocks. The FRAW 
model was formed with rod and bar elements, as 
discussed previously, while the HASTRAM model used 
CROD, CBAR, and CSHEAR elements (Reference 3). As 
before, all skin panels were assumed fully 



effective in reacting axial load and this effective 
area was lumped into the adjacent stringers. 

Including ballast, to replace removed appenda- 
ges resulted in a' substantial improvement in the 
correlation, particularly in frequency prediction 
as shown in Table III. Significantly, ballast 
eliminated the difficulties identified as sensitivi- 
ty to modeling assumptions and local frame modes in 
the absence of representative mass distributions. 
The average error in predicting the frequencies of 
fundamental fuselage bending modes and the trans- 
mission pitch mode was 3-W for both the FRAN/Vibra- 
tion Analysis and NASTRAH. In addition the shape 
correlation for these modes varied from good to ex- 
cellent. The analyses also were able to predict 
accurately the significant changes in the charac- 
teristics of the fuselage and transmission modes 
resulting from the addition of the ballast, Figures 
7, 8 and 9- To achieve this degree of correlation, 
modeling of the ballast flexibilities was required. 
This modeling was successfully accomplished in the 
vertical/pitch direction, Figure 10, but did not 
prove successful in the lateral/torsion direction, 
Figure 11. The contrast between these two results 
establishes the ability of finite-element analyses 
to predict accurately the characteristics of fuse- 
lage and transmission modes when the structural 
data base is defined with sufficient accuracy. Fur- 
ther improvement in the correlation could have been 
achieved if a more detailed definition of the bal- 
last flexibilities had been acquired from measure- 
ments of static deflections . 

Reasonable success has been achieved in pre- 
dicting higher frequency, ramp-controlled modes, 
Figures 12 and 13. However, some margin. does exist 
for further improvements in shape and frequency 
prediction. From the standpoint of modeling, it 
appears that the 200 dynamic degree of freedom 
limit established in this study is inadequate for 
predicting the shell-type modes of the cargo ramp 
structure. In addition, the test procedure em- 
ployed, namely the use of a single rotorhead 
shaker, does not provide a means of uncoupling the 
forced response characteristics of modes at the 
higher frequencies, Figure 2. 

Conclusions 



3. 



70 



Finite element analyses can predict accurate- 
ly the frequencies and mode shapes of complex 
helicopter structures, provided the structur- 
al data base is defined accurately. 

Complete stripping of a vehicle for correla- 
tion purposes may make the analysis overly 
sensitive to normally minor modeling assump- 
tions . 

Significant changes can be predicted accurate- 
ly in the character and frequency of fuselage 
and transmission modes due to changes in mass 
distributions and structural characteristics. 

The modeling techniques established by this 
study can be used during aircraft design re- 
gardless of the finite-element analytical 
system' being used. 



Recommendations 



References 



1) A full-scale shake test correlation should 
he performed on a fully assembled flight 
vehicle to establish and validate modeling 
techniques for those appendages removed 
during this study. 

2) Appendages not amenable to accurate or eco- 
nomical structural analysis should he 
tested statically to determine flexibility 
data required for dynamic analysis. 

3) Integrated structural design systems should 
he developed to couple static and dynamic 
analyses and thus provide the accurate 
structural data required for defining vibra- 
tory response characteristics as early as 
possible during aircraft design. 

k) Use of additional shaker locations should be 
incorporated in the test procedure to pro- 
vide a means of uncoupling higher frequency 
modes. Alternatively, more sophisticated 
means of processing shake test data (e.g., 
system identification techniques described 
in Reference k) should be employed. 



1) Kenigsberg, I. J., CH-53A FLEXIBLE FRAME 
VIBRATION ANALYSIS/TEST CORRELATION, 
Sikorsky Engineering Report SER 651195, 
March 28, 1973. 

2) IBM 7090/709!+ FRAN FRAME STRUCTURE ANALYSIS 
PROGRAM (7090-EC-OlX). 

3) McCormick, C. W. , ed. , THE NASTRAN USER'S 
MANUAL, (level 15), NASA SP-222(0l), June 
1972. 

1+) Flannelly, W. G., Berman, A., and Barnsby, 
R. M., THEORY OF STRUCTURAL DYNAMIC TESTING 
USING IMPEDANCE TECHNIQUES, USAAVLABS 
TR 70-6A,B, June 1970. 

5) Willis, T., FRAN CORRELATION STUDY, Sikorsky 
Report SYTR-M-36, July 1969. 



Illustrations 



Horizontal 
Stabilizer 



Tail Sotor 
Attachment 




Cargo 
Ramp Door 



-Nose Gear 



Sponsons 



Figure 1 CH-53A General Arrangement 

71 



o 
o 



3 

x 

(O 

CO 
_i 
o 
o 
o 

o 
-H 



0-. 



% •■ 



FULL SCALE 



3G/I000 LB. 
FULL SCALE" 




SECOND TRANSMISSION 
VERTICAL VERTICAL 



FULL SCALE 
I I 



■4- 



100 



200 300 400 500 



600 700 800 900 1000 1100 1200 1300 1400 1500 1600 1700 
FREQUENCY, CPM 



Figure 2 Typical Vertical Response to Vertical Excitation 



TABLE I 
PHASE I SHAKE TEST CORRELATION SUMMARY 



TEST 










ANALYSIS 
















18 Bay 
30 Stringer 
Reduced DOF 




9 Bay 
30 Stringer 
Reduced DOF 




6 Bay 
30 Stringer 
Reduced DOF 




6 Bay 
30 Stringer and 
60 Stringer 
Basic DOF 




Mode 


Freq. 
(CPM) 


Freq. Error 


Shape 


Freq. 


Error 


Shape 


Freq. 


Error 


Shape 


Freq. Error 


Shape 


1st Lateral 


910 


1207 33$ 


G 


1U66 


60% 


P 


1U35 


58% 


P 


11*1*0 58% 


P 


1st Vertical 


1155 


1175 2% 


E 


1282 


11% 


E 


121*2 


8% 


E 


121*1 8% 


E 


XSSN Pitch 


1U90 


1709 13% 


E 


1710 


13% 


E 


17^8 


17% 


E 


1758 17% 


E 


2nd Vertical 


1950 


2150 10% 


G 


2390 


22% 


F 


2505 


28% 


F 


2577 32% 


F 


XSSN Roll 


2000 
2150 


21405 20% 
2250 k% 


P 

F 


2870 


1*3% 


P 


2900 


1*5% 


P 


2891* 1*5% 


P 


XSSN Vertical 








Torsion 


2300 


2763 20% 


F 


2l*28 


6% 


P/F 


2U22 


6% 


P/F 


21*1*5 6% 


P/F 



72 



632 /V- RIGID FRAME 




Figure 3 Modular Representation 
of Helicopter Structure 



-RIGID FRAME 



F.S. tlO 




Figure *t Finite-Element Model, Transmission 
Support and Bamp Areas 



WL 163 
BL -5L91 



WLJ39. 
8L-52.92 



WLI09 
BL-52 



WL 189.7 
BL 31.81 



WL I9t 
BL ^20 



BASIC 



WL 163 
BL^!98 



WL 139 
BL 52.92 



WLI57 

BL-52.38 



iiy] 



WL JEt 
BL- 52.97 



71 T4 

WL 87.55 WL87 WL 87.55 

BL-35.7 BL BL 35.7 



V — L 



"7^ 



REDUCED 



>' 



-© : — : — S "§- 

1 A^ I 



mj57 

BL 52.! 



WL 139 
BL 52.9 



WLJ2I 
BL 52.97 



Figure 5 Degree of Freedom Locations for Basic and 
Reduced Dynamic Degree of Freedom Models 



WL 87.0J 
BL 22.86 




Figure 6b Tail Ballast 
Installation 



Figure 6c Nose Ballast 
Installation 



Figure 6 Phase II Ballast Installations 



74 



TABLE II - SHAKE TEST FREQUENCIES 



MODE 

1st Vertical Bending 

1st Lateral Bending 

Transmission Pitch 

Hose Block Lateral/Roll 

Hose Block Vertical/Coupled 

Forward Cabin/Nose Block Lateral 

Hose Block Vertical 

Second Vertical Bending 

Torsion 

Transmission/Ramp Vertical Bending 

Ramp Vertical Bending 



PHASE II 
TEST FREQUENCY CPM 


PHASE I 
TEST FREQUENCY 


kko 


1155 


615 


910 


7^0 


1U90 


930 





970 





990 





1050 





1290 


1950 


1310 


2300 


lte5 


2150 


16^0 






TABLE III 



PHASE II CORRELATION SUMMARY 



VERTICAL/PITCH MODES 



MODE 

1st Vertical Bending 

Transmission Pitch 

Hose Block Vertical/ 
Transmission Pitch 

Hose Block Vertical 

Second Vertical 

Transmission Vertical/ 
Ramp Vertical 

Ramp Vertical 



Frequency 


Test 


FRAN 


kko 


1*38 


TkO 


751 


970 


933 


1050 


101*3 


1290 


1523 


1U25 


1563 


161*0 


139 s * 



Error 



1.555 



10? 
15? 



Shape 


HASTRAN 


E 


1*53 


G 


785 


G 


956 


F 


1063 


F 


1608 


F/G 


181*3 


P/F 


1355 



Error 



1.5% 

155 
25? 

29? 

17? 



Shape 

E 
G 

G 
F 
F 

F/G 
P/F 



1st Lateral Bending 615 

Hose Block Lateral/Roll 930 

Forward Cabin Lateral/ 

Nose Block Lateral 990 

Torsion 1310 



LATERAL/T0RSI0H MODES 
659 7? & 
735 21? P 



858 
1601 



13? 
22? 



P 
P 



595 

812 

970 
1325 



3? 
13? 



G 
P 



75 





FRAN-1207 opm 



Figure 7a Correlation of First Lateral Bending 
Mode, Phase I - Stripped 




NASTBAU-595 cpm 



Figure 7b Correlation of First Lateral Bending 
Mode, Phase II - Ballasted 



76 




FHM-11T5 opm 



Figure 8a Correlation of First Vertical Bending 
Mode, Phase I - Stripped 




NASTBAN-453 opm 



•Figure 8b Correlation of First Vertical Bending 
Mode, Phase II - Ballasted 



77 




PRM-1709 cpm 



Figure 9a Correlation of Transmission Pitch Mode, 
Phase I - Stripped 




NASTRAN-785 cpm 



Figure 9t> 



Correlation of Transmission Pitch Mode, 
Phase II - Ballasted 



78 




HASTRM-956 cpm 



Figure 10 Correlation Hose Block Vertical/Transmission 
Pitch Mode, Phase II - Ballasted 




MASTRAU-812 cpm 



•Figure 11 Correlation of Hose Block Lateral/Roll 
Mode, Phase II - Ballasted 



79 




Phase II test-1290 cpm 




HASTMM-1608 cpm 




NASTRAH-1355 cpm 



Figure 12 Correlation of Second Vertical Bending 
Mode, Phase II - Ballasted 



Figure 13 Correlation of Ramp Vertical Bending 
Mode, Phase II - Ballasted 



80 



COUPLED ROTOR/AIRFRAME VIBRATION PREDICTION METHODS 

J. A. Staley 
Senior Dynamics Engineer 

J. J. Sciarra 
Senior Structures Engineer 

Boeing Vertol Company 
Philadelphia, Pa. 



Abstract 

The problems of airframe structural 
dynamic representation and effects of 
coupled rotor/airframe vibration are dis- 
cussed. Several finite element computer 
programs (including NASTRAN) and methods 
for idealization and computation of air- 
frame natural modes and frequencies and , 
forced response are reviewed. Methods for 
obtaining a simultaneous rotor and fuse- 
lage vibratory response; determining 
effectiveness of vibration control devices, 
and energy methods for structural optimi- 
zation are also discussed. Application of 
these methods is shown for the vibration 
prediction of the Model 347 helicopter. 

Notation 

A - airframe mobility matrix 

B - rotor impedance matrix 

EI - blade bending rigidity 

F - force 

GJ - blade torsional rigidity 

I - identity matrix 

k - rotor frequency multiple, 1, 2, etc. 

K - stiffness matrix 

M - mass matrix 

q - airframe mode generalized coordinate 

X - airframe displacements 

rn - airframe mode generalized mass 

w - airframe mode natural frequency 

$ - airframe mode shape (eigen vector) 

fl - rotor frequency 

[] - matrix 

{ } - column vector 

Subscripts 

A - absorber , airframe 

c - cosine component amplitude 

H - hub 

k - rotor frequency multiple, 1, 2, etc. 

n - airframe mode number 

o - zero hub motion 

R - rotor 

s - sine component 

Presented at the AHS/NASA-Ames Specialists' 
Meeting on Rotorcraft Dynamics, February 
13-15, 1971. 

Part of the work presented in this paper 
was funded by the U.S. Army Research 
Office - Durham, North Carolina under 
Contract DAHC04-71-C-0048. 



Superscripts 

. - velocity 
. . - acceleration 
T - transpose 

Prediction of helicopter airframe 
vibration involves two major problem 
areas s 

• Prediction of rotor vibratory 
hub loads 

• Prediction of airframe dynamic 
characteristics . 

The effects of vibratory hub motion on 
vibratory hub loads and effects of vibra- 
tion control devices and resulting air- 
frame fatigue stresses must also be con- 
sidered. 

Methods for independent prediction of 
vibratory hub loads and airframe dynamic 
characteristics have been developed pre- 
viously and are discussed briefly below. 
Independent determination of rotor vibra- 
tory loads and airframe vibratory response 
to these loads does not account for any 
interaction between airframe vibratory 
motion on rotor vibratory loads. One 
approximate method for accounting for 
these interactions is to assume that an 
effective rotor mass is attached to the 
airframe at the rotor hub. A more direct 
method is to compute (or measure) the 
rotor hub impedance and determine compat- 
ible vibratory hub loads and hub motions. 
This method is discussed below. A simple 
example of compatible rotor load-hub 
motion is given for a single rotor heli- 
copter with vertical hub motion. In 
addition, flight test results for the 
Model 347 helicopter are compared with 
vibration predictions obtained using a 
coupled rotor/airframe vibration computer 
program. 

Rotor Vibratory Hub Loads 

Methods and digital computer programs 
have been developed for prediction of 
rotor vibratory hub loads for constant 
speed level flight conditions 1,2,3. 
Rotor blades are represented by lumped 
parameter analytical models as indicated 



8.1 



in Figure 1. Iteration techniques are 
used to compute individual blade deflec- 
tions and aerodynamic and inertia load 
distributions at integer multiples of the 
rotor rotating frequency. The total 
rotating and fixed system rotor vibratory 
hub loads are obtained by summing indivi- 
dual blade root shears and moments. The 
vibratory hub loads may be computed assum- 
ing no hub motion. If the vibratory hub 
motions are known, effects of these 
motions may be included when computing 
blade aerodynamic and inertia loads. 

Airframe Dynamics 

Structural Model, Natural Modes and Fre- 
quencies , ' ana Forced Response 

Finite element methods have been used 
in the helicopter industry for some time 
for prediction of airframe dynamic charac- 



teristics^ 



As indicated in Figure 2, 



developing a finite element airframe model 
consists of: 

• Defining nodal data 

• Defining elastic properties of 
members connecting nodes 

• Defining mass properties asso- 
ciated with each node. 

Nodal data and properties of struc- 
tural members are used to develop stiff- 
ness matrices for individual members. 
These matrices relate forces at each node 
to nodal displacements. The stiffness 
matrices for individual members are super- 
imposed to obtain the stiffness matrix 
for the entire airframe. 

Most of the degrees of freedom are 
reduced from the airframe gross stiffness 
matrix. Mass properties are concentrated 
at the remaining (retained) degrees of 
freedom. Equations (1) are the airframe 
equations of motion with the gross stiff- 
ness matrix. Equations (3) are the air- 
frame equations of motion, in terms of 
the reduced stiffness matrix. 



(1) 



fcj = [ K n] - L K 12] fri "^21] < 2 > 

[M] {xj + [Kj {X X } ={f r } _ (3) 



M 0" 




Xl" 


+ 


hi K i2 


M 


• — 


Fr 


0_ 




1*2 




?21 K 22. 


l X 2j 








The solution for natural modes and 
frequencies is made using the reduced 
stiffness matrix and the mass matrix 
associated with the retained degrees of 
freedom. 

The airframe motions are expressed 
in terms of natural modes s 



M = [♦] M 



(4) 



and, after assuming sinusoidal motion 
with no external forces, Equation (3) 
becomes : 



-1 



(5) 



The modal generalized mass is then 
computed. A value of modal damping is 
assumed for each mode, and these modal 
properties are used to compute airframe 
response to vibratory hub loads: 



*» = N T I M ] W 



(6) 



q n + 25 n w n q n + w n q n = [4^ {Fr}/^ (7) 



Substructures Method 

A large saving in computer time can 
be realized by performing the matrix 
reduction process on several smaller sub- 
structure stiffness matrices instead of 
on the large stiffness matrix for the 
entire airframe. In one application, use 
of the substructures method reduced com- 
puter running time from about ten to two 
hours on an IBM 360/65 computer. 

The airframe is divided into several 
substructures, and all but mass and 
boundary degrees of freedom are reduced 
from the stiffness matrix of each sub- 
structure. The stiffness matrices of the 
substructures are then merged (super- 
imposed or added just as they are for 
individual members) to form a stiffness 
matrix for the entire airframe. Any 
degree of freedom on the boundaries may 
be reduced after merging the substructure 
matrices (Figure 3) . 

NASTRAN 

New developments in finite element analy- 
sis have been occurring on a continuous 
basis. New programs and new structural 
elements, both dynamic and stress analy- 
sis capability, FORTRAN programming cap- 
ability by the engineer within the finite 
element program, and greater problem size 



82 



capability have been developed . NASTRAN 
(NASA Structural Analysis) 6 is a govern- 
ment developed, maintained, and continu- 
ally updated finite element program which 
has apparently provided a solution to the 
difficulties of developing and maintain- 
ing finite element programs by private 
contractors. NASTRAN is similar to other 
finite element computer programs except 
that it generally provides additional 
capability: 

• More types of structural elements 

• Common deck for stress and 
dynamic analysis 

• User programming capability 

• Transient vibration analysis, 
buckling, non-linear, and static 
capability 

• Unlimited size capability for 
mass and stiffness matrices. 

For a nominal fee, this program and 
manuals describing the program and its 
use are available. NASTRAN provides a 
standard for airframe dynamic analysis 
and relieves contractors of some of the 
problems of maintaining the most up-to- 
date methods for airframe structural 
analysis. 

Energy Methods for Structural Optimization 

One further development related to 
airframe dynamics is the Damped Forced 
Response Method 7 / 8 . The airframe forced 
response is computed, and structural 
members with significant strain energy 
are identified. These members are 
changed to reduce vibration response for 
modes with frequencies above and below 
the rotor exciting frequency. This 
method is outlined in Figure . 4 . 

Vibration Control Devices 

Vibration control devices such as 
absorbers are often used to reduce vibra- 
tion in local areas of the airframe. 
The force output for an absorber may be 
computed by expressing the vibration as 
the sum of vibration due to rotor forces 
and the vibration due to the force output 
by the absorber. 



{ f a} - -[ a aa] _1 |^ar| { f r} 



(9) 



X* 



X r 



A AA a ar 



a ra a rr 



(8) 



The absorber force output required to null 
vibration at the absorber attachment 
point is 



The corresponding motions at the rotor 
hub are 

{%} " [ A RR- A RA a aa A AR ] { F f} ("J 

The mobility matrices in the above equa- 
tions may be obtained analytically using 
computed modal properties (Equations (1) 
through (7)) or by applying unit vibra- 
tory loads to the airframe in a series of 
shake tests. 

This method was applied to predic- 
tion of cockpit vibration with a vertical 
cockpit absorber for the Model 347 heli- 
copter°. Analytical and flight test re- 
sults are compared in Figure 5*. 

Coupled Rotor/Airframe Analysis 

Theory 

Any vibratory motion of the rotor 
hub will change the rotor blade vibratory 
aerodynamic and inertia forces which are 
summed to obtain vibratory hub loads. 
Changes in hub loads will in turn cause 
changes in vibratory hub motions^'-*-". 

Airframe Motion is assumed to be 
related to vibratory hub loads by a 
mobility matrix for a particular exciting 
frequency: 



^ks 



%c 



A kll A kl2 
A k21 A k22 



r ks 
F kc 



= [Ak] 



■ks 



-kc 



(11) 



where 
{ x k}={ x ks} sin knt + ( X kc( cos knt 

{ F k} = { F ks} sin knt + { F kc} cos knt 

The airframe mobility data are air- 
frame responses to unit vibratory hub 
loads; these data may be obtained analy- 
tically by using theoretical modal proper- 
ties (Equation (4) through (7)), or by 
conducting an airframe shake test. It is 
emphasized that these are airframe 
response characteristics for no blade 
mass attached to the airframe at the 
rotor hub. All blade inertia effects 
will be included in the rotor vibratory 
hub loads as modified by vibratory hub 
motion. 

In general, six sine and six cosine 
components of shaking forces and moments 
exist at each rotor hub; a tandem rotor 
helicopter would have a total of 24 



83 



components of vibratory forces. If only 
the rotor hub motions are considered, the 
relationship between hub motion and hub 
forces is: 



24x1 



Tiks 



Hike 



24x24 24x1 



= [%k] 

\ 



; ks 



? kc 



(12) 



The vibratory hub loads are assumed 
to be loads with no hub motion plus an 
increment of hub loads proportional to hub 
motion : 




24x1 
F kso 
F kco 

F kso 
F kco 



24x24 
B kll B kl2 
Bk21 B k22 



24x1 

.1 

x Hks 
x Hkc 



(13) 



H 



*-Hks 



x Hkc 



The coefficients of the B matrix are 
obtained by making several computations of 
vibratory hub loads : 

• Components of vibratory hub loads 
are computed assuming no hub 
motion 

• Components of vibratory hub loads 
are computed assuming a small 
vibratory hub motion at the fre- 
quency for each degree of freedom 
of hub motion at each rotor 

• Changes in sine and cosine com- 
ponents ' of vibratory hub forces 
per unit vibratory' hub motion in 
each rotor hub degree of freedom 
are then computed. 

The coupled rotor/airframe solution 
for compatible rotor hub motions and rotor 
hub loads is obtained by substituting ' 
Equation (13) in Equation (12) and solving 
for vibratory hub motions: 



*Hks 



x Hkc 



-r-1 



[I]-[A][B] 



[A] 



b kso 



e kco 



(14) 



Once a solution for Equation (14) is 
obtained, the total vibratory hub loads 
may be computed using Equation (13) and 
the vibration for the entire airframe may 
be computed using Equation (11) . 



Single Rotor Example 

Figure 6 shows a simple example of 
the coupled rotor airframe method applied 
to a single rotor helicopter vertical 
vibration analysis. Hub vertical vibra- 
tion response and the vertical vibratory 
hub loads are computed at a frequency of 
four times rotor speed (4/rev) . The air- 
frame is represented by its rigid body 
vertical mode and one flexible mode. 
Figure 6b shows airframe mobilities vs 
flexible mode natural frequency for 4/rev 
vertical hub forces. Hub vertical shak- 
ing forces vs hub vertical motion are 
shown in Figure 6c. The vibratory hub 
loads are seen to vary approximately 
linearly at least up to .005 inches of 
motion at the 4/rev frequency. Figures 
6d and 6e show compatible rotor hub 
vertical vibration amplitudes and rotor 
hub shaking forces vs flexible mode 
natural frequency. 

For this example, the rotor vibra- 
tory hub motions and forces both peak 
when the flexible mode natural frequency 
is just above the rotor hub force excit- 
ing frequency. This is not a general 
result, but depends upon the relation- 
ships between hub shaking forces and hub 
motions . 

Coupled Rotor /Air frame Analysis Computer 
Program (D-65) 

Figure 7 shows the flow-diagram for 
the Boeing Vertol D-65 Coupled Rotor/ 
Airframe Analysis computer program. This 
program links three major computer 
programs-*^. 

Trim analysis program A-97 

Rotor vibratory hub loads analy- 
sis program D-88 

- Airframe forced response analy- 
sis program D-96. 

Compatible fuselage motions and 
vibratory hub loads are obtained using 
this program with the method discussed 
above. In its current state, the D-65 
program computes three vibratory rotor 
forces and three vibratory rotor moments 
at each rotor for either single or tandem 
rotor helicopters. Response to trans- 
lational and rotational vibratory hub 
forces is computed for the airframe, but 
compatibility of hub forces and motions 
is satisfied for hub translational 
degrees of freedom only in the current 
version of the program. The program will 
be modified in the near future to provide 
compatibility for hub rotational degrees 
of freedom. 



84 



Analysis vs Test Results for the Model 347 
Helicopter 

The D-65 coupled rotor/airframe pro- 
gram was used to predict Model 347 flight 
vibration levels. Figure 8 shows the 
model used to predict airframe dynamic 
characteristics. Figure 9 compares pre- 
dicted vertical and lateral cockpit 
vibration levels vs vibration levels 
measured in flight. Vertical vibration 
levels are in reasonably good agreement 
at high airspeeds where vibration levels 
may become significant. Lateral vibra- 
tion levels are higher than predicted. 

Conclusio ns 

Methods have been developed indepen- 
dently for prediction of rotor vibratory 
hub loads and airframe dynamic character- 
istics. Methods are available for in- 
cluding effects of vibration control 
devices on airframe vibration and for 
optimizing the airframe structure. The 
substructure method is available for 
minimizing computer running time in 
analysis of airframe structures, and 
NASTRAN now provides a common finite 
element structural analysis program avail- 
able to all aerospace contractors. Rotor 
hub vibratory motions can modify rotor 
hub vibratory forces acting on the air- 
frame. A linear coupled rotor/airframe 
analysis method provides an approach for 
determing compatible hub motions and hub 
shaking forces . This method should be 
studied further to determine its 
validity. A method of this type should 
be considered in applications of NASTRAN 
for prediction of helicopter vibration; 
the user programming feature in NASTRAN 
should permit a coupled rotor/airframe 
solution of this type within NASTRAN. 

Figure 10 shows a scheme for solving 
for rotor trim, rotor forces with no hub 
motion, and the rotor impedance matrix 
using a rotor analysis program. NASTRAN 
would be programmed to use these mobili- 
ties and the rotor analysis results to 
solve for compatible rotor/airframe loads 
and motions. The NASTRAN airframe analy- 
sis could include airframe installed 
vibration control devices either in the 
initial airframe analysis or in the 
coupled rotor/airframe solution. Finally, 
results of these analyses could be used 
to determine optimum changes to the air- 
frame structural elements for minimizing 
airframe vibration. 



10. 



D.C., 1954. 

Boeing Vertol Company, D8-0614, 
AEROELAST1C ROTOR ANALYSIS, D-95, 
Thomas, E., and Tarzanin, F., 1967. 

Boeing Vertol Company, D210-10378-1, 
& -2, AEROELASTIC ROTOR ANALYSIS, 
C-60, Tarzanin F. J., Ranieri, J. 
(to be published) . 

Sciarra, J. J., DYNAMIC UNIFIED 
STRUCTURAL ANALYSIS METHOD USING 
STIFFNESS MATRICES, AIAA/ASME 7th 
Structures and Materials Conference, 
April 1966. 

The Boeing Company, D2-125179-5, THE 
ASTRA SYSTEM — ADVANCED STRUCTURAL 
ANALYSIS, Vol. 5, User's Manual. 

NASA SP-222 (01) , NASTRAN USER'S 
MANUAL, McCormick, Caleb W. , 
National Aeronautics and Space 
Administration, Washington, D.C., 
1972. 

Sciarra, J. J., and Ricks, R. G. , USE 
OF THE FINITE ELEMENT DAMPED FORCED 
RESPONSE STRAIN ENERGY DISTRIBUTION 
FOR VIBRATION REDUCTION, ARO-D 
Military Theme Review, Moffett Field, 
California, U.S. Army Research 
Office, September 1972. 

Sciarra, J. J., APPLICATION OF IMPE- 
DANCE METHODS TO HELICOPTER VIBRA- 
TION REDUCTION, Imperial College of 
Science and Technology, London, 
England, July 1973. 

Gerstenberger, W. , and Wood, E. R. , 
ANALYSIS OF HELICOPTER CHARACTERIS- 
TICS IN HIGH SPEED FLIGHT, American 
Institute of Aeronautics and Astro- 
nautics Journal, Vol. 1, No. 10, 
October 1963, pp 2366-2381. 

Novak, M. E ., ROTATING ELEMENTS IN 
THE DIRECT STIFFNESS METHOD OF DYNA- 
MIC ANALYSIS WITH EXTENSIONS TO 
COMPUTER GRAPHICS, 40th Symposium on 
Shock and Vibration, Hampton, 
Virginia, October 1969. 



References 

Leone , P . F . , THEORY OF ROTOR BLADE 
UNCOUPLED FLAP BENDING OF AERO- 
ELASTIC VIBRATIONS, 10th American 
Helicopter Society Forum, Washington, 



85 



ACTUAL BLADE 




APPROXIMATION 



BLADE SECTION 
BOUNDARIES 




EQUIVALENT SYSTEM 



CONSTANT EI 6 GJ 
ELASTIC BAY 



APPLIED 
AIRLOADS 



A 



\rrttYfrmtmroj j mw((rmtt^- n( tmli( mf ^--^-Y 



/ MASS BAY 

^EQUIVALENT MASS 



-BLADE SECTION 
BOUNDARIES 



-PITCH AXIS 



Figure 1. Rotor Blade Analytical Model 



• AIRFRAME INPUT DATA 

-NODAL COORDINATES AND CONSTRAINTS 
-STRUCTURAL ALEMENT PROPERTIES 
-MASS AND INERTIA PROPERTIES 


. 


- 


• STRUCTURAL ANALYSIS 

-FORM MEMBER STIFFNESS MATRICES 

AND ADD TO OBTAIN THE AIRFRAME GROSS 

STIFFNESS MATRIX 
-REDUCE NON-MASS DEGREES OF FREEDOM 

FROM GROSS STIFFNESS MATRIX 


' 




• COMPUTE AIRFRAME NATURAL MODES AND 
FREQUENCIES AND GENERALIZED MASSES 








• ROTOR VIBRATORY 
FORCES 






1 


i 


■ . . . .... 


• COMPUTE AIRFRAME FORCED RESPONSE 



Figure 2. Uncoupled Airframe Dynamic 
Analysis 



INPUT 



• NODE NUMBERS, CONSTRAINTS 
RETAINED, REDUCED DEGREES 
OF FREEDOM 

• STRUCTURAL PROPERTIES OF 
MEMBERS CONNECTING NODES 

• MASSES AND INERTIAS TO BE 
CONCENTRATED AT RETAINED 
DEGREES OF FREEDOM FOR 
MASS MATRIX 



ELEMENTS 

AXIAL 

1« «2 

BEAM 



SKIN 



^ 



GENERATE STIFFNESS MATRIX, K 

• GENERATE MEMBER STIFFNESS 
MATRICES AND ADD TO OBTAIN 
AIRFRAME GROSS STIFFNESS MATRIX 

• REDUCE GROSS STIFFNESS MATRIX 
TO RETAINED DEGREES OF FREEDOM 



N = 



K 11 K 12 



_ K 21 K 22 



~ 2000 X 2000 
1 



[Kll] = [Kii-K 12 K22~ 1 K2i]~200 X 200 

• LARGEST COMPUTER TIME ASSOCIATED 
WITH REDUCTION PROCESS 



T 



COMPUTE NATURAL MODES , ^ 
AND FREQUENCIES , u n r 



INPUT 



ROTOR FORCES, F R 



COMPUTE FORCED 
RESPONSE , {x} 



A n n n n 

T 



n n 



{ >nR} (Fr) -fn 

K} T [>] {♦„}" W n 

• COMPUTE RESPONSE 
OF EACH MODE; 
ADD TO OBTAIN 
TOTAL RESPONSE 

q n = MODE GENERALIZED 
COORDINATE 



Figure 2. Continued 



86 



AIRFRAME INPUT 
DATA, SUBSTRUCTURE i 



OPTION 1 



• GENERATE SUBSTRUCT. 
i GROSS STIFFNESS 
MATRIX 

• REDUCE ALL BUT MASS 
AND BOUNDARY DEGREES 
OF FREEDOM 



• MBRGE SUBSTRUCTURE REDUCED 
STIFFNESS MATRICES 

• REDUCE NON -MASS BOUNDARY 
DEGREES OF FREEDOM 



• COMPUTE AIRFRAME NATURAL 
MODES AND FREQUENCIES 



• ROTOR FORCES 



• COMPUTE AIRFRAME 
FORCED RESPONSE 





1 

. - i 




K AA | 




rn 


1 


_o-_J 


1 
1 


I . _ 

1 K BB- 


1 
1 


1 


r~+--~ 


1 


-4- J 




| K cc 
I 




Figure 3. 



BOUNDARY DEGREES 
OF FREEDOM 

Substructure Method for Gener- 
ating Airframe Reduced Stiff- 
ness Matrix 













S-74 








' 


DYNAMIC ANALYSIS 
NORMAL MODE 
METHOD 




INTERNAL LOADS 
TAPE 

F s = Kx s 
F c = KX C 








STRUCTURAL 
DATA 




S-74 

STRESS 

ANALYSIS 






OPTION 2 




ALL DEFLECTIONS 
OBTAINED 






» GENERATE ELEMENT 
STIFFNESS MATRICES 

• PICK UP ELEMENT 
END DEFLECTIONS 




NASTRAN 
NORMAL MODE 
ANALYSIS- 
CALCULATE 
ALL DEFLECTIONS 


















COMPUTE MAX 

X T KX FOR 

EACH ELEMENT 


_ 






ACCEPTABLE 
OPTIMIZATION 




* 


1 




STRAIN ENERGY 

SORT - CALCULATE 

WEIGHTS AND 

STRAIN DENSITY 

- SORT 






X£it> 




CRITERIA 

1. VIBRATION LEVEL 

- MIN. 

2. WEIGHT PENALTY 

- MIN. 

3. WITHIN ALLOWABLE. 
STRESSES 




I 






MODIFY STRUCTURE, 
RUBINS METHOD, RE- 
RUN 










' 


NO 



RESPONSE OF 
.ORIGINAL FUSELAGE - 
WITH RIGID BODY MOTION 




RESPONSE USING RUBIN 
METHOD FOR STRUCTURAL 
MODIFICATION-RIGID BODY 
MOTION INCLUDED 



Figure 4. Damped Forced Response Method 
for Airframe Optimization 



87 



MODEL 347 COCKPIT VERTICAL 
(NO ABSORBERS) 



to 



.5 r 



+1.4 - OBJECTIVE 



8 

H 
Eh 



H 

H 
U 

3 



FLIGHT DATA 



ANALYSIS 




40 80 120 160 
AIRSPEED - KNOTS 

MODEL 347 COCKPIT VERTICAL 
WITH ABSORBER 



O 
+ \ 

§ 

H 

w 

H 
U 

3 



OBJECTIVE 



ANALYSIS 



FLIGHT DATA. 



— . * * 5=gF 




40 80 
AIRSPEED 



120 
KNOTS 



160 



z H4s 



A ll A 12 



A 21 A 22 



! Z4s 



Z4c 



A 22 = A ll 



A 21 A 12 



A ll' A 12 
10"" 4 IN/LB 



.4 
.3 


*ll/ 


^Exciting 

Frequency 
i = 4fi 
< 
t 


.2 
2 


/ 

/ 


\ 
\ 
\ 
\ 


.1 

n 


/ 
/ 


/ v \^ 
1 ^ 




i 4 £ 


.1 


A 12 j 


2 


.2 


>v / K 


.3 







(b) Airframe -Hub Mobilities 



Figure 5. Predicted Vs. Measured 

Cockpit vibration Reduction 
with a Vertical Cockpit 
Absorber 



iy a = 44.5 RAD/SEC 

T F , 2 
1 Zk Eh 




(a) Single Rotor Helicopter Vertical 
Vibration 



F Z4 " F Z4c cos 4S2t + F Z4s sin 



4f2t 



Z H4 = Z H4c COS 4flt + Z H4s sin 4fit 




-200. 



200. 



H4S, .001 IN. 



H4S, .001 IN. 



(c) Hub Forces Vs. Hub Motion 



Figure 6. Coupled Rotor /Airframe Analysis 
for a Single Rotor Helicopter 
Vertical Vibration 



Figure 6. Continued 



88 



Z = (Z 2 + Z 2 )H 2 





.0010 


j 


EXCITING -i 


.08 








I frequency! 








j/ 


.06 




z 


.0005 






.04 


Z 

4 


IN. 









.02 



G's 



F 
Z4 

LB. 



(d) Vibratory Hub Motion 



F Z4 = ( F l 4 c + Fg.s) 



II 2 




(e) Vibratory Hub Force 



A- 9 7 TRIM 
ANALYSIS 



D-88 INPUT 
FORWARD AND AFT 
ROTORS 



PARAMETRIC STUDY 
READ D-88 AND D-96 
PARAMETER CHANGES 



i~_ 



BASIC D-88 
AEROELASTIC ROTOR ANALYSIS 



Vibratory Hub Loads 
x' My' 



F x' ?y *"z' 



M. 



ML 



MATRIC 
CAPABILITY 



VIBRATION 
DEVICES 



D-96 INPUT 
(Fuselage) 

1. Masses 

2. Natural Frequencies 

3. Degrees of Freedom 




BASIC D-96 
DAMPED, FORCED RESPONSE OF 
A COMPLEX STRUCTURE 



OUTPUT FOR EACH 
DEGREE OF FREEDOM 

1. Displacement 

2. Phase Angle 

3 . G-Levels 



Figure 6. Continued 



Figure 7. D-65 Coupled Rotor/Airframe 
Program Flow Diagram 



89 



. STRUCTURAL IDEALIZATION 

HAS:" 1061 STRINGERS 
' 1089 SKIN ELEMENTS 

38' BEAMS 

SZ1 NODES (STRUCTURAL! , 
' 1849 DEGREES OF FREEDOM 
51 MASS NODES 
139 RETAINED O.O.F. 




Figure 8. Model 347 Airframe Dynamic Model 



.5 

/4 



STA. 95 C/L VERTICAL 
4/REV VIBRATION 
-NO ABSORBERS 




.5 
.4 

.3 

G's 

.2 

.1 




"40 60 80 100 120 140 160 180 

AIRSPEED, KNOTS 



STA. 95 C/L LATERAL 
4/REV VIBRATION 
-NO ABSORBERS 



FLIGHT TEST 




4o d rt ioo ria i4o iso i 



80 



AIRSPEED, KNOTS 

Figure 9. Model 347 Flight Data Vs. D-65 
Coupled Rotor/Airf rame Analysis 
Results 



• TRIM ANALYSIS 



• VIBRATORY ROTOR LOADS , 
NO HUB MOTION {f r0 }, 

• VIBRATORY ROTOR LOADS 
WITH UNIT VIBRATORY 
HUB MOTIONS 



• ROTOR IMPEDANCE 
MATRIX, B 



, AIRFRAME SUBSTRUCTURE ANALYSIS 
(INCLUDE MODELS OF VIBRATION 
CONTROL DEVICES) 

• MERGE; SUBSTRUCTURE STIFFNESS 
MATRICES 

• COMPUTE" -AIRFRAME MODES, FREQUENCIES, 
AND GENERALIZED MASSES WITH NO BLADE 
MASS AT ROTOR HUBS 



•COMPUTE AIRFRAME RESPONSE TO UNIT 
VIBRATORY HUB LOADS: 



*a 




A RR 
. A AR. 




F R 



• COMPUTE COMPATIBLE VIBRATORY HUB 
MOTIONS AND FORCES 

{X R } -[[l]-[A RR ] [B]]~ [A RR ] {f ro ] 

compute total hub forces 
{fr} = {fro} + [b]{x r ] 

•compute motions at other airframe 
degrees of freedom 



{x A } - [a ar ] (f r ) 

r 



•IDENTIFY STRUCTURAL CHANGES 
TO MINIMIZE AIRFRAME 
VIBRATION USING STRAIN 
ENERGY METHODS 



Figure 10. Coupled Rotor/Airf rame/NASTRAN 
Analysis 



90 



HELICOPTER GUST RESPONSE CHARACTERISTICS 
INCLUDING UNSTEADY AERODYNAMIC STALL EFFECTS 

Peter J. Arcidiacono 

Chief Dynamics 

Sikorsky Aircraft Division of United Aircraft Corporation 

Stratford, Connecticut 

Russell R. Bergquist 
Senior Dynamics Engineer 
Sikorsky Aircraft Division of United Aircraft Corporation 
Stratford, Connecticut 

W. T. Alexander, Jr. 

Aerospace Engineer 
U. S. Army Air Mobility Research and Development Laboratory 

Eustis Directorate 
Fort Eustis, Virginia 



Abstract 

The results of an analytical study to eval- 
uate the general response characteristics of a 
helicopter subjected to various types of discrete 
gust encounters are presented. The analysis em- 
ployed was a nonlinear coupled, multi-blade rotor- 
fuselage analysis including the effects of blade 
flexibility and unsteady aerodynamic stall. Only 
the controls-fixed response of the basic aircraft 
without any aircraft stability augmentation was 
considered. A discussion of the basic differences 
between gust sensitivity of fixed and rotary wing 
aircraft is presented. The effects of several 
rotor configuration and aircraft operating param- 
eters on initial gust-induced load factor and 
blade vibratory stress and pushrod loads are dis- 
cussed. The results are used to assess the accu- 
racy of the gust alleviation factor given by MIL- 
S-8698. Finally, a brief assessment of the rela- 
tive importance of possible assumptions in gust 
response analyses is made and a brief comparison 
of gust and maneuver load experiences in Southeast 
Asia is presented. 

The results confirm that current gust alle- 
viation factors are too conservative and that the 
inclusion of unsteady stall effects result in 
higher initial load factors than predicted using a 
steady stall aerodynamic analysis. 

Notation 

An gust alleviation factor; see Equation (l) 



l * 



and (3) 



ft. two-dimensional lift curve slope of rotor 
blade section 

b number of blades 

B tip loss factor 

XL blade chord, ft 

C' T vertical force coefficient, Thrust/jjpn. t R i *' 

GW gross weight, lbs 

Presented at the AHS/NASA-Ames Specialists' Meet- 
ing on Rot or craft Dynamics, February 13-15, 197^. 



Xg blade mass moment of inertia about flapping 
hinge, slug - ft 2 

R blade radius , ft 

S fixed wing area, ft 2 

"t a , partial derivative ^ ^ e ' 

V forward velocity, knots or ft /sec 

V,Lta average characteristic velocity for helicop- 
ter rotor 

Vbl maximum vertical velocity of gust, positive 
up, ft/sec 

(X angle between shaft and relative wind, posi- 
tive tilted aft, radians 



ft . 
)f blade lock number, f <*-<cft/X 



B 

. ,- , MA* THMST ■ 

^ n incremental rotor load factor ; — — ■ B — • 

j&fij incremental rotor load factor predicted by 
linear steady theory for sharp edge gust 
instantaneously applied to entire lifting 
device . 

A inflow ratio, ( VSin* - ^)/xiK 

ytx advance ratio, V c-* 4 -'* /siH. 

1/ rotor induced velocity, positive up, ft /sec 

f air density, slugs /cubic foot 

<T rotor solidity, ^"^/n R 

XL rotor angular rotational velocity, radians/ 
second 



id 



three dimensional lift curve slope for fixed 



wing 
Subscripts 
P" denotes fixed wing 
" denotes helicopter 



91 



< 
o 



M .0.5 - 




2 .4-6 8 10. 

.'.'- DISC LOADING, LB/FT 2 

Figure 1. Current Gust Alleviation Factor. 

Current procedures for predicting helicopter 
gust-induced loads involve computing rotor loads 
by means of a simplified linear theory and modify- 
ing these loads by a gust alleviation factor de- 
fined in Specification MIL-S-8698 (AEG). The al- 
leviation factor is shown in Figure 1 and is a 
function of rotor disc loading alone. Further, no 
alleviation is allowed for disc loadings greater 
than 6.0 - a value exceeded by many modern heli- 
copters. Attempts to verify the accuracy of this 
approach through flight test have been complicated 
by uncertainties regarding the gust profiles. 
This has led to side.-by-slde flight tests of fixed 
and rotary-wing aircraft (Reference l) in order to 
build a semi-empirical bridge between the rela- 
tively straight forward fixed wing situation and 
the more complex situation associated with rotary 
wings. Limited qualitative results on aircraft of 
comparable gross weight indicated that the heli- 
copter was less gust sensitive than the fixed wing 
aircraft, but extensive quantitative data from 
this type of test are, obviously expensive and 
difficult to obtain. Analytical confirmation of 
the MIL-S-8698 (ARG) gust alleviation factor has 
been hampered by the lack of an analysis which can 
handle both the transient response of the helicop- 
ter and the aeroelastic response of the rotor 
blades, while, simultaneously, providing a reason- 
ably complete modeling of the rotor aerodynamic 
environment. An improved gust response analysis 
(described in Reference 2) has indicated that cur- 
rent procedures are too conservative. The primary 
objectives of this investigation were (l) to 
develop a similar computerized analysis based on 
the rotor aeroelastic and unsteady stall aerody- 
namic techniques developed at Sikorsky Aircraft 
and the United Aircraft Research Laboratories and 
(2) to apply the analysis to predict rotor gust 
alleviation factors for comparison with those 
given in Specification MIL-S-8698 (ARG) and in 
Reference 2. The principal contribution of this 
analysis relative to that of Reference 2 is the 
inclusion unsteady stall aerodynamics. The re- 
sulting computer program was designed to function 
on the CDC 6600 computer and is catalogued at both 
the Langley Research Center and the Eustis 
Directorate. 



Comparison of Helicopter and Fixed 
Wing Gust Response 

Before proceeding with a detailed analysis 
of the helicopter gust response characteristics, 
it is instructive to compare fixed wing and heli- 
copter characteristics in relatively simple terms. 
Such a comparison follows. 

In analyzing the response of fixed wing air- 
craft to discrete sharp edge gusts, (eg. Reference 
3) the concept of a gust alleviation factor is em- 
ployed. The gust alleviation factor is simply the 
ratio of the "actual" incremental load factor pro- 
duced by the gust to the incremental load factor 
computed from simple steady-state theory. The 
"actual" load factor may represent a measurement 
or may be computed from some more rigorous theory 
applicable to the unsteady gust encounter situa- . 
tion. Thus, if Ag is defined as the. gust allevia- 
tion factor, we have 






for fixed wing aircraft we have 



(l) 



(2) 



Following the same approach for a helicopter 
having a rotor as its sole lifting element, we can 
write: 



(a*) h = (*» s ) H (A<j) H 



(3) 



Using steady, linear rotor theory results from 
Reference k, and assuming a sharp edge gust in- 
stantaneously applied to the entire rotor, 
is given by: 






(a» s ) h . - kf C^iMM* 



&W/bcR 



k« ^V ft -^ 



6^/bX!R 



* 



(k) 



(5) 



(6) 



92 



Hence, the actual load factor is given by 



'\ 



^/ bA R 



(7) 



Equation (7) is of similar form to the correspond- 
ing fixed wing equation (Equation 2). Further, by 
comparing the two equations, it is clear that the 
characteristic or average velocity for the rotor 
is given by tj , XJ-R and that the characteristic 
area for the rotor is the total blade area. The 
characteristic velocity V<wj, of the rotor is pre- 
sented in Figure 2. A typical value of Va.^ is 
about 0.5ilR and the effect of advance ratio (or 
forward speed) is seen to be small. This con- 
trasts with the fixed wing case where the charac- 
teristic velocity is equal to the aircraft's for- 
ward velocity. 





0.8 


7 = 15,B=0.97 ; 


o a 






lE;f 






O o 

fcF 


0.6 


■ __ 


m 5 




— - — "" 


tr. or 




^ 


0.4 




o t- 






<u 






£s 


0? 




X UJ 






o > 








0.2 0.4 
ADVANCE RATIO, /J- 



0.6 



Figure 2. Rotor Characteristic Velocity Ratio. 

Now, attempts have been made to measure the 
gust alleviation factors of helicopters through 
side-by-side flights with fixed wing aircraft. 
However, the relative alleviation factors so de- 
termined are only meaningful if the Gust Response 
Parameter for the two aircraft are equal. This 
equality of Gust Response Parameters is shown in 
Equation 8: 






(8) 



6-W/bxiR 



If the relation above is satisfied, Equations (2) 
and (T) indicate that the following equality also 
holds : 






(9) 



Assuming the two aircraft encounter the same gust 
velocity profile, Equation (9) reduces to 






do) 



Thus, the ratio of the gust alleviation factors 
will be in proportion to the measured load factors 



for the two aircraft . If the fixed wing gust al- 
leviation factor is known, (Ag)g can then be de- 
termined. • '• 

If Equation (8) is not satisfied, then the ' 
gust alleviation factor for the helicopter can be 
determined from the following relation: 



M^^ri v § 



(A*), 



%'w 




IS Aw \\ 



*"L*> 






.(ID- 



Typical values of the Gust Response Parameters of 
Equation (8) are presented in Figure 3. The re- 
sults of Figure 3 indicate that a helicopter having 
a blade loading of 100 lb /ft 2 and operating at a 
forward speed of 250 fps will exhibit approximately 
the same sensitivity to a gust as a fixed wing air- 
craft having a wing loading of about 60 lb /ft 2 , 
provided, of course, that the gust alleviation fac- 
tors for both aircraft are equal. In practice, 
for this example, the gust alleviation factor for 
the fixed wing will be significantly higher (mean- 
ing higher acceleration) than that for the helicop- 
ter. 



SEA LEVEL 

FIXED WING 6C L /3a =4.5) 

HELICOPTER (a = 5.73, JIR = 700 FPS) 




100 200 300 

FORWARD VELOCITY, V- FT/SEC 



Figure 3. 



Fixed Wing and Helicopter Gust 
Response Parameters . 



Factors Influencing Helicopter Gust Response 

The computation of gust induced loads for 
helicopters is a difficult analytical task because 
the rotary wing lifting system is a complex aero- 
elastic mechanism operating in complicated aerody- 
namic environment. Principal factors which can be 
expected to influence the gust response of a heli- 
copter are described briefly below. 

a. Rotor blade response - Helicopter rotors 

differ from fixed wings in that the blades 
(wings) of the rotor are relatively flexible 
and, in many cases, are articulated relative 
to the fuselage. The blades, therefore, are 
much more responsive to gust loads than is 
the aircraft as a whole and react in such a" 
way as to reduce or isolate (at least tem- 
porarily) the fuselage from the impact of 
the gust. Thus, while the blades axe re- 
sponding to the gust, the fuselage has time 
to build up vertical velocity which, in 
turn, redUces the effective velocity seen by 
the rotor. A simple example illustrating 



93 



the magnitude of the various forces con- 
tributing to the fuselage acceleration is 
shown in Figure k for a sharp edge gust 
applied instantaneously to a rotor having 
nonelastic flapping blades and operating in- 
hover. In this extreme case, because of the 
overshoot of the blade flapping response, 
the peak acceleration experienced by the 
body is about the same as it would have been 
had the blades been completely rigid (i.e. 
equal to the acceleration given by the gust 
term alone). As seen in Figure h, the 
forces associated with the blade dynamic re- 
sponse are large; hence any factor influenc- 
ing the blade is potentially important. 



r.io HOVER 

SHARP-EDGE GUST, ZERO TIME PENETRATION 
NONELASTIC BLADES 



GUST TERM 




100 200 300 

BLADE AZIMUTH, DEG 



Figure h. 



Comparison of Terms Contributing to 
Fuselage Acceleration. 



Fixed wing response - If the helicopter is 
fitted with fixed wings (compound configura- 
tion), additional gust loads are, of course, 
generated. These can be treated using 
available fixedr-wing techniques and are not 
of primary concern in this study. 



be phased so that the peak loads for each 
blade occur at different times (see Figure 5)> 
As a result, finite-time penetration of the 
gust reduces the peak fuselage accelerations 
produced by a given gust profile. 



SHARP-EDGE GUST 
FINITE TIME PENETRATION 
NONELASTIC BLADES 




BLADE AZIMUTH, DEG 

Figure 5. Finite-Time Penetration Causes Peak 

Blade Forces to be out of Phase. 
Control system inputs - The ultimate effect 

of gust on the helicopter must be influenced 
by any reaction of the pilot or stability 
augmentation system to the initial loads pro- 
duced by the gust. It is possible (but un- 
likely with a properly designed system) that 
the largest loads produced by the gust will 
not be the initial loads but, rather, those 
associated with the longer term response of 
the coupled system represented by the air- 
craft, pilot, and stability augmentation sys- 
tem (See Schematic in Figure 6). These longer 
term effects depend on the specific design 
characteristics of the aircraft system and no 
attempt was made to model them in the present 
study. Hence, the gust-induced loads con- 
sidered are the initial loads caused by the 
gust for a controls-fixed rotor operating con- 
dition. 



Rotor Aerodynamic Modeling - The ability of 
a rotor to generate load factor during a 
gust encounter will depend on the proximity 
of the blade trim angle of attack distribu- 
tion to stall. A rotor operating on the 
verge of stall prior to a gust encounter can 
be expected to generate less additional lift 
due to the gust than can a rotor initially 
operating further away from stall. The 
modeling of stall aerodynamics is important; 
therefore, the impact of unsteady aerody- 
namics on rotor stall was investigated in 
this study. 

Gust Characteristics - Gust profile and am- 
plitude are, of course, potentially impor- 
tant factors. In addition, the speed of the 
helicopter as it penetrates a given gust 
front can be expected to be significant. 
Figure k indicated the fuselage acceleration 
for a gust applied instantaneously to the 
entire disc. With a finite-time penetration 
of the gust front, the contribution of each 
blade to the fuselage loading will not be 
identical (as in Figure It) but rather will 



SECONDARY LOAD PEAK 
RESULTING FROM INTERACTION 
OF GUST AND AIRCRAFT- 
PILOT- SAS SYSTEM^ 



INITIAL MAXIMUM 

LOAD PRODUCED 

BY GUST (CONTROLS -FIXED) 




Figure 6. 



Schematic of Possible 
Load Factor Time Histories. 



94 



Brief. Description of the Analysis' 



.Simple,' Linear Gust theory 



Complete documentation of the equations used 
in the analysis is given in Reference 5, while 
procedures for running the associated computer 
program may be found in Reference 6. Both of 
these references can he obtained from the Eustis 
Directorate of USAAMRDL. 

Briefly, the analysis is essentially a digi- 
tal flight simulator that can be used to determine 
the fully coupled rotor - airframe response of a 
helicopter in free flight. This is accomplished 
by the numerical integration of the blade - air- 
frame equations of motion on a digital computer. 
The principal technical assumptions and features 
of the analysis are listed below. 

1. The blade elastic response is determined 
using a modal approach based on the equa- 
tions defined in References 7 and 8. The 
number of modes used consisted of three 
flatwise,, two ehordwise and one torsion for 
each .blade. 

2. The .aerodynamic modeling of the blade in- 
cludes unsteady aerodynamic effects based on 
the. equations ■ and tabulations defined in 

'.Reference 8 which assume that the lift and 
. moment coefficients can be expressed as 
as functions of instantaneous angle of at- 
tack and its first two time derivatives. 
Steady-state drag was used, however, because 

■ of a lack of data on unsteady drag in stall . 

3. Rotor inflow is assumed constant for this 
study although provision for time-varying 
induces velocities is available. The con- 
stant value is determined from classical 

' momentum theory and was invariant with 
either position on the disc or with time. 

■ In view of the short times required for 
peak loads to be achieved, this assumption 
is considered reasonable. 

h. The response of each individual blade is 
considered. 

5. The fuselage is a rigid (nonelastic) body 
having six degrees of freedom. Provisions 
for fixed wings are included. The aerody- 
namic .forces on the wings are computed using 
simple, -.finite-span wing theory, neglecting 
stall and unsteady effects. 

6. Fuselage aerodynamic forces and moments are 
determined using steady-state nonlinear, 
empirical data. 

7. The gust is assumed to be both two dimen- 
sional (i.e. does not vary along the lateral 
axis of the rotor) and deterministic in 
nature. • Although three dimensional and ran- 
dom gust effects may prove important, their 
inclusion was beyond the scope of this 
study. 



, As stated earlier, it was des'ired to cast the 
results obtained in this investigation in terms' of , 
correction factors (gust alleviation ' factors ) that 
could be applied to results dbtained . f rom a simple, 
specification eventually evolved. 

The simple "theory used is that defined in '- . 
Reference k, in which blai# stall and'compressi-' 
bility effects are; neglected. In -addition, it is • 
assumed here that the' gust is" sharp'^edged and is 
instantaneously applied' 'to -the, entire rotor. The, 
increment in rotor load f actpr prtiduced by the gust' 
is then given by Equation (5) . ' Using the relation, 



ew «f^r i$y 



(1?) 



the incremental rotor, load factor given by simple 
theory is 



The ratio of the ( A*\ ) H . computed by the more 
complete analysis described herein to the value 
given by Equation (13) represents a gust allevia- 
tion factor which can be used to correct the load 
factors results given by Equation (13). Thus: 






(i>0 



Values of Ag presented in this paper are based on a 
rotor blade lift curve slope, a, of 5.73. Hope- 
fully, if Ag shows reasonably consistent trends, it 
can be used with some confidence to rapidly predict 
rotor load factors for combinations of parameters 
other than those considered in this study. 

Scope of Study 

Gust load factors , blade bending moments , vi- 
bratory hub loads, and rotor control loads were 
calculated for a range of values for rotor thrust 
coefficient- solidity ratio, blade Lock number, ad- 
vance ratio, and blade flatwise and torsional 
stiffness. The effect of adding a wing was also 
investigated. The responses associated with three 
types of vertical gusts were investigated: sine- 
squared, ramp, and sharp-edged. The sine^squared 
gust and the ramp gust reached a maximum value of 
fifty feet per second at a penetration distance of 
ninety feet. The gust profiles are displaved in 
Figure 7- Three types of rotor systems were eval- 
uated: articulated, nonarticulated (hinse3.ess) , 
and gimbaled. Emphasis in this paper is placed on 
the results for the reference articulated rotor. 
The reader is referred to Reference 10 for details 
of the other configurations studied. The articu- 
lated rotor properties can be found in Table I, 
together with the natural frequencies of the 
blades. As indicated, the number of modes used 



95 



consisted of three flatwise, two ehordwise, and 
one torsional modes. 







SIN 


1 ■ 

! -SQUARED OUST 








































40 































*s 


/ 


















1 
























r 









1 

RA1 


1 

P 61 


1 

JST 












































40 

























































































f 

















#»= 


0.5,REFER£NC£ ARTICULATED ROTOR, 


:,/. 


..06 




































__ — =— — Sf HE -SQUARED CU 
SHARP-EDGED GUS1 




SUIT 
























































- 


























A 




r 




















:- 






| 24 000 










!\ 


A 


r 


•"' 












K 


H 


[/ 












,■!■ 








„' 


r 


^ 


■' 


vy 


ft 






















1 








// 


•& 




\ 






/ 


\ 


V'-.N 




\ ,' 


":»\ /; 


1 6 ODO 
































£ 


vw 


~v 17 






r 






























\ 


















==. 


'/ 


























\ 




\ 


£ 




^K 


? 
































V 


i 


I 8 000 












































l. 














































1 











































































































































1 






160 




320 
ROTOR BLADE 


480 ft 
*IAZiMUTf 

1 


640 
, DEC 






800 


9C 



s 


HARI 


1 1 

-ED! 


ED I 


UST 












































































































ft 
























,» 





40 80 120 160 200 

PENETRATION DISTANCE , FT 

Figure T. Gust Profiles. 



Discussion of Results 

Effect of Oust Profile 

The effect of gust profile on incremental 
rotor thrust force for the reference articulated 
rotor was evaluated by the penetration of three 
gusts with profiles as shown in Figure 7. The 
helicopter was assumed to penetrate a stationary 
gust with a velocity of 350 feet per second. This 
corresponds to an advance ratio of 0.5. 

The time history of the rotor thrust asso- 
ciated with each of the gust profiles is shown in 
Figure 8. It may he seen from this figure that 
while the actual gust wave form has little impact 
on maximum rotor force and consequently on rotor 
load factor, the particular time histories behave 
differently, although in an expected manner. Ini- 
tially, the sine-squared and the ramp gust shapes 
result in similar peak rotor loads at approximate- 
ly the same time. The sharp-edge gust induces a 
greater peak load with a faster build up. As the 
penetration distance increases, the loads produced 
by the sharp-edge gust and the ramp gust tend to 
merge since their respective velocities are hoth 
50 fps while the value of the sine-squared gust 
velocity has dropped back towards zero. 

Analysis of the computed results forming the 
basis for Figure 8 indicates that at the time the 
maximum rotor vertical force and load factor is 
reached, the helicopter fuselage has had time to 
develop only a modest amount of vertical velocity. 
The vertical velocities at the peak load points of 
the helicopter associated with the sine-squared 
and sharp-edged gusts are 6fps, and 3 fps,3 fps, 
respectively. These compare to the '50 fps gust 
velocity, indicating that little gust alleviation 



GUST FRONT IS GUST FRONT REACHES MAXIMUM GUST VELOCITY 

TANGENT TO HELICOPTER CENTER REACHES HELICOPTER 

ROTOR DISC OF GRAVITY CENTER OF GRAVITY 

Figure 8. Rotor Force Time Histories. 

is being produced by fuselage motion for the con- 
dition analyzed. 

The three types of gust profiles evaluated did 
not produce greatly different peak rotor loads. 
While the sharp-edge gust does produce the largest 
loads.it is probably the least realistic of the 
three profiles. Since other studies, such as 
Reference 2, have used a sine-squared gust, the 
remainder of the results presented are based on 
this profile. 

Effect of Rotor and Flight Condition Variables 

The variation of gust alleviation factor, (as 
computed from Equation .1*0 , with initial rotor 
loading is shown in Figure 9 for the three types 

/x= 0.002378 ftR=700FPS o-=0.085 SINE 2 GUST 





1.2 











NOH 


ART 


CUL 


ITEt 


' ROTOR 
























itlL- 


S-8( 


98 




















































CUR 


REN 


' ST 


UDY 











































AR' 


" 

ICUl 


ATE 


D R 


3 TOR 
























«IL- 


>-8i 


98 


























































CUR 


SEN 


ST 


JDY 



































.04 .06 .08 .10 

ROTOR THRUST COEFFICIENT/ SOLIDITY, Ct/o- 

Figure 9. Gust. Alleviation Factors for 
Different Rotors. 



96 



types of rotors analyzed. Rotor iloading in this 
figure has been expressed t>oth in terms of rotor 
thrust coefficient solidity ratio and rotor disc 
loading. It should he remembered that disc load- 
. ing is not a unique function of Op/8" hut rather 
depends on the value of density, tip speed and 
solidity of the rotor. Values for these quanti- 
ties are noted on the figure. 

The results of Figure 9 indicate that in- 
creasing Oj^r leads to a large reduction of gust 
alleviation factor, This is similar to the trend 
noted in Reference 2 and is believed to be re- 
lated to the loss in average additional lift 
capability at the higher Op/a* due to the occur- 
rence of stall. The influence of rotor configura- 
tion is seen to be of rather secondary importance. 
Rotor configuration would be expected to influ- 
ence fuselage motion through the transmittal of 
differing rotor pitching moments to the airframe, 
depending on the degree of rotor articulation. 
The relative insensitivity of the results to con- 
figuration is believed to be due to the short 
time in which the initial, controls-fixed load 
factor is generated. As a result, the fuselage 
response to the differing moments is not large 
and the load factor tends to be dominated by the 
rotor blade dynamic response, which is roughly 
the same . for all rotors . This result is also 
similar to that observed in Reference 2. 

It is also evident from Figure 9 that the 
gust alleviation factors defined in MIL-S-8698 
(ARG) are too high (i.e. result in loads which 
are too high) . The conservatism of the current 
specification is particularly evident at the high 
thrust coefficient-solidity ratios where rotor 
stall becomes a factor limiting gust-induced 
thrust generating capability. On this basis, one 
would expect the gust alleviation factor for up- 
ward gusts to be different from those for down- 
ward gusts (i.e. gusts which unload the rotor). 
While downgusts are not critical from a structural 
loads viewpoint, they could prove more important 
from a passenger - comfort point of view. 

The results of Figure 9 are for typical 
reference rotor configurations (see Reference 10 
and Table I herein) . As part of this study, cal- 
culations were made to examine the sensitivity of 
the computed gust alleviation factors to separate 
variations in blade Lock Mumber, bending stiff- 
ness and torsional stiffness. Ranges of the 
parameters considered are noted below: 



ratios and higher Lock numbers being associated 
with the lower gust alleviation factors. The rela- 
tively small effect of blade stiffness variations 
is perhaps not surprising inasmuch as the total 
blade stiffness tends to be dominated by centrifu- 
gal stiffening effects. The variations shown for 
Oj/t of 0.06 are believed to be representative of 
those at other Of/tr values; however, this should be 
verified. 

Articulated rotor load factors predicted using 
the results of Figure 9 are presented in Figure 10 
where they are also compared to the results of 
Reference 2. Load factors predicted by the current 
study are seen to be higher than those of Reference 
2. This increase appears to be due to the use of 
unsteady aerodynamics in the current study. 









articulated' rotor 
sine^gust 


































^ 
























■-. 


-^^ 


*** 




^_ 






















~~ 


3-1. 


























































CURRENT STUDY (UNSTEADY STALL AERO) 







CURRENT STUDY{STEADY STALL AERO 

L L_^_J ! J J 









ROTOR THRUST COEFFICIENT/SOLIDITY, C T /«• 



DISC LOADING, LB/FT 1 



Figure 10. Comparison of Results with Reference 2 

Gust-Induced Blade Stresses, Control Loads and 
Vibration 

The effects of a gust encounter on other 
quantities of interest to the designer such as 
blade stresses, control loads, and aircraft vibra- 
tion were briefly examined. In examining these 
effects, an attempt was made to generalize the re- 
sults to a limited degree by relating the maximum 
values produced by the gust to the initial trim 
values. Results are based on the trim condition of 
Cy/<r = 0.06 and an advance ratio of 0.3 are pre- 
sented in Figure 11. Detailed analysis of the 
trends shown were beyond the scope of this paper. 
The reader is referred to Reference 2 for a more 
detailed discussion. 



Lock Humber: reference, 0.7 ref. 1.3 ref. 

Bending stiffness: reference, 0.5 ref. 

Torsional stiffness: reference, 0.5 ref. 

The parameter variations listed above were made 
at an advance ratio of 0.3 and 0.5 for a Oj/a- of 
0.06. The range of results is also shown in 
Figure 9 and as can be seen, the effect of blade 
Lock number and stiffness is relatively small. 
Lock number and advance ratio account for most of 
the small variation shown, with the lower advance 



97 



.' 40 

30 

-w 20 

■' q: 

"- 

o 

w 10 



§ 6 

H 

in 4 H 
o 

i 2 

X 

I ° 

u_ 

° 3.0 

o 

2 2.0 



■ 4P 


VERTICAL 


VIBRATORY' FORCE 


- '. 


i 


•^ y= 7 



Reference 2 and previously presented in Figure 10 . 



vibratory control load 
St 



10 

y=i3^- 



4 



i.o - 





VIBRATORY FLATWISE MOMENT 



7= 7, 10,13 



0.2 



0.3 0.4 

ADVANCE RATIO 



0.5 



Figure 11. Effect of Gust on Vibration, 

Control Load and Flatwise Moment, 
C T = 0.06, 50 fps SINE 2 Gust 

Sensitivity of Results to Assumptions 

As discussed in an earlier section of this 
report, many factors could potentially influence .' 
rotor gust response characteristics. To account . 
for all of these factors leads to a time consum- 
ing, complex, digital analysis. In the following '• 
paragraphs, the results of a brief examination of 
the importance of some of these factors are dis- 
cussed. Only the reference articulated rotor at 
one operating condition is considered. Any con- 
clusions drawn from these results must, therefore, 
be considered preliminary and should be substanti- 
ated by further investigation. A summary of the 
results obtained is presented in F'igure 12. Shown 
is the percentage change in the predicted gust al- 
leviation factor resulting, from the separate elim- ' 
ination of fuselage motion, blade elastic torsion^ 
finite time penetration j and unsteady stall ef- 
fects in the analysis. The baseline value cor- ' ■'. 
responds to value for' the complete' analysis. A -. 
positive change in Ag means that the effect 
eliminated causes an increase iri predicted load- .. 
ing. It is evident that the- unsteady stall aero-' 
dynamic and finite-time gust penetration effects 
are most important* Excluding- unsteady aerody- 
namics reduces the predicted value of Ag by about 
29%.: This; is because the maximum lift capability 
of the rotor base"d on' steady aerodynamic "stall, . 

'characteristic's is lower than that "based on un- . 
steady- characteristic's (see Reference 9). The re-! 

■ duction is consistent with the observed lower' ' 
values 'of -predicted -load factors obtained in 



70 1— 
60 - 
50 - 
40 - 
30 - 
20 - 

10 ■ 


-10 
-20 



-30 



TRIM C T Aj- =0.06 

50 FPS GUST AMPLITUDE 

M-0.5 

ARTICULATED ROTOR 



. NO FINITE 
TIME 
PENETRATION 
. (SHARP-EDGE) 



FUSELAGE t&HJIu 

MOTION , TORSION 

(SINESgUST) (SINEgQUST) 




Figure 12. 



Sensitivity of Gust Alleviation to 
Analytical Assumptions. 



The largest change in predicted gust allevi- 
ation factor was produced by the elimination of 
the finite time penetration of the gust front. As 
might be expected, when the gust is assumed to af- 
fect all blades simultaneously, the blade forces 
are all in phase and large values of Ag (and hence 
loading) result. 

It should be emphasized that the results pre- 
sented in Figure 12 were determined for only one 
reference trim condition. Further worK is required 
to substantiate the generality of the- results. 

Gust Load Factor Experience .in SEA 

The earlier portions of this .paper have been 
• devoted to analytical techniques, appropriate for 

determining- the effects of gust encounters on heli- 
1 copter response variables . One point' which has 
' been made is that rotor blades , because of their 
•f.lexib."ility,. tend to reduce the impact of the gust 
on the fuselage. Resulting gust alleviation fac- 
tors have been found to be low and, hence, one 
.would expect that gust-induced loads on the fuse- - 
lage could be' reduced in importance. Experimental 
evidence supporting this contention has been ac- 
•qulred .by the U. S. Army in SEA. A brief discus- 
sion' of .that 'data is presented below. 

' • ' The U. S. Army has been acquiring usage data; 
on ' its combat operational helicopters' in Vietnam . 

," since early 1966. Beginning with both, the cargo 
arid armored versions of the CH-1*7A, the CH-5l*A, 
Aft-IG; ■ nd.UH-6A helicopters were instrumented to' 
rec6r'd .the history of their actual combat Usage. 

-• Since .control positions and c;g. accelera- 
tions were among the parameters 'measured 'and the '■' 
data were, recorded in analog format, .occurrences • 

, of gus'trinduced loads' were- identified- and isolated . 

.'from pilot-induced (maneuver) accelerations by 

•analyzing those particular trace recordings. Gust- 
induced acceleration, peaks, therefore, were-identi- 

. f ied as" those, accelerations occurring when both 



n 



the cyclic and collective stick traces were steady 
or, if stick activity was present, the sense of 
the peaking acceleration had to be in opposition 
to that expected from the stick control motion. 

A total of lVf7 hours of flight data were 
acquired during the measurement programs for the 
cited aircraft (References 11-13). The conclu- 
sive finding in each of these programs was that 
normal loads attributed to gust encounters were 
of much lesser magnitude and frequency than 
maneuver loads. Further, when the total load 
factor experience was statistically examined for 
each aircraft, the loads directly attributed to 
gust encounters were found to be only a small 
percentage of the total experience. These points 
are graphically illustrated in Figure 13. The 
maneuver load scatter band was obtained from 
References Ik and 15. 

It should be pointed out that while gust- 
induced load factors are smaller than typical 
maneuver load factors for military aircraft, gust 
loadings can be an important consideration from a 
ride comfort standpoint in commercial appliea- 
tions . 



10,000 



SUST INDUCED 
O CH-54A 
* AH-IG 
« 0H-6A 



i 









M 


u 


4 




/ 


i 


* ' 




A / £ 


\ V fk 


A h/ / 


\ \ \ 


A \\\W i 


^ m\\\\\\\\k 


m\\W t> 


\* I Bi 


A\\\\W /« 


\ \ 


• 411/ '* 


°\ \ 




MANUEVER ' 




Mo- SCATTER 


, 


BAND 

1 ' 



•5 .5 

INCREMENTAL LOAD FACTOR 



Figure 13. 



Gust-Induced toads are Significantly 
Less than Maneuver Loads. 
Conclusions 



The following conclusions were reached as 
a result of this study. It should be noted that 
Conclusions 1-3 are based on the computation of 
initial gust-induced load factors for various 
rotor systems mounted on a single fuselage and 
operating with the controls fixed throughout the 
gust encounters. 

1. The results of this study generally con- 
firm those of Reference 2, indicating that 
the current method for computing gust-in- 
duced load factors for helicopter rotors 
(Specification MIL-S-8698 (ARG)) results 
in realistically high values and should be 
revised. 



cause retreating blade angles of attack 
greater than the two-dimensional, steady- 
state stall angle, the inclusion of unsteady 
aerodynamic effects based on the model of 
Reference 8 results in gust-induced load fac- 
tors which are higher than those based on a 
steady aerodynamic model such as that used in 
Reference 2, 

Principal parameters influencing gust-in- 
duced load factor appear to be nondimen- 
sional blade loading, proximity of the rotor 
trim point to blade stall, and rate of pene- 
tration of the rotor into the gust. 

Gust loadings on military helicopters appear 
to be significantly lower than those due to 
maneuvers . 

References 

Crim, Aimer D., GUST EXPERIENCE OF HELICOP- 
TER AND AN AIRPLANE IN FORMATION FLIGHT, 
NACA Technical Note 335**, NACA, 195U. 

Harvey, K. W., Blankenship, B, L. Drees, 
J. M. , ANALYTICAL STUDY OF HELICOPTER GUST 
RESPONSE AT HIGH FORWARD SPEEDS. USAAVLABS 
Technical Report 69-I, September I969. 

Bisplingoff, R. L. , H. Ashley and R. L. 
Halfman, AEROELASTICITY, Addison-Wesley Pub- 
lishing Company, Inc., Cambridge, Mass. 1955. 

Bailey, F. J., Jr., A SIMPLIFIED THEORETICAL 
METHOD OF DETERMINING THE CHARACTERISTICS OF 
A LIFTING ROTOR IN FORWARD FLIGHT. NACA 
Report No. 716. 

Bergquist, R. R., Thomas G. C. TECHNICAL 
MANUAL FOR NORMAL MODES AEROELASTIC COMPUTER 
PROGRAM, July 1972. 

Bergquist, R. R., Thomas, G. C. USER'S 
MANUAL FOR NORMAL MODE BLADE AEROELASTIC 
COMPUTER PROGRAM, July 1972. 

Arcidiacono, P. J. , PREDICTION OF ROTOR IN- 
STABILITY AT HIGH FORWARD SPEEDS, VOLUME 1. 
STEADY FLIGHT DIFFERENTIAL EQUATIONS OF 
MOTION FOR A FLEXIBLE HELICOPTER BLADE WITH 
CHORDWISE MASS UNBALANCE. USAAVLABS Techni- 
cal Report 68-18A, February I969. 

Arcidiacono, P. J., Carta, F. 0., Cassellini, 
L. M. , and Elman, H. L., INVESTIGATION OF 
HELICOPTER CONTROL LOADS INDUCED BY STALL 
FLUTTER. USAAVLABS Technical Report 70-2, 
March 1970. 

Bellinger, E. D., ANALYTICAL INVESTIGATION 
OF THE EFFECTS OF UNSTEADY AERODYNAMICS VARI- 
ABLE INFLOW AND BLADE FLEXIBILITY ON HELI- 
COPTER ROTOR STALL CHARACTERISTICS. NASA 
CR-1769. 



If the gust amplitude is sufficient to 



99 



10. Bergquist, R. R. , HELICOPTER GUST RESPONSE 
INCLUDING UNSTEADY STALL AERODYNAMIC EF- 
FECTS. USAAVLABS Technical Report 72-68, - 
May 1973. 

.11. Giessler, F. Joseph; Nash, John F.; and 

Rockafellow, Ronald I., FLIGHT LOADS INVES- 
TIGATION OF AH-1H HELICOPTERS OPERATING IN 
SOUTHEAST ASIA, Technology, Inc., Dayton, 
Ohio; USAAVLABS Technical Report 70-51, 
U. S. Army Aviation Materiel Laboratories, 
Fort Eustis, Virginia, September 1970, AD 
, 878039. ■..•;' 

12. Giessler, F. Joseph; Nash John F.; and 
Rockafellow, Ronald I., FLIGHT LOADS INVES- 
TIGATION OF CH-5U HELICOPTER OPERATING IN .' 
SOUTHEAST ASIA, Technology, Inc., Dayton, 
Ohio; USAAVLABS Technical Report 70-73, 
Eustis Directorate, U. S. Army Air Mobility 
Research and Development Laboratory, Fort 
Eustis, Virginia, January 1971, AD 881238. , 

13. Giessler, F, Joseph; Clay, Larry E.; and 
Nash, John F., FLIGHT LOADS INVESTIGATION \ 
OF 0H-6A HELICOPTERS OPERATING IN SOUTHEAST. 
ASIA, Technology, Inc., Dayton, Ohip; 
USAAMRDL Technical Report 71-60, Eustis • '. 
Directorate, U. S. Army Air Mobility Re- . 

. Search and Development Laboratory, Fort 

, Eustis, Virginia, October 1971, AD 7308202;' 

Ik. Porter fields John D., and Maloney, Paul F., ' 
■ .. EVALUATION OF HELICOPTER FLIGHT SPECTRUM 
DATA, Kaman Aircraft Division, Raman Corpo- 
' ration, Bloomfield, Connecticut; USAAVLABS 
Technical Report 68-68, U. S. Army Aviation 
Materiel Laboratories , Fort Eustis , Vir- 
ginia, October 1968, AD 680280. ' 

15. Porterfield, JotmD., Smyth, William A. and 
Maloney, Paul F. , THE CORRELATION AND EVAL- 
UATION OF AH-1G, CH-5ltA, and 0H-6A FLIGHT 
. SPECTRA DATA FROM SOUTHEAST ASIA OPERA- 
TIONS, Kaman Aircraft Division, Kaman Corpo- 
ration, Bloomfield, Connecticut; USAAVLABS 
Technical Report 72-56, Eustis Directorate, 
U. S. Army Air Mobility Research and De- 
velopment Laboratory, Fort Eustis, Virginia, 
October 1972, AD 75555^- 



' : TABLE I, Reference Articulated . 
■ ' . Rotor Characteristics 

Density Slugs/ft 2 .002378 

Tip speed, ft/sec 700. 

Radius, ft 25 

; No. of blades h. 

Blade Chord, ft : 1.67 

Flap hinge off set ratio • 0.0k 

Twist, deg •'. . -8.0 

Young's Modulus, psi 10' 

Mass per unit length at. 0.75R slugs/ft 0.18 

Lock Number 10.0 

.Rigid body flatwise frequency 1.03P 

First bending flatwise frequency 2.66P 

Second bending flatwise frequency 5.06P 

Third bending flatwise frequency * 8.50P 

Rigid body chordwise frequency 0.25P 

First bending chordwise frequency 3.68P 

Second bending chordwise 10.20P 

First bending torsional frequency 5.72P 



100 



APPLICATION OF ANTIEESONANCE THEORY TO HELICOPTERS 

Felton D. Bartlett, Jr. 
Research Engineer 

William G. Flannelly 
Senior Staff Engineer 

Kaman Aerospace Corporation 
Bloomfield, Connecticut 



Abstract 

Antiresonance theory is the principle 
underlying nonresonant nodes in a struc- 
ture and covers both nonresonant nodes 
occurring naturally and those introduced 
by devices such as dynamic absorbers and 
antiresonant isolators . The Dynamic 
Antiresonant Vibration Isolator (DAVI) 
developed by Kaman Aerospace Corporation 
and the Nodal Module developed by the Bell 
Helicopter Company are specific examples 
of the applications of transfer anti- 
resonances. A new and convenient tech- 
nique is presented to numerically calcu- 
late antiresonant frequencies. It is 
shown that antiresonances are eigenvalues 
and that they can be determined by matrix 
iteration." 

Novel applications of antiresonance 
theory to' helicopter engineering problems,, 
using the antiresonant eigenvalue equation 
introduced in this paper, are suggested. 



Notation 

f force vector 

K stiffness matrix 

M mass, matrix 

y response vector 

Z impedance matrix 

6 antiresonant eigenvector 

on forcing frequency 

to antiresonant frequency 



In forced vibrations an antiresonance 
or "off- resonance node", is that frequency 
for which a system has zero motion at one 
or more points. A nodal point in a normal 
mode is a special case of an antiresonance. 
Driving point antiresonances have a 
readily grasped physical interpretation 
since they are the resonances of the 
system when it is restrained at the 
driving point. However, transfer anti- . 
resonances are not all real and, in 
general, have not been susceptible to 
analysis except in special cases. The 
eigenvalue equation for antiresonances 
used in this paper renders them as amen- 
able to analysis as are resonances. The 
mathematics for analyzing resonances are . • 
conventional and well-known^. 

Although general analytical methods 
for transfer antiresonances were not here- 
tofore commonly used, the existence of 
both driving point and transfer antires- 
onances in the forced vibration of a 
string were described by Lord Rayleigh 2 . 
The invention of the dynamic vibration 
absorber in 1909 gave antiresonances some ! 
practical engineering importance^. The 
absorber is an appendant dynamic system 
which has a driving point antiresonance 
at its fixed base natural frequency and it 
therefore reacts the forces at its base in 
the direction in which it acts. Isolating 
devices based on transfer antiresonances 
were not invented until this decade^. 
Sometimes natural fuselage transfer anti- 
resonances for major hub excitations 
occurred near a critical point and at the- 
proper frequency (e.g., the pilot's seat 
at blade passage frequency) by fortuity 
of helicopter design. Occasionally, 
engineers have manipulated transfer 
antiresonance frequencies and positions 
in design through lengthy trial-and-error 
response analyses." • However, the industry . 
has not used a direct analytical method 
for calculating the positions and fre- 
quencies of natural. antiresonances. 



Presented, at the AHS/NASA-Ames Special- 
ists' Meeting on Rotorcraft Dynamics, 
February 13-15, 1974. 



101 



Structures have antiresonances as an 
intrinsic "natural" property much as they 
have "natural" resonant frequencies. 
Natural transfer, or "of f -diagonal", anti- 
resonances are as important to structural 
dynamics engineering as are resonances. 
Unfortunately, many of the theorems which 
underly conventional analyses do not apply 
to transfer antiresonances. The anti- 
resonant dynamical matrix is in general 
nonsymmetrical and therefore not positive 
definite. This results in both left- 
handed and right-handed eigenvectors which 
are unequal and require a new orthogonal- 
ity condition for the calculation of 
successive eigenvectors. The antiresonant 
frequencies of the transfer antiresonance 
determinants are not necessarily real and 
the imaginary roots do not have a simple 
physical interpretation. These matters, 
along with the lack of an engineering 
eigenvalue formulation for antiresonances, 
may, in part, account for the relatively 
little attention given to natural anti- 
resonances over the years. 



Z m f k i Z . 
mn J irrj 

2_?f_i I _SJ_^_IS 

Z kn n * * ! Z kj 



.fA__j_ Z C_ 
Z R ] Z kj 



(4) 



If the impedance matrix is similarly 
partitioned so that the upper left-hand 
matrix does not contain the j-th row or 
the k-th column, then 



Z = 






(5) 



It follows from Equations (4) and (5) that 

T 



Z A = Z A 



From Equation (3) we obtain 



z A y- o 



(6) 



(7) 



Theory 

The steady-state equations of motion 
for an undamped spring-mass system vi- 
brating in the vicinity of equilibrium 
are: 



A kj antiresonance is defined such 
that for a force at k alone, the response 
at j is zero. Normalizing y and sub- 
stituting for Z A in Equation (7) results 
in the antiresonance eigenvalue equation. 



(K - to M)y = f 



(1) 



where the impedance matrix is defined as 

(2) 



Z = [8f i /3y j ] = (K 



a) 2 M) 



Let all the forces be zero except the 
force acting at the k-th generalized 
coordinate and further impose the re- 
straint of zero motion for the j-th 
generalized coordinate. The resulting 
eigenvalues are jk antiresonances of 
Equation (1) . Since Z is real and 
symmetric the antiresonance eigenvectors 
are real and the jk and kj antiresonance 
eigenvalues are real (positive or nega- 
tive) and equal. 

Partition Equation (1) so that the 
kj-th element of the impedance matrix 
appears in the lower right-hand corner. 



Z__ m f k I Z . 

mn ' i mj 

iLL'sL^fe 



-i- 



. kn 



where 



it, . 



o 



(3) 



102 



M A 9 r = -T- K A 9 r 
w r 



(8) 



A jk antiresonance eigenvalue equation is 
similarly defined by considering Equation 
(5) and making use of Equation (6). 



l sW " ~Tf *,\ 



(9) 



Equations (8) and (9) constitute a 
set of right-handed and left-handed eigen- 
vectors. Since Z^ is not symmetrical, 
the jk eigenvectors are not orthogonal 
but instead are biorthogonal with the kj 
eigenvectors-'-. Premultiply Equation (8) 
by ~ T, postmultiply Equation (9) by 9 , 

s 
and subtract to obtain 



(- 



0) 



0) 



T* VVr 



r s 
when s ^ r we have 



9 s K A 9 r = ° 



(10) 



(11) 



Thus , the kj antiresonance eigenvector is 
biorthogonal to the jk antiresonance 
eigenvector. 



When s = r the corresponding gen- 
eralized' mass "and stiffness are defined as 



3 M-9- 
r A r 



M_ 



9 K,9 = K 
r A r r 



(12) 



(13) 



Successive antiresonance eigenvectors are 
found by applying the biorthogonality 
condition and using classical matrix 
iteration techniques. The (n + l)st jk. 
antiresonant eigenvector is obtained from 
Equation (14) , 



(K, 



n 

- E 



. , K. 
1=1 l 



-)M„e j_, 

A n+1 



u n+l 



n+1 



(14) 



which establishes the method of sweeping.5 



Discussion of Theory 

Each antiresonant eigenvector con- 
sists of a pair which is biorthogonal with 
respect to both mass and stiffness. For 
driving point antiresonances (j = .k), the 
two eigenvectors are, obviously i the same. 
An N-degree-of-freedom system has N 2 
possible antiresonant eigenvectors cor- 
responding to all possible forcing and 
response coordinates. 

Since the mass and stiffness matrices 
are nonsymmetric in the antiresonance 
eigenvalue problem and consequently, not 
positive definite when j ^ k, the anti- 
resonant generalized masses and stiff- 
nesses may be either positive or negative. 
In other words, the antiresonance fre- 
quencies are not necessarily real. When 
j = k the antiresonant mass and stiffness 
matrices are symmetrical and positive 
definite, resulting in at least U-l, 
positive real antiresonances. As shown 
in Reference 6 the driving point anti- 
resonances lie between the natural 
resonant frequencies. 



Applications of Antirfesonance.' Theory 

To illustrate- the practical potential 
of antiresonance theory,- consider a ten- 
degree-of- freedom beam specimen with . 
springs to ground at stations- 3. and 9 and 
mass and stiffness parameters simulating.' 
a 9000 pound helicopter. Antiresonances 
are continuous functions of frequency and 
position and Figure 1 presents- a typical 
position spectrum plot of the specimen 
forcing at station 3 alone. ' The dashed 
vertical lines are the natural resonant 
frequencies determined conventionally.- 



When an antiresonance line crosses a 
natural frequency line there is a nodal 
point in the "natural mode". 




Figure 1. Antiresonance Lines Forcing at Station 3 



With the same techniques of altering 
masses and stiffnesses to avoid undesir- 
able natural resonances , the engineer can 
manipulate natural antiresonances.- The 
stiffness between stations 2 and 3 was 
increased by 11.8% in the K 2 3 term of the 
stiffness matrix and Figure 2 illustrates 
this effect in the natural frequencies 
and antiresonance lines. Similar changes, 
in the mass of the structure have a sim- 
ilar effect. This possibility for re- . 
sponse control indicates a profitable' ■'.. 
area for further exploration. 




-i 1 — t — i — i i 1 1 i 

10 
■ . FREQUENCY -HZ 



Figure 2. Antiresonance Lines with Stiffness Change 
Forcing at Station 3 



Conventional Use of the Dynamic Absorber 

A dynamic absorber .is an appendant 
dynamic system attached to a helicopter, 
usually at a point, as shown in Figure 3. 
When we eliminate the i-th. row and column., 
corresponding to the 'attachment point 
(see Figure 3) we obtain two uncoupled 
•systems. 



103 



o j z 

I aa 



is not at the tuned frequency of the 
(15) "resonator" and does not necessarily 
produce an antiresohance- at j for 
excitations along generalized 
coordinates other than k. 






N — ** " 














r- 




1 


- 








— - — 


Zff 


! ^ 

- 1 ■ — ■ 


1 
1 

--t- - 







Yf 




f 


z if 


! z h 


1 
1 

-4-- 
1 
1 


Z ia 


< 


Yj 


Ki 


► 





! Z ai 


z aa 




Ya 







L 




1 






L J 




L J 



The aforementioned system and equa- 
tions of motion are shown in Figure 4. 
To obtain an antiresonance at j for a 
force at k we eliminate the k-th row and 
j-th column from the equations of motions. 
This results in the antiresonant eigen- 
value equation, 



z ff 


fVj i z i 

1 1 
f jfk j f £ k I 





fVj i z. . i z. 
! z . j z 

1 ai I aa 



(17) 



which is of the form of Equation (7) 



Figure 3. Conventional Absorber 



The antiresonant eigenvalue equation is 
obtained from Equation (15) as 



[K M ]6 
aa aa J r 



2 r 



(16) 



which is of the form of Equation (7). 

If the absorber system were attached 
at I points, instead of one, we would 
eliminate the I rows and columns corres- 
ponding to the attachments and find the 
simultaneous antiresonant frequencies of 
all I points. 



Unconventional Use of the Dynamic Absorber 

In some instances there may be only 
one significant unreacted force on the 
helicopter as, for example, when an in- 
plane isolation system or in-plane hub or 
flapping absorbers leave small hub moments 
but a relatively large vertical oscilla- 
tory force. We can use a dynamic absorber 
in the fuselage at some point i as a 
"resonator" to shift antiresonance lines 
so that there exists an antiresonance at 
another point j (e.g., the pilot's seat) 
for the one remaining large force or 
moment along the k-th generalized coor- 
dinate. This is creating a jk antires- 
onance by manipulation of a "resonator" 
at point i. The jk antiresonant frequency 



( i 

V. .i . — 








r iiii 




— — 




— — 


i i i 
z «i i z fj i 2 * i ° 

f*k | f ^ k| f5>k| 
**\ | I I 




Y f1 







i __• __!__ 

Z kf 1 | Z kj | Z ki | ° 

f 1 ! L + .._4.__ + __ 

Z if1 , 1 Z ij I Z « | Z ia 

o ; o ! z ai j 2 aa 


l Yi I r 
._. 


L Y aJ L 



Figure 4. Antiresonance at Station j from a Resonator at Station i 

This technique of using a remote 
dynamic absorber as a "resonator" allows 
the engineer to obtain an antiresonance, 
to a given excitation, at points where 
structural limitations prevent installa- 
tion of an absorber. When the new res- 
onant frequency introduced by the 
"resonator" cuts across a natural anti- 
resonance line, the shifts are dramatic 
as shown in Figure 5. Figure 5 illustrates 
the antiresonance lines in the specimen, 
forcing at station 3, when an absorber of 
77.2 pounds tuned to 7.7 Hz is added to 
station 2. The natural frequency intro- 
duced by the absorber intersects the 
antiresonance line of Figure 1 and pro- 
duces new antiresonances at all stations, 



104 



forcing at station 3 

10 




10 

FREQUENCY -HZ 
Figures. Antiresonance Lines with Dynamic Absorber at 
Station 2, Forcing at Station 3 



The effect at station 5 of the 77.2 
pound absorber located at station 2 and 
tuned to 7.7 Hz, in terms of both anti- 
resonant frequency and bandwidth, is the 
same as the effect produced by a 193 pound 
absorber located at station 5 itself and 
tuned to 8.0 Hz. Bandwidth is here 
defined as the difference between the 
antiresonance frequency and the nearest 
natural frequency. This comparison is 
presented in Figure 6. The approximately 
two to one reduction in absorber weight 
does not imply that such savings are 
always obtainable. 



iu 5 




— 77.2 LB ABSORBER AT 

STATION 2 TUNED TO 7.7 HZ 

1\-193 LB ABSORBER AT 

I \ STATION 6 TUNED TO 8.0 HZ 



i — i — i — i 1 1 1 — | r 

7.0 7.S 8.0 8.S 

FREQUENCY -HZ 

Figure 6. Comparison of Antiresonance Lines 
for Two Absorbers 



Antiresonant Isolators 

Passive antiresonant isolation 
devices have received considerable 
attention from the industry in recent 
years. Notable among these are Bell 
Helicopter's Nodal Module, Kaman's BAVI 
series, and the Kaman COZID. 

Figure 7 illustrates the antiresonant 
isolation system and corresponding equa- 
tions of motion. The excited structure 

105 







r- — 




*- — 




Y . 




\ 


- 


Y 
Y 


,=< 







L J 




L .) 



Figure 7. Antiresonance isolation 

is coupled to another structure through, 
and only through, the antiresonant 
isolation system which has inertial and 
elastic elements. Any isolator with a 
single input and single output, or a 
symmetrical arrangement having the same 
effect, has antiresonant frequencies given 
by the eigenvalues of 



fDl_j^DD 
K OI J K OD 



-1 



M, 



M I M 
T>I | OD 









e 


V 


e 



(18) 



where I, .0, and D represent the input, 
output and internal isolator degrees-of- 
freedom, respectively. The two-dimen- 
sional and three-dimensional DAVIs have, 
respectively, each two and three un- 
coupled equations of the form of Equation 
(18) . Two outputs displaced with dynamic 
symmetry from a given input, or the con- 
verse, are also described by Equation (18) 
because the roots are not changed by 
transposing a matrix. 

It is possible to solve for simul- 
taneous antiresonances on arbitrarily 
placed multiple outputs for an equal 
number of arbitrarily placed multiple 
inputs by letting I and O be greater than 
one in Equation (18) . However, such 
simultaneous antiresonances will, in the 
general case, occur only for those dis- 
tributions of input forces given by the 
product of the rectangular impedance 
matrix of rows corresponding to the forced 
degrees of freedom and the vector of dis- 
placements. This is the reason why 
multiple input-output antiresonant isola- 
tors are not used in engineering. It is 
observed that the impedance matrix of 



Figure 7, is, in general, nonsymmetric 

while the impedance matrix of Figure 3 

is necessarily symmetric. That is the 1. 

mathematically distinguishing feature 

between absorbers and antiresonant 

isolators . 

2. 
It is obvious from Equation (18) that 
an infinite number of mechanical systems 
exist which will produce antiresonant 
transmissibilities at more than one fre- 
quency. Such systems can be analytically 3. 
synthesized using desired antiresonant 
frequencies , the biorthogonality conr 
dition, and the methods of Reference 7. 
However, not all such synthesized systems 4. 
will be physically realizable and not all 
of the physically realizable synthesized 
systems will be practical from an 
engineering standpoint. 

An immediately practical application 
of Equation (18) would be the investiga- 5. 
tion of physical multi-input antiresonant 
isolators with internal coupling using 
simpler engineering arrangements for 
multi -harmonic antiresonances than has yet 
been achieved. 6. 



Conclusion 

This paper has presented a solution 7. 
to the antiresonant eigenvalue problem. 
It has been shown that antiresonances can 
be determined by < matrix iteration tech- 
niques. Antiresonant nodes introduced by 
dynamic absorbers and antiresonant iso- 
lators have been discussed to illustrate 
the novel application of the theory to 
helicopter engineering problems. 



' References 

Meirovitch, L. , ANALYTICAL METHODS IN 
VIBRATIONS, McGraw-Hill Book Co., New 
York, 1967. 

Strutt, J.W. , Baron, Rayleigh, THE 
THEORY OF SOUND, 2nd Edition, Volume 
1, Sec. 142a, Dover Publications, 
New York, 1945. 

Den Hartog, J.P. , MECHANICAL VIBRA- 
TIONS, 4th Edition, McGraw-Hill 
Publishing Co., New York, 1956. 

Kaman Aircraft Report RN 63-1, 
DYNAMIC ANTIRESONANT VIBRATION 
ISOLATOR (DAVI) , Flannelly, W.G. , 
Kaman Aircraft Corporation, 
Bloomfield, Connecticut, November 
1963. 

Rehfield, L.W. , HIGHER VIBRATION 
MODES BY MATRIX ITERATION, Journal 
of Aircraft , Vol. 9, No. 7, July 1972, 
p. 505. 

Biot, M.A. , COUPLED OSCILLATIONS OF 
AIRCRAFT ENGINE-PROPELLER SYSTEMS, 
Journal of Aeronautical Society , 
Vol. 7, No. 9, July 1940, p. 376. 

USAAMRDL Technical Report 72-63B, 
RESEARCH ON STRUCTURAL DYNAMIC 
TESTING BY IMPEDANCE METHODS, 
Giansante, N. , Flannelly, W.G. , 
Berman, A., U. S. Army Air Mobility 
Research and Development Laboratory, 
Fort Eustis, Virginia, November 19 72. 



106 



THE EFFECT OF CYCLIC FEATHERING MOTIONS 

ON 

DYNAMIC ROTOR LOADS 

Keith W. Harvey 

Research Engineer 

Bell Helicopter Company 

Fort Worth, Texas 



Abstract 

The dynamic loads of a helicopter 
rotor in forward flight are influenced 
significantly by the geometric pitch 
angles between the structural axes of the 
hub and blade sections and the plane of, 
rotation. 

The analytical study presented in- 
cludes elastic coupling between inplane 
and out-of-plane deflections as a function 
of geometric pitch between the plane of 
rotation and the principal axes of inertia 
of each blade. In addition to a mean col- 
lective-pitch angle, the pitch of each 
blade is increased and decreased at a one- 
per-rev frequency to evaluate the dynamic 
coupling effects of cyclic feathering mo- 
tions. The difference in pitch between 
opposed blades gives periodical coupling 
terms that vary at frequencies of one- and 
two-per-rev. Thus, an external aerody- 
namic force at n-per-rev gives forced res- 
ponses at n, n±l, and n+2 per rev. 

The numerical evaluation is based on 
a transient analysis using lumped masses 
and elastic substructure techniques. A 
comparison of cases with and without cyclic 
feathering motion shows the effect on com- 
puted dynamic rotor loads. The magnitude 
of the effect depends on the radial loca- 
tion of the pitch change bearings. 

Introduction 

For a stiff -in-plane rotor system, 
the blade chordwise stiffness may be 20 
to 50 times greater than the blade beam- 
wise stiffness. The elastic structure 
tends to bend in the direction of least 
stiffness, resulting in dynamic coupling 
between out-of-plane and inplane motions 
as a function of the geometric pitch 
angles due to collective pitch, built-in 
twist, forced cyclic feathering motions of 
the torsionally-rigid structure, and elas- 
tic deformation of the blade and control 
system in the torsional mode. 

Typical cruise conditions for a mod- 
ern helicopter require collective pitch 
angles of 14 to 16 degrees at the root, 
depending on the amount of built-in twist. 
Cyclic feathering motions of 6 to 7 de- 
grees are required to balance the one-per- 
rev aerodynamic flapping moments. In cur- 
rent design practice, elastic torsional 



deflections of the blade and control sys- 
tem of a stiff -in-plane rotor are generally 
less than one degree. The largest part of 
the angular motion in the blade-torsion de- 
gree of freedom, therefore, is the forced 
feathering motion due to cyclic pitch. 

Periodic variations of the inplane/ 
out-of-plane elastic coupling terms are 
caused when the geometric pitch angle of 
each blade is increased and decreased at a 
frequency of one cycle per rotor revolution. 
When one blade is at high pitch and the op- 
posed blade is at low pitch, an asymmetri- 
cal physical condition exists with respect 
to a reference system oriented either to 
the mast axis or to the plane of rotation. 
One-half revolution later, the reference 
blade is at low pitch and the opposed blade 
is at high pitch. Thus, periodic dynamic 
coupling occurs at the principal frequency 
of one-per-rev with respect to a rotating 
coordinate system. The coupling terms are 
nonlinear functions of blade pitch; hence, 
these terms also have 2-per-rev content. 

Both the steady and periodic coupling 
terms have been treated in an analytical 
study of the effects of one-per-rev cyclic 
feathering motions on dynamic rotor loads. 
Equation's have been derived and programmed 
for a digital computer solution of the 
transient response of an elastic two-bladed 
rotor. 

The rotor is modeled by elastic sub- 
structure elements and lumped masses, for 
which the accelerations and velocities are 
integrated over small time increments to 
determine time histories of deflections, 
inertia loads, bending moments, etc. The 
time-variant analysis includes the capa- 
bility to calculate rotor instabilities. 
The present computer program has been 
tested for this capability, but further 
discussion of instabilities is beyond the 
intent of the paper. 

Dynamic rotor loads have been calcu- 
lated for a parametric series of rotors, 
where the coupled natural frequencies were 
tuned over the range of contemporary de- 
sign practice for teetering rotors. A de- 
scription of the analysis and a summary of 
computed results is presented. 



107 



Objective 



A primary consideration in the design 
of a helicopter rotor is to minimize os- 
cillatory bending loads, or at least to 
reduce the loads to a level that will en- 
sure satisfactory fatigue life. During 
early stages of design, the principal 
method of evaluating the dynamics of a 
proposed rotor is to calculate its coupled 
rotating natural frequencies. If required, 
design changes are made to achieve suffi- 
cient separation between the natural fre- 
quencies and harmonics of the rotor .oper- 
ating speed. 

Current practice at Bell Helicopter 
Company is to require a separation of 0.3 
per rev for all flight combinations of 
rotor speed and collective pitch. One 
purpose of the present analytical develop- 
ment is to determine whether the separa- 
tion rule may be relaxed due to beneficial 
effects of cyclic feathering motions on 
rotor dynamic response. 

Collective and Cyclic Modes 

The calculation of natural frequen- 
cies for semi-rigid rotors uses a coor- 
dinate system that is based on the plane 
of rotation. The orientation of the cen- 
trifugal force field, the angular motion 
allowed by the flapping hinge(s), and the 
constraints of opposing blades lead to the 
segregation of natural frequencies into 
collective modes, cyclic modes, and (for 
four-bladed rotors) scissor or reaction- 
less modes. This procedure allows the use 
of continuous-beam theory for a single 
blade, where the centerline boundary con- 
straints are imposed from conditions of 
symmetry or asymmetry to match deflections, 
slopes, shears, and moments for the other 
blades . 

The centerline boundary conditions for 
the collective mode (Figure 1) are: 

- zero vertical (out-of-plane) slope 
change , 

- vertical deflection constrained by 
mast tension/compression, 

- inplane slope constrained by mast 
torsion, and 

- zero inplane translation. 

The centerline boundary conditions for 
the cyclic mode (Figure 2) are: 

- vertical slope change unrestrained 
(except with flapping springs), 

- zero vertical deflection, 

- zero inplane slope change, and 



inplane deflection constrained by 
mast shear. 




ROTOR SPEED 300 

FIGURE 1. TYPICAL COLLECTIVE MODE 

FREQUENCIES AND MODE SHAPES. 




ROTOR SPEED 
FIGURE 2 



300 



TYPICAL CYCLIC MODE FRE- 
QUENCIES AND MODE SHAPES. 



108 



For the reactionless modes, the cen- 
terline boundary conditions are: 

- zero slope change and zero trans- 
lation in both the inplane and 
vertical directions. 

Uncoupled frequencies are determined 
by setting the geometric pitch angle of 
each elastic element to zero. The un- 
coupled frequencies are shown in Figures 
1 and 2 by the labeled curves. Note that 
the frequencies of the vertical (out-of- 
plane) modes are highly dependent on rotor 
speed, and that the frequencies of the in- 
plane modes are only slightly dependent 
on rotor speed. 

Coupled natural frequencies are shown 
as small circles in the figures. Typical 
collective modes have very small frequency 
shifts as a function of collective pitch. 
However, the cyclic modes (Figure 2) 
couple significantly with collective pitch. 
Note that the inplane frequency decreases 
and the vertical frequencies increase with 
collective pitch. The method of deter- 
mining these coupling effects is given in 
Reference 1. 

By using only one blade plus appro- 
priate boundary conditions, this method of 
calculating rotor natural frequencies is 
based on one explicit assumption, i.e., 
all other blades are at the same geometric 
pitch angle as the reference blade. If 
the blades are at different pitch angles, 
then the conditions of symmetry or asym- 
metry are not present. The inclusion of 
cyclic feathering motion, therefore, re- 
quires that the analysis treat separately 
each blade of the rotor and provide a 
means of matching the centerline slopes 
and deflections. 

Elastic Substructured Rotor Analysis 

A digital computer program has been 
developed to study the effects of cyclic 
feathering motions on dynamic rotor loads. 
The Bell Helicopter computer program is 
identified by the mnemonic ESRA for Elas- 
tic Substructured Rotor Analysis. 



The analysis is a transi 
of elastic rotor blade motion 
coupling terms for each blade 
separately and the necessary 
are imposed on each blade to 
and deflection continuity at 
centerline. Each blade is co 
being divided into a discrete 
segments, with uniform weight, 
ness properties over the leng 
ment. The geometric pitch an, 
segment is a function of roto 
position and input values of 
lateral cyclic pitch. To rep 
hub structure that is inboard 



ent solution 
s , where the 

are treated 
constraints 
insure slope 
the rotor 
nsidered as 

number of 

and stiff- 
th of a seg- 
gle of each 
r azimuth 
fore/aft and 
resent the 

of the pitch- 



change bearings, the inboard elastic ele- 
ment may be specified as an uncoupled 
element (geometric pitch equals zero). 

All forces are applied at the ends of 
the elastic elements. Slope and deflection 
changes over the length of a segment are 
based on linear moment distributions versus 
span. For compatibility, shear over the 
segment length must be constant, which re- 
quires concentrated forces for both inter- 
nal and external forces. In its simplest 
form, the analysis follows a lumped-mass 
approach. All of the important rotor dy- 
namic characteristics may be retained with 
this method, however, by using the con- 
cepts of equivalent structural segments. 

The dynamic response equations are 
solved by a step-by-step iterative method in 
order to include transient conditions. If 
the initial deflections and velocities are 
specified (spanwise distributions for each 
blade), then the internal bending moment 
distributions are found with respect to the 
rotating reference system. Internal shear 
distributions are obtained from the moment 
distributions, and summed with applied air- 
load forces and inertial components of the 
centrifugal force field to determine span- 
wise distributions of accelerations. 

The first estimates of deflection and 
velocity changes are calculated for con- 
stant acceleration during the integration 
time step. Then bending moments, shears, 
and accelerations are calculated for the 
end of the time step. Subsequent deflec- 
tion and velocity estimates are based on 
accelerations changing linearly with time, 
and the iterations continue until a pres- 
cribed error limit is satisfied for the 
entire set of accelerations, or until a 
limit is reached on the number of itera- 
tions . 

In recognition of the problems inher- 
ent with this type of numerical integration, 
the initial development of the ESRA com- 
puter program has been limited to a quali- 
tative study of cyclic feathering effects. 
The current program represents each blade 
with four elements, each with beamwise and 
chordwise bending elasticity. Only the 
forced rigid-body motion is allowed in the 
blade-pitch degree of freedom, i.e., elas- 
tic blade torsion is not considered. 



The current computer 
ited to two-bladed rotors, 
al impedance of the drive 
to be zero. In practice, 
a two-per-rev torque from 
proportional to the drive- 
times the Hooke ' s- joint an, 
which is a function of rot 
rotor is the predominant i 
of the drive system, and a 
mation for two-bladed roto 



program is lim- 

and the torsion- 
system is assumed 
the rotor senses 
the mast that is 
system impedance 
gular oscillation, 
or flapping. The 
nertia component 
good approxi- 
rs is to assume 



109 



that the true axis of rotation remains per- 
pendicuLar to the tip-path plane even when 
the tip path plane is not perpendicular to 
the mast. Thus, Coriolis accelerations 
equal to the product of coning times flap- 
ping are not appropriate in a two-bladed 
rotor analysis. 

Bending deflections of the elastic 
elements are linearized; therefore, Cori- 
olis accelerations from radial foreshort- 
ening are excluded also. Vertical and in- 
plane translational motions of the rotor 
center are not included in the current 
version of the program. 

Referring to the description above, 
the formulation of the analysis allows the 
removal of these limiting assumptions. 
For instance, nonlinear bending deflections 
and Coriolis accelerations may be included 
by a direct addition to the inertial forces 
acting on each mass. Translation of the 
rotor centerline, additional blades, con- 
trol system flexibility, elastic blade 
torsion, and nonlinear hub and control 
kinematics also may be added within the 
existing computational method. 

With the limitation of four elastic 
elements for each blade, plus provisions 
for slope and deflection continuity at the 
rotor centerline, the current ESRA pro- 
gram allows 15 distinct vibration modes 
for the rotor: 

- 3 rigid -body modes (flapping, mast 
torsion, blade pitch) 

- 6 coupled elastic collective modes 
(3 vertical, 3 inplane) 

- 6 coupled elastic cyclic modes (3 
vertical, 3 inplane) 

In attempts to predict rotor loads for 
two-bladed rotors, emphasis is placed on 
response components at least up to the 
third harmonic of rotor speed. Three-per- 
rev airloads excite the cyclic mode that 
derives from the first elastic asymmetric 
mode in the out-of-plane direction. Four 
elastic elements for each blade should 
provide a very satisfactory dynamic repre- 
sentation for this frequency range. At a 
frequency of f ive-per-rev , the second elas- 
tic mode would be excited and computed 
loads may be marginally valid. Current 
design practice is to minimize higher fre- 
quency loads by proper tuning of the rotor 
natural frequencies, as discussed earlier. 

Numerical Evaluation 

A parametric computer study was ac- 
complished to resolve a basic question: 

With respect to the natural fre- 
quency of the first coupled vertical 



elastic cyclic mode at or near 
3 per rev, how much does cyclic 
feathering motion affect 3-per- 
rev dynamic rotor loads? 

Selection of Rotor Dynamic Characteristics 

Corresponding to a Huey main rotor, 
the computer study was based on a 48-foot 
diameter 2-bladed semi-rigid rotor, oper- 
ating at 300 RPM. Two basic design ap- 
proaches were selected as end points for 
the evaluation. ; 

1. A constant blade weight distri- 
bution of 1.20 lb/in. with no dynamic - 
tuning weights was picked to simulate the 
early production Huey rotors. Uniform 
beamwise and chordwise stiffness values 
were determined to locate the two lowest 
coupled cyclic mode frequencies at 1.40/rev 
(inplane) and 2.60/rev (vertical) for a 
collective pitch of 14.75 degrees. The fan 
plot of cyclic mode natural frequencies for 
this rotor is shown in Figure 3. 



BASIC SECTION WT 
=1.20 LB/IN 

NO TIP WEIGHT 

_ 14.75° COLL. PITCH 
o 

25 
U 

g. 




-1/REV 



ROTOR SPEED, RPM 



300 



FIGURE 3. CYCLIC MODE, TUNED 
BELOW 3/REV. 



2. A very recent rotor development at 
Bell (the Model 645 rotor) was simulated by 
a configuration with a constant blade weight 
distribution of 1.00 lb/in plus a dynamic 
tuning weight of 100 pounds located at the 
blade tip. Uniform beamwise and chordwise 
stiffness values were determined to locate 
the coupled cyclic mode frequencies at 
1.40/rev (inplane) and 3.40/rev (vertical), 
as shown in Figure 4, again for 14.75 de- 
grees of collective pitch. 



110 




ROTOR SPEED, RPM 



300 



FIGURE k. CYCLIC MODE,, TUNED 
ABOVE 3/REV. 

Between the two basic configurations, 
a series of intermediate rotor parameters 
was established by stepping the uniform 
blade weight from 1.20 down to 1.00 by 
increments of 0.025 lb/in., while increas- 
ing the tip weight from 0. to 100. by in- 
crements of 12.5 pounds. Beamwise and 
chordwise stiffnesses were varied to hold 
the coupled inplane frequency at 1.^0/rev 
while tuning the coupled vertical fre- 
quencies from 2.60/rev to 3.40/rev in in- 
crements of 0.10/rev. Thus, to compensate 
for the program restriction of no hub mo- 
tion, proper placement of the coupled in- 
plane frequency was maintained by the 
selection of rotor stiffness. This ap- 
proach affects the spanwise. distribution 
of inplane bending moments, but is entire- 
ly adequate for a qualitative evaluation. 

All of the above frequencies were 
tuned with the first segment uncoupled 
(hub structure to .25 radius), which maxi- 
mized the coupling of the vertical mode 
near 3/rev and minimized the coupling of 
the inplane mode near 1/rev. 

Additional input data was taken 
directly from the Bell Helicopter Rotor- 
craft Flight Simulation, program C81-68 
(References 2,3), for a Model 309 King- 
Cobra flying at 150 knots. Data used in 
the present computer evaluation included 
a collective pitch setting at the root of 
lif.75 degrees, a total cyclic pitch of 
6.30 -degrees, and the spanwise distributed 
airloads up to and including the third* 
harmonic components. 



The study results presented below are 
based, therefore, on full-scale parameters 
that are realistic with regard to current 
helicopter design practice. Although di- 
rect correlation with measured loads is not 
possible because of the simplifying assump-, 
tions , it may be noted that the magnitude 
of calculated bending moments is well with- 
in the expected range. 

Computed Results 

The forced response was computed for 
the series of nine parametric rotor con-' 
figurations, where the inplane coupled fre- 
quency was held at 1.40/rev and the verti- 
cal coupled frequency was varied from 2.60/ 
rev to 3.40/rev. The dynamic rotor loads 
for each configuration were calculated 
twice, once with cyclic feathering and 
once without cyclic feathering. 

Figure 5 shows the 2/rev vertical 
bending moment at the rotor centerline as 
a function of natural frequency of the ver- 
tical elastic cyclic mode. The 1/rev vari- 
ation in structural coupling due to cyclic 
pitch, and the 3/rev applied airloads pro- 
duce a 2/rev component of bending moment. 
This additional component peaks and changes 
sign as the vertical mode is tuned through 
3/rev. For the two-bladed rotor, 1/rev and 
3/rev vertical bending moments at the cen- 
terline are negligible. 



3 
i 

z 



o 

o 
o 



z 



45- 



4tt 



35 



30 



S 25| 

o 

z 



a 

z 
w 



20 



2/REV 



WITHOUT CYCLIC FEATHERING 



WITH CYCLIC FEATHERING 



15 



2.6 2,8 3,0 3.2 3.4 
COUPLED VERTICAL NAT. FREQ. , PER REV 
FIGURE 5. VERTICAL MOMENT AT CENTERLINE 



111 



InpLane bending moments at the rotor 
centerline are shown in Figure 6. The 
large peak in the overall oscillatory mo- 
ment occurs as the coupled vertical mode 
is tuned through resonance at 3/rev. Note 
that the coupling associated with cyclic 
feathering increases the 1/rev response by 
about 5 percent for the vertical frequency 
tuned to 2.6/rev. In other respects, the 
effect of cyclic feathering appears to be 
minimal. 

Beamwise moments and chordwise mo- 
ments at midspan are shown in Figures 7 
and 8, respectively. Two-per-rev moment 



WITHOUT CYCLIC FEATHERING 
WITH CYCLIC FEATHERING 



340 

320-1 

300 

280- 

260- 

240- 

220 



3 200-1 
i 

25 





'/ OVERALL N 
'' OSCILLATORY 



--,_ 1/REV 



o 
o 
o 



H 

z 



180 
160 



g 140-1 

s •• 

o 

5- 120 

a 
z 
w 

« 100 



80 
60 



3/REV 




V— 



2.6 2.8 3.0 3.2 3.4 
COUPLED VERTICAL NAT. FREQ. , PER REV 
FIGURE 6. INPLANE MOMENT AT CENTERLINE 



components are not shown in the figures 
because of their small magnitudes. The 
significance of the cyclic feathering ef- 
fects at midspan is consistent with that 
indicated in earlier figures for the rotor 
centerline. 



80 



70 



m 

►J 

z 


60 


o 
o 
o 

1—1 


50 


z 


40 


1 


30 


o 

z 

M 

Q 


20 


W 





10 



WITHOUT CYCLIC FEATHERING 
WITH CYCLIC FEATHERING 



OVERALL 
OSCILLATORY 



°VS 




n 3/REV 



2.6 2.8 3.0 3.2 3.4 
COUPLED VERTICAL NAT. FREQ. , PER REV 
FIGURE 7. BEAMWISE MOMENT AT MID SPAN 



140 



pq 

•J 120 

z 



glOO 



H 

Z 



o 

o 

z 

I— I 
Q 
Z 
W 
« 



8a 

60 
40 
20 



OVERALL 
OSCILLATORY 




oU 



WITHOUT CYCLIC FEATHERING 
WITH CYCLIC FEATHERING 



2.6 2.8 3.0 3.2 3.4 
COUPLED VERTICAL NAT. FREQ. , PER REV 
FIGURE 8 . CHORDWISE MOMENT AT MID SPAN 



112 



As discussed in a previous section, 
the aeroeLastic effect of blade bending 
velocity was excluded from this study by 
basing the response- calculations on a pre- 
scribed set of airloads. No inference is 
intended regarding the magnitude of aero- 
dynamic damping that may be associated with 
elastic bending velocities. Conversely, 
the procedure was selected so that the 
time-variant structural couplings could be 
studied in an analytical environment that 
does not include other sources of damping. 

The computed responses appear as un- 
damped resonances centered at 3/rev, from 
which it follows that cyclic feathering 
motions do not provide any significant 
amount of equivalent damping to suppress 
3/rev dynamic loads. Regarding the verti- 
cal cyclic mode near 3/rev, in particular, 
the effect of cyclic feathering motion 
does not provide relief for the design rule 
that requires 0. 3/rev separation of coupled 
frequencies from excitation harmonics of 
rotor speed. 

The results presented above are all 
based on rotor structural simulations with 
the inboard 25 percent radius treated as 
non-feathering hub structure. This option 
of the program was selected to maximize 
the coupling (as a function of collective 
pitch) of the vertical cyclic mode near 
3 per rev. As noted, the largest change 
in rotor loads due to the inclusion of 
cyclic feathering motions was a 5 percent 
increase in inplane bending moments at the 
rotor centerline. 

The pitch-change or feathering bear- 
ings of production two-bladed main rotors 
are located typically at about 10 percent 
radius. In this respect at least, the 
above results are based on a dynamic model 
that is not representative of actual de- 
sign practice. 

To evaluate the importance of the 
radial location of the bearings, another 
set of rotor loads was computed for a 
case where the entire radius is in the 
feathering system. 

A constant blade weight distribution 
of 1.20 lb/in with no dynamic tuning 
weights was selected, as before, to simu- 
late the early production Huey rotors. 
The structural properties of the rotor 
were modified to maintain a 1.40/rev 
natural frequency for the coupled inplane 
cyclic mode. For the modified parameters, 
the natural frequency of the coupled verti- 
cal cyclic mode is 2.87/rev. 

The computed results are shown in 
Figures 9 through 12 for the case in which 
the feathering bearings are located at 
zero percent radius. The bar< graphs show 
first, second, and third harmonics plus 



overall levels of oscillatory bending mo- 
ments. The open bars are for the condition 
of no cyclic pitch, i.e., the geometric 
pitch of the elastic elements held fixed 
at the specified value of collective pitch. 
The closed bars are for the condition that 
the geometric, pitch of the elastic struc- 
ture is a function of both collective pitch 
and cyclic pitch. 

Vertical and inplane oscillatory bend- 
ing moments at the rotor centerline are 
shown in Figures 9 and 10, respectively. 
The vertical moments are not changed 



cn 



z 



o 
o 
o 



Z 



s 

o 

z 

M 

o 

z 
w 
PQ 



100- 
80- 
60 
40 
20 




WITHOUT CYCLIC 
WITH CYCLIC 



1 2 3 OVERALL 
HARMONIC (PER REV) 

FIGURE 9. VERTICAL MOMENT AT CENTERLINE 



9 



300 



o 

2 200 



z 



3 

o 

z 



100- 



w 



CZ1 WITHOUT CYCLIC 
5 WITH CYCLIC 



FIGURE 10. 



1 2 3 OVERALL 
HARMONIC (PER REV) 

INPLANE MOMENT AT CENTERLINE 



113 



signif icantLy by the inclusion of cyclic 
pitch. However, the inplane centerline 
moments increase by 57 percent, with both 
the first and third harmonics contributing 
to the increase. 

Beamwise and chordwise oscillatory 
bending moments at 50 percent radius are 
shown in Figures 11 and 12. Most of the 
increase in the overall oscillatory mo- 
ments at mid-span is due to an increase 
in 3/rev response. 



Due to cyclic feathering motions, a 
significant increase (57 percent) in rotor 
loads is indicated with the feathering 
bearings at zero percent radius; a mini- 
mal increase (5 percent) is indicated with 
the feathering bearings at 25 percent 
radius. This suggests that the radial lo- 
cation of the feathering bearings may have 
a controlling influence on the magnitude 
of the cyclic-feathering effect. Further 
study of this relationship is in progress. 

Conclusions 



« 100 

z 



o 
o 
o 



H 
55 



I 



w 



80- 



60- 



40 



20- 



WITHOUT CYCLIC 
WITH CYCLIC 




12 3 OVERALL 
HARMONIC (PER REV) 

FIGURE 11. BEAMWISE MOMENT AT MID-SPAN 



1. The cyclic feathering motions of 
a helicopter rotor cause time-dependent 
elastic coupling due to asymmetrical pitch 
on opposed blades. The effect of these 
motions on dynamic loads may be calculated 
by modeling the rotor with elastic sub- 
structure elements, by providing individual 
treatment of each blade , and by matching 
slopes and moments at the rotor centerline. 

2. Cyclic feathering "motion of the 
elastic blade structure does not cause any 
significant damping effect on the 3-per-rev 
dynamic loads of a two-bladed semi-rigid 
rotor. The design rule requiring 0.3? rev 
separation between coupled natural fre- 
quencies and aerodynamic excitation fre- 
quencies should not be relaxed on the 
expectation of beneficial effects from 
cyclic feathering. 

3. The inplane one-per-rev rotor loads 
of a stiff -in-plane rotor are affected sig- 
nificantly by cyclic feathering of the 
elastic structure. The magnitude of the 
effect is decreased as the feathering bear- 
ings are moved radially away from the rotor 
centerline. 



a 100 



55 

M 
O 

O 

o 



E-t 

55 



55 

M 

Q 

55 
W 

pq 



80- 



60 



40 



20- 



WITHOUT CYCLIC 
WITH CYCLIC 



m n^ rm i m 



1 2 3 OVERALL 
HARMONIC (PER REV) 

FIGURE 12. CHORDWISE MOMENT AT MID-SPAN 



References 

Blankenship, B. L. and Harvey, K. W. , 
A DIGITAL ANALYSIS FOR HELICOPTER PER- 
FORMANCE AND ROTOR BLADE BENDING MO- 
MENTS , Journal of the American Heli- 
copter Society , Vol,! 7 , No . 4 , 
October 1962, pp 55-69. 

Duhon, J. M. , Harvey, K. W. , and 
Blankenship, B. L. , COMPUTER FLIGHT 
TESTING OF ROTORCRAFT , Journal of the 
American Helicopter Society , Vol. 10, 
No. 4, October 1965, pp 36-48. 

USAAVLABS Technical Report 69-1, 
ANALYTICAL STUDY OF HELICOPTER GUST 
RESPONSE AT HIGH FORWARD SPEEDS, Harvey, 
K. W. , Blankenship, B. L. , and Drees, 
J. M. , U.S. Army Aviation Materiel 
Laboratories, Ft. Eustis, Virginia, 
September 1969. 



114 



CONTROL LOAD ENVELOPE SHAPING BY LIVE TWIST 

F. J. Tarzanin, Jr. 
Senior Engineer, Boeing Vertol Company, Philadelphia, PA 

P. H. Mirick 
Aerospace Engineer, Eustis Directorate, USAAMRDL, Ft. Eustis, VA 

Abstract 



For flight conditions at high blade 
loadings or airspeeds, the rotor control 
system experiences a rapid load growth, 
resulting from retreating blade stall. 
These loads frequently grow so large that 
the aircraft flight envelope is restricted 
long before the aircraft power limit is- 
reached. A theoretical study of one flight 
condition and a limited model test have 
shown that blade torsional flexibility 
plays a major role in determining the mag- 
nitude of these large, stall-induced con- 
trol loads. Recently, an extensive 
analytical investigation* was undertaken 
to determine the effect of changing blade 
torsional properties over the rotor flight 
envelope. The results of this study showed 
that reducing the blade stiffness to intro- 
duce more blade live twist** could signi- 
ficantly reduce the large retreating blade 
control loads. Too much live twist, how- 
ever, may increase the control loads by 
introducing a large nose -down advancing 
blade torsional moment. Still, signifi- 
cant control load reductions can be 
achieved and the flight envelope can be 
expanded by increasing live twist to reduce 
retreating blade stall loads, but not 
enough to greatly increase advancing blade 
loads. 

Introduction 

For any practical' helicopter design, 
the level-flight, steady-state loads should 
be below the endurance limit (infinite life 
load) so that sufficient life will be avail- 
able to absorb the larger maneuver loads . 
A major design objective is to produce an 
aircraft with a flight envelope limited by 
aircraft power and not by structural limits 
Frequently, however, the operational flight 
envelope is limited by a rapid growth of 
stall- induced control loads that exceed the 
endurance limit. Therefore, the flight 
envelope is limited by control loads, and 
the available power cannot be fully util- 
ized. 



* Work performed under Contract DAAJ02- 
72-C-0093, Investigation of Torsional 



Natural Frequency on Stall-Induced 
Dynamic L o ading , by The Boeing Vertol 
Company, U. S. Army Air Mobility Researc 
and Development Laboratory (USAAMRDL) . 

** Live twist is the steady and vibratory 
elastic pitch deflection that results 
from blade torsional loads. 



The rapid control load growth is at- 
tributed to stall flutter which is a 
consequence of high angles of attack and 
resulting blade stall. Visual confirmation 
of the large stall loads can be found in 
pitchlink or blade torsional gage waveforms 
on which characteristic stall spikes appear 
in the fourth quadrant of the blade azimuth. 
These high loads result from an aeroelastic, 
self-excited pitch motion in conjunction 
with repeated submersion of a large portion 
of the rotor blade in and out of stall. 

An aeroelastic rotor analysis program^ 
was developed, using unsteady aerodynamic 
theory that could preduct the large stall - 
induced control loads. Limited analytical 
studies of a single flight condition, using 
this program 2 and another study by Sikorsky 
Aircraft?, indicated that modifications to 
the blade torsional properties could 
significantly reduce the stall-induced 
control loads. These encouraging theoreti- 
cal results led to a model test3 to verify 
the control load reduction. The test re- 
sults showed that, by reducing the blade 
torsional natural frequency from 5.65 to 
3 per rev, the model stall flutter torsion 
spike was reduced 73 percent, giving a 
first verification of the analytical trend. 



Next, an extensive study was under- 
taken to explore the impact of modified 
blade torsional properties on blade tor- 
sional loads over the flight envelope. The 
study had two major parts- -the first part 
compared model test results of blades with 
different torsional properties with analy- 
tical results to evaluate the analysis; 
while the second part analytically explored 
the variation of control loads for flight 
conditions of hover and 125, 150, and 175 
knots with blade loadings (C T /a) from 0.05 
to 0.18. This paper summarizes the results 
of this study. 

Theory and Test Comparison 

For useful analytical results, confi- 
dence in the theory must be established to 
show that the fundamental phenomena are 
properly accounted for. The aeroelastic 
rotor analysis has been successfully cor- 
related with control loads obtained from 
full-scale CH-47C flight data for both 
stalled and unstalled conditions. Addi- 
tional correlation with the model rotors 



115 



test was performed to further evaluate the 
analysis. 

The model test used three six-foot 
diameter rotor sets. Each rotor set had 
three articulated blades with identical 
airfoil and planform, but each set had a 
different torsional natural frequency. The 
first set of blades had a torsional natural 
frequency of 4.25 per rev and was con- 
structed of fiberglass, using conventional 
crossply torsion wrap. The second set of 
blades had mass properties similar to the 
first blade set, but had a torsional' nat- 
ural frequency of 3.0 per rev. These 
blades were constructed with fiberglass, 
using a uniply torsion wrap which substan- 
tially reduced the blade torsional 
stiffness. The third set of blades had a 
torsional natural frequency of 5.65 per 
rev and was constructed of carbon composite. 
Although the carbon blades were not signi- 
ficantly stiffer than the first set of 
blades, they had significantly lower tor- 
sional inertia which accounted for the 
higher torsional natural frequency. 

A number of runs were made for each 
rotor set at full-scale tip speeds and an 
advance ratio of 0.3. Due to the differ- 
ence in torsional properties, the blade 
live twist of each rotor set was different, 
resulting in propulsive force and thrust 
differences for identical collective, 
cyclic, and shaft angle. One run for each 
rotor set was selected such that the rotors 
would have similar propulsive force varia- 
tions with thrust. From each of these runs, 
five test points were selected which covered 
the range of available blade loading (Cx/a) 
and which provided at least one flight con- 
dition below stall, one condition in 
transition, and two stalled conditions. A 
detailed description of the test conditions 
and all the model blade physical properties 
are given in Reference 4. 

The variation of test and analytical 
blade torsion amplitude with blade loading 
(Cf/a) is shown in Figures 1, 2, and 3 for 
the low stiffness, standard reference, and 
carbon blades, respectively. Each blade 
was instrumented to record blade torsion 
data. Due to gauge failures, only two 
gauges on the standard reference blades and 
only one gauge on the carbon blades were 
operational. In general, the analysis 
correctly preducts the trend of blade tor- 
sional load amplitude with blade loading 
for both stalled and unstalled flight 
conditions. The preducted stall inception 
agrees well with test for the low stiffness 
blade. For the standard reference blade 
and the carbon blade, stall inception is 
predicted about 0.01 Ct/ct too early (see 
Figures 2 and 3) . The analysis predicts 
approximately the correct torsional load 
growth rate in stall; but, because the 
stall inception is predicted early, the 



« 16 



















• 










• 










6 

• 










1 




J * * 


xU 


• ANALYSIS 

, RANGE OF TEST 








I DATA FOR 3 
1 INSTRUMENTED 
1 BLADES 

O TYPICAL TEST 
VALUE 




0.04 



C T /o - BLADE LOADING 



Figure 1 . Comparison of Measured and Calculated Blade Torsion Amplitude for 
the Low Stiffness Blade (3 per rev). 



28 








_ ______ 


ft 1 

I 


24 


































• 














• I 




EQ 








1 




z 








[ 




z 
o 
w 16 

EC 

o 

h- 
a 

z 

r- 
< 

1 12 

LLI 

1- 

< 








1 










I 










I 




Q 
< 

co 8 






• 








• 


,I lX 








i 


i 




C ANALYSIS 


4 








T RANGE OF TEST 








[ INSTf 
BLAD 


1UMENTED 
ES 



0.O4 0.06 0.08 0.10 0.12 0.14 

C T /o - BLADE LOADING 

Figure 2. Comparison of Measured and Calculated Blade Torsion Amplitude for 
the Standard Blade (4.25 per rev). 



116 









w — 










• 








• 


°; 












, 














n , 


» / 

• 


ANALYSIS 




" 








O 


TEST VALUE F 
NSTRUMENTS 
INSTRUMENT 
ON OTHER BL 
HASFAILEDI 


OR ONE 
D BLADE 
ATION 
ADES 

■ 



C T /a- BLADE LOADING 

Figure 3. Comparison of Measured and Calculated Blade Torsion Amplitude 
for the Carbon Blade (5.65 per rev). 

stall loads are overpredicted for the two 
stiff er blades. 

A number of possible explanations for 
the analytical overprediction are discussed 
in Reference 4. The two most likely expla- 
nations are first, questionable model blade 
physical properties (possible 0.02 chord 
error in the carbon blade center of gravity, 
0.06 chord error in the standard blade 
shear center, and -0.07 chord error in the 
carbon blade shear center) secondly, the 
unsteady aerodynamic plunging representa- 
tion is inadequate (a lack of cyclic pitch 
led to model blade flapping of 12 , the 
theory may be overly sensitive to these 
large plunging velocities) . 

Even though correlation is not as good 
as desired, it is clear that the analysis 
predicts the large stall-induced control 
load increase (stall flutter) , approximates 
the control load increase with increasing 
rotor thrust, and defines the proper load 
trend for changes in blade torsional prop- 
erties. These results are sufficient to 
provide a degree of confidence in the 
theoretical trends, indicating that the . 
qualitative results of the flight envelope 
investigation are meaningful. 

Analytical Investigation of 
Stall -Induced Dynamic Loading 

The analysis and test of Reference 2 
showed that changes in blade torsional prop- 



erty can change the stall -induced control 
loads. For this discovery to have a prac- 
tical application, it must be shown that 
realistic changes in the blade torsional 
properties will lead to a significant 
reduction in the large stall-induced con- 
trol loads throughout the flight envelope. 
To determine the variation of control loads 
over the flight envelope, an extensive 
analytical study was performed. The air- 
craft used for this study was a single- 
rotor helicopter with CH-47C blades. A 
20.1 square-foot frontal area was assumed 
for the fuselage and a tail rotor suffi- 
cient for trim purposes was added. The 
main rotor consisted of three articulated 
30- foot radius blades with a constant chord 
of 25.25 inches. The blade cross-section 
is a cambered 23010 airfoil with 9.137 
degrees of linear twist along the blade 
span. 

To perform this study, blade torsional 
properties were modified by changing blade 
torsional stiffness (GJ) , changing control 
system stiffness, and changing blade pitch 
inertia. Using the CH-47C blade's tor- 
sional natural frequency of 5.2 per rev as 
a baseline, the frequencies were adjusted 
from 3 per rev to 7 per rev. 

Natural frequencies of 3 rev, 4 rev, 
and 7 rev were obtained by multiplying the 
torsional stiffness distributions by 0.25, 
0.5 and 3.3, respectively. A control 
system stiffness of 1650 pounds per inch 
generated a frequency of 3 per rev; 11,850 
pounds per inch was used for the basic 
blade, and an infinite stiffness produced 
a 6 per rev frequency. Pitch inertia 
changes resulted in blades with frequencies 
of 3, 4 and 7 per rev due to scaling the 
pitch inertia by factors of 3.08, 1.7 and 
0.55, respectively. All changes were made 
by only varying the blade properties in- 
dicated, while holding all other parameters 
at the nominal CH-4 7C values. 

Figure 4 shows the relationship be- 
tween pitch-link load amplitude and 
torsional frequency (3 per rev to 7 per 
rev) at an airspeed of 125 knots and a 
blade loading of 0.115 for the three meth- 
ods of varying frequency. Each method 
produced approximately the same trend of 
increasing pitch-link loads with increasing 
torsional natural frequency. Figure 5 
illustrates the variation of pitchlink load 
amplitude with natural frequency at 150 
knots airspeed and blade loadings of 0.115. 
For both airspeeds, the variation in tor- 
sional stiffness leads to larger changes 
in pitch- link loads, than do changes in 
control stiffness or pitch inertia. 

Since blade torsional stiffness changes 
resulted in the largest change in control 
load and the lowest loads, the effect of 
blade torsional stiffness changes will be 



117 



* 2000 
1 



BLADE LOADING = 0.115 



TORSIONAL STIFFNESS VARIATION 



PITCH INERTIA VARIATION 




CONTROL SYSTEM STIFFNESS 
VARIATION 



2 3 4 5 6 7 

TORSIONAL NATURAL FREQUENCY - PER REV 

Figure 4. Variation of Pitch-Link Load Amplitude with Natural Frequency 
at 125 Knots. 



6000 



5000 



o 

I- 




3000 



1000 



3 4 5 6 7 

TORSIONAL NATURAL FREQUENCY - PER REV 

Figure 5. Variation of Pitch-Link Load Amplitude with Natural Frequency 
at 150 Knots. 

explored in greater depth. It is apparent 
from these results that it is not the 
reduced torsional frequency alone that 
reduces pitch-link loads, but also the in- 
crease in blade live twist, resulting from 
reduced torsional stiffness. 

Four blades with different torsional 
natural frequencies (i.e., 3 per rev, 4 
per rev, 5.17 per rev and 7 per rev) were 
analyzed for 24 flight conditions to in- 
vestigate the interactive effects of 



torsional stiffness, blade loading (Ct/ct) , 
and airspeed. The airspeeds ranged from 
hover to 175 knots and blade loading from 
0.05 to 0.018. 

Hover 



Figure 6 shows the variation of pitch - 
link load amplitude with blade loading 
for four sets of rotor blades. One degree 
of cyclic pitch was used to provide some 
means of introducing a cyclic load varia- 
tion. If this were not done, the analysis 
would predict only steady loads. At blade 
loadings of 0.115 and 0.12 the pitch-link 
load has a 1-per-rev waveform with an amp- 
litude of about 100 pounds for all four 
blades. These loads represent an unstalled 
condition, and there is virtually no load 
variation with blade loading or torsional 
natural frequency. At a blade loading of 
0.15, the loads increase to between 200 
pounds and 300 pounds, with the 3 per rev 
and 4 per rev blades having the lowest 
load. At this condition, the rotor power 
is around 4000 horsepower which is well 
beyond the available rotor power. 

At a blade loading of 0.165, the pitch- 
link load for the 3 per rev blade increases 
sharply to 1000 pounds. The major portion 
of this load is a 950 pound, 8-per-rev 
component. Since the blade torsional 
natural frequency is 3 per rev, it was 
surprising to observe that there was little 
3-per-rev load and a very large 8-per-rev 
load. Further examination revealed that 
the blade second torsional natural frequency 
is almost exactly 8 per rev, explaining the 
source of the large load. It is not known 
why the torsionally soft blade prefers to 
oscillate in its second mode. Further 
investigation is necessary. 




Cjh - BLADE LOADING 

Figure 6. Variation of Pitch-Link Load Amplitude with Blade Loading in Hover 
for One Degree of Cyclic Pitch. 



118 



The 4-per-rev blade, at the same 
flight condition, has a pitch -link load of 
340 pounds which is the lowest of the four 
blades. The 5.2-per-rev and 7-per-rev 
blades had approximately the same load at 
about 400 pounds. The required rotor power 
for all blades is approximately 5000 horse- 
power which is 66 -percent more than the 
available rotor power of a CH-47C rotor. 
Since the required power is so high, res- 
ults at this (0.165) and higher blade 
loadings probably have no practical appli- 
cation. 

At a blade loading of 0.18, the 3 per 
rev blade pitch-link load increases to 
4500 pounds, with the 8-per-rev component 
again providing the largest load. The 4- 
per-rev blade also shows a large load in- 
crease, reaching a load of about 4000 
pounds. However, this blade's large tor- 
sional loads occurred at the first torsional 
natural frequency (3600 pounds at 4 per 
rev). The 5.2-per-rev blade has a load of 
650 pounds and the 7-per-rev blade load is 
540 pounds. At this condition, the loads 
reduce with increasing torsional frequency. 
The required rotor power for this flight 
condition is over twice the available power, 
indicating that rotor stall has reached a 
larger portion of the blade. 

These results indicate that in hover, 
increased torsional frequency (i.e., tor- 
sional stiffness) delays the inception of 
stall flutter. This conclusion generally 
agrees with propeller experience. However, 
the large power required at a blade loading 
of 0.16 (i.e., 50 percent above available 
power) implies that this flight condition 
and higher blade loadings do not apply to 
current aircraft. If current power to 
rotor solidity ratios are therefore used, 
there is very little difference between 
the torsional loads of the four blades up 
to reasonable blade loadings (for this 
discussion, approximately 0.15 blade 
loading) . 

125 Knots 



a large, high-frequency torsional load 
component which generally appears between 
an azimuth position of 270 degrees to 60 
degrees and usually determines the load 
amplitude as shown in Figure 9. The 
stalled pitch-link load continues to 
rise to 2650 pounds at a blade loading 
of 0.11. Increasing the blade loading 
beyond this point results in a load 
reduction. This reversal of the load 
trend may at first appear surprising, but 
it has been observed in model data (see 
Figure 1) and full-scale results (see 
Figure 10) . 

The 7-per-rev blade has generally the 
same pitch-link load trend, with blade 
loading as the basic blade. There is an 
unstalled load region up to a blade loading 
of 0.09 (with a typical waveform given in 
Figure 8) , a stalled load region typified 
by a large load increase with blade loading 
(with a typical stalled waveform at a blade 
loading of 0.10 as shown in Figure 8), and 
a load reversal at a blade loading of 0.11. 
However, as far as control loads are con- 
cerned, the 7-per-rev blade is significantly 
worse than the basic blade. In the un- 
stalled region, the loads are about the 
same; in stall the 7-per-rev blade loads 
are 65-percent larger. Stall inception 
occurs at a blade loading of about 0.095 
which is 0.008 before the basic blade. 

The 4-per-rev blade has a significantly 
different pitch-link load trend with in- 
creasing blade loading than the two blades 

? 



The variations of pitch-link loads 
with blade loading for each of the four 
different torsional frequency blades are 
shown in Figure 7 for an airspeed of 1.25 
knots. The basic blade (with a torsional 
natural frequency of 5.2 per rev) pitch- 
link load increases slowly with increasing 
blade loading up to a value of 0.10. In 
this region, the pitch-link load waveform 
is predominantly 1 per rev (see Figure 8) 
and the loads are classified as unstalled 
(even though some stall is present) . Stall 
inception occurs at a blade loading of 
about 0.103. The stall inception repre- 
sents the flight condition in which the 
control loads begin to exhibit the rapid 
increase, due to blade stall. In this 
region, the pitch- link load waveform has 







0.06 0.07 0.08 0.09 0.10 0.11 0.12 

C T /o - BLADE LOADING 

Figure 7. Variation of Pitch-Link Load Amplitude with Blade Loading at 125 Knots. 



119 



AIRSPEED 125 KNOTS 
BLADE LOADING 



TORSIONAL 
FREQUENCY 



Cf/o = 0.09 



4/REV 1, 



5.2/REV " 




270 360 

BLADE AZIMUTH - DEG 



270 360 

BLADE AZIMUTH - DEG 



Figure 8. Pitch-Link Load Waveforms for 125 Knots, at Blade Loadings of 
0.09 and 0.10. 



AIRSPEED 125 KNOTS 



4 






TORSIONAL 3 
FREQUENCY 


C T /o = 


.12 


2 


\ 


A / 


3/REV 


% 


J\ 


■W 


r v 


-2 


\y 




-3 


■ 




-4 






-5 






-6 


■ 




-7 


L 





4/REV < 



Cy/o».12 







90 ISO 270 360 

BLADE AZIMUTH - DEG 



90 180 270 

BLADE AZIMUTH - DEG 



BORON 
BLADE 

6.B3/REV 




GLASS BLADE 
5.45/REV 



Figure 9. Pitch-Link Load Waveforms for 125-Knots at High Blade Loadings. 



0.07 0.08 0.09 0.10 0.11 0.12 

NONDIMENSIONAL BLADE LOADING - C T /o 
(IN WIND AXIS SYSTEM) 

Figure 10. CH-47C Advanced-Geometry Blade Flight Test Data at an Advance 
Ratio of 0.2. 

previously discussed. While there is the 
typical unstalled region with little load 
growth up to a blade loading of 0.09 (a 
typical unstalled waveform is given in 
Figure 8), there is an irregular, but mod- 
erate, load growth between 0.09 and 0.12. 
At a blade loading of 0.12, the torsion load 
is only 1641 pounds and the waveform just 
attains a fully stalled characteristic (see 
Figure 9) . There is a large load increase 
(1360 in-lbs) as the blade loading increases 
from 0.12 to 0.13. Examining the 0.13 
pitch-link load waveform (see Figure 9) 
shows that the large load is not caused by 
stall flutter; instead, it is due to a 
large nose -down moment generated by the 
advancing blade combined with moderate 
stall spikes. 

The 3-per-rev blade has a load trend 
similar to the 4-per-rev blade. The un- 
stalled load region extends to a blade 
loading of 0.115, and a typical unstalled 
waveform is given in Figure 8. Even at a 
blade loading of 0.12, the pitch-link load 
waveform does not show a fully stalled 
waveform (see Figure 9). However, at 
0.125, the pitch-link load increases by 
150 percent to 5273 pounds, by far the 
largest load of any blade. The waveform 
at 0.125 (see Figure 9) shows large stall 
spikes with an amplitude of 2500 pounds; 
however, the large load increase is due to 



120 



a 5000 pound load at 90° azimuth which re- 
sults from a large nose -down pitching 
moment. The load growth is so large at 
this condition that it may represent the 
lower boundary of an instability. 

It is clear from these results that 
stall inception occurs earlier as the 
blade torsional frequency is increased. 
Further, the maximum retreating blade 
stall-induced pitch-link loads are larger 
for blades with higher torsional frequen- 
cies (i.e., the 7-per-rev blade has the 
largest stall-induced loads) . 

The CH-47C flight test data substan- 
tiating the conclusion that stall inception 
occurs at higher blade loadings as blade 
frequency is decreased. Figure 10 shows 
the results of a CH-47C advanced- geometry 
blade flight test for aft rotor blades with 
a boron filament spar and a fiberglass spar 
at an advance ratio of 0.2. The glass 
blade has a stall inception delay of 0.0085, 
due to reducing the torsional frequency 
from 6.53 per rev to 5.45 per rev. The 
single-rotor study results show a stall 
delay of 0.008 for reducing the torsional 
frequency from 7 per rev to 5 . 2 per rev at 
an advance ratio of 0.3. 

The large advancing blade loads ex- 
perienced by the 3-per-rev and 4-per-rev 
blades, beyond a blade loading of 0.12, 
are not due to the stall -flutter phenomenon 
which results from retreating blade stall 
and unstall cycles. This load is associated 
with the negative lift on the advancing 
blade tip and appears to be a divergence- 
like phenomenon. Large, negative tip lift 
causes the blade to bend tip down; the 
high tip drag coupled with the flap de- 
flection causes a nose -down moment. The 
moment causes elastic nose -down pitch which 
leads to more negative lift, resulting 
eventually in even larger loads. 

150 Knots 



At 150 knots ( see Figure 11) , the 
basic CH-47C blade has an unstalled load 
region up to about 0.08 which is typified 
by a slow increase of pitch-link load with 
blade loading and a predominantly 1-per- 
rev waveform. In the region between 0.08 
and 0.1, a different load trend is observed. 
The load increases gradually from 1500 to 
2000 pounds, but at a faster rate than the 
sub-stall load growth, and the waveforms 
show significant evidence of stall spikes 
for the retreating blade (see Figure 12) . 
Stall inception (i.e., rapid load growth) 
appears to occur around 0.10 3, reaching 
maximum load near 0.11 (see Figure 12). 
The load drops at a blade loading of 0.115, 
showing a load reversal as observed at 125 
knots . 




POWER 
LIMIT 



0.05 0.06 0.07 0.08 0.09 0.10 0.11 0.12 

C T /o- BLADE LOADING 

Figure 1 1 . Variation of Pitch-Link Load Amplitude with Blade Loading at 
150 Knots. 



TORSIONAL 
FREQUENCY 



AIRSPEED 150 KNOTS 






90 180 270 360 

BLADE AZIMUTH - DEG 




90 180 270 360 

BLADE AZIMUTH - DEG 



Figure 12. Pitch-Link Load Waveforms at 150 Knots for Blade Loadings of 
0.09 to 0.11. 



121 



The 7-per-rev blade shows a similar 
load trend with blade loading as the basic 
blade, but with significantly larger stall 
loads. The unstall loads occur up to 0.07. 
Stall inception occurs at approximately 
0.075, reaching a fully stalled waveform 
at 0.09 (see Figure 12). The loads level 
out at 0.10, reach a second stall inception 
near 0.105, and the load begins to grow 
again. Examining the pitch- link load wave- 
form at 0.11 (see Figure 12) shows that the 
load increase is due to a large stall spike 
occurring in the to 50-degree azimuth re- 
gion, not to retreating blade stall spikes. 

With an expected unstalled waveform 
the 4-per-rev blade has a typical unstalled 
control load growth up to a blade loading 
of about 0.09. Between 0.09 and 0.115, 
there is an irregular load growth. In this 
region, the waveforms show evidence of 
retreating blade stall (see Figure 12) , but 
no large load increase. At a blade load- 
ing of 0.11, the torsion load is 2100 
pounds and the waveform just attains a 
fully stalled characteristic (see Figure 
12). There is a 1230-pound load increase 
as the blade loading increases from 0.115 
to 0.12. Examining the 0.12 pitch-link 
load waveform (see Figure 13) shows that 
the large load increase is caused by a 
large, advancing-blade nose -down spike 
combined with retreating blade stall spikes. 

The 3-per-rev blade shows a reasonable 
pitch-link load through a blade loading of 
0.11. However, at 0.115, the blade is 
apparently unstable since the loads have 
grown so large that the blade would proba- 
bly fail. The pitch-link load waveform 
at 0.11 (see Figure 13) contains relatively 



TORSIONAL 
FREQUENCY 



AIRSPEED 130 KNOTS 
BLADE LOADING 
■0.11 Ct/o-0.12 




small retreating-blade, stall-induced 
spikes. There is, however, a large com- 
pression load for the advancing blade at 
90 degrees blade azimuth. By examining 
the pitch link load waveform for the un- 
stable flight condition, it appears that 
the biade divergence involves a large ad- 
vancing blade compression load that con- 
tinually increases with each rotor 
revolution." 

The 3-per-rev blade is experiencing 
an additional problem which is not apparent 
by simply observing the load trend. For 
all the load conditions calculated at 150 
knots, the required power exceeds the 
available power. Apparently, the blade is 
experiencing so much live twist that there 
is a significant increase in rotor drag. 
The other blades, by contrast, exceed the 
available power only at a blade loading of 
0.115. It is, therefore, obvious that the 
3-per-rev blade is not an acceptable con- 
figuration for the 150-knot flight condi- 
tion. 

175 Knots 

At 175 knots (see Figure 14), the 
basic blade pitch-link load trend shows 
unstalled loads continuing to a blade 
loading of 0.07 and stall inception occur- 
ring about 0.075. The stalled load 
increases with a moderate growth rate up 
to 0.09. Figure 15 illustrates the pitch- 
link load waveform at 0.09, showing the 
retreating blade stall spikes and a large 
nose-down load at 90 degrees azimuth. Be- 
yond a blade loading of 0.09, the load 
does not reverse as it does for previous 
airspeeds (even though the retreating 
blade stall spike is significantly reduced 
at a blade loading of 0.11 as shown in 
Figure 15). Instead, the load continues 
to increase at about one half the previous 
growth rate, due to an increasingly large 
nose-down load at 90 degrees azimuth. 



The 7 -per 
trend is almost 
blade trend up 
shows , the wave 
frequency stall 
shows a slight 
but then resume 
typical stalled 
growth at 0.11 
large stall spi 
azimuth and an 
at 90 degrees a 



rev blade pitch-link load 
identical with the basic 
to 0.09. As Figure 15 
form exhibits typical high- 
spikes. The torsion load 
load reversal beyond 0.09, 
s the load increase at the 

load growth rate. The load 
is due to a combination of a 
ke at around 30 degrees 
increasing nose -down load 
zimuth (see Figure 15) . 



90 180 270 360 
BLADE AZIMUTH - DEG 



90 180 270 360 

BLADE AZIMUTH - DEG 



Figure 13. Pitch-Link Load Waveforms at 180 Knots for the 3-per-rev and 
' 4-per-rev Blades at High Blade Loadings. 



The 4-per-rev blade has a typical 
substall load growth up to 0.07 and gener- 
ally follows the load trend of the 7-per- 
rev blade and the basic blade up to 0.08. 
Beyond this point, the load growth rate 
drops significantly. At a blade loading of 
0.09, the pitch link load is 500 pounds 



122 




TORSIONAL 
FREQUENCY 



AIRSPEED 175 KNOTS 

BLADE LOADING 

0,7s -0.09 Of/a -0.11 



POWER 
LIMIT 



0.05 0.06 0.07 0.08 0.09 0.10 0.11 0.12 

Cjh - BLADE LOADING 

Figure 14. Variation of Pitch-Link Load Amplitude with Blade Loading at 
1 75 Knots. 

below the other two blades. Also 0.09, 
the pitch-link load waveform shows little 
evidence of retreating blade stall (see 
Figure 15) , but does show that the major 
component of the load results from a nose- 
down moment at 90 degrees blade azimuth. 
Beyond this point, the load growth rate in- 
creases sharply from 2500 pounds at 0.09 to 

4200 pounds at 0.11. 

The 3-per-rev blade is not seriously 
considered at this airspeed. The loads 
are 1500 pounds beyond any of the other 
blades, and an advancing blade instability 
is apparent at a blade loading of 0.09. 
Further, the required rotor power exceeds 
the available CH-47C power for all 175- 
knot flight conditions examined. 

Examining the 175-knot pitch-link load 
waveforms at a blade loading of 0.11 
clearly shows that all four blades experi- 
ence increased advancing-blade compression 
loads when compared to the 15 -knot wave- 
forms (compare Figures 12 and 14) . The 
7-per-rev blade shows a 1'300-pound advan- 
cing-blade load increase for the 25-knot 
airspeed increase. The basic 5.2-per-rev 
blade load increase is 2200 pounds, the 
4-per-rev blade load increase is 3500 
pounds, and the 3-per-rev blade has di- 
verged. Therefore, the blades experience 
advancing blade load problems which are 
intensified as airspeed is increased and 
blade torsional stiffness is reduced. 




270 
BLADE AZIMUTH - DEG 



90 180 270 

BLADE AZIMUTH - DEG 



Figure 1 5. Pitch-Link Load Waveforms for 1 75 Knots at Blade Loadings of 
0.09 and 0.11. 

When comparing the 175-knot results 
for the 4-per-rev, 5.2-per-rev and 7-per- 
rev blades, it should be noted that the 
rotor power limit for a single CH-47C rotor 
is reached just beyond a blade loading of 
0.09. For conditions below lthe power 
limit, the 4-per-rev blade is slightly 
better than the others, since the maximum 
pitch-link load is 500 pounds lower. The 
three blades appear to have adequate man- 
euver margin, although the 4-per-rev blade 
may experience larger maneuver loads. 

These results show that a significant 
reduction of the basic blade control loads 
can be realized over a considerable range 
of advance ratios and blade loading and 
that these reductions lead to a significant 
extension of the control load-limiting 
aircraft flight envelope. These results 
can be summarized by obtaining the flight 
condition (as a function of Cj/a and p) 
that first experiences a 2500-pound pitch- 
link load. The 2500-pound load approximates 
the original pitch- link endurance limit load 
for the CH-47C control system. These flight 
conditions lead to a blade loading versus 
advance ratio envelope for the 2500-pound 



123 



pitch-link load or an endurance -limit 
flight envelope. Figure 16 compares the 
endurance limit flight envelopes for each 
of the four different frequency blades 
investigated. 

As Figure 16_ shows , the blade with a 
torsional natural frequency of 4 per rev 
(dashed line) has the best flight envelope 
and represents a significant improvement 
: over the basic blade configuration. The 
4-per-rev flight envelope has essentially 
the same shape as the basic blade, but it 
occurs at a higher blade loading. The 
basic blade envelope occurs at a blade 
loading of 0.016 below the 4-per-rev blade 

at an advance ratio of 0.29 . At an. 

advance ratio of 0.38, the basic and 4-per- 
rev blades are approximately equal; but at 
an advance ratio of 0.4, the 4-per-rev 
blade envelope is expanded beyond the basic 
blade by a blade loading of 0.005. 

The 3-per-rev blade (short dashes) 
shows a different flight envelope. At an 
advance ratio below 0.29, the 3-per-rev 
blade reaches the endurance limit at a 



blade loading of' 0.123. However, the 3- 
per-rev envelope drops sharply with 
increasing advance ratio and eventually 
falls below the three other blades at a 
0.375 advance ratio. The sharp boundary 
reduction of this blade at the higher ad- 
vance ratios is due to the large advancing - 
blade load growth which eventually becomes • 
an instability.* These instabilities show 
that the 3-per-rev blade is clearly un- 
acceptable, at least for the current pitch- 
link-controlled configuration. 

The 7-per-rev blade clearly has the 
poorest flight envelope up to an advance 
ratio of about 0.37. At the higher advance 
ratios above 0.35, the 7-per-rev blade has 
the smallest reduction of blade -loading 
capability with increasing advance ratio. 
In this region, the 7-per-rev blade sur- 
passes the 3-per-rev blade at an advance 
ratio of 0.37, surpasses the basic blade 
at 0.40, and will probably surpass the 
4-per-rev blade around 0.44. Therefore, 
a torsionally stiff blade may be required 
to attain a reasonable flight envelope 
beyond advance ratios of 0.44. 



o.i3r 



0.12 - 



0.11 - 



9 o.io 
q 



5 



0.09 - 



0.08 - 



0.07 



0.06 - 



0.2 



K 




\ 


4.04/REV 


3/REV-'*\ 


\ PITCH LINK LOAD 


\ 


> ABOVE 2500 LB 


\ 

\ 


\ 


5.165/REV 


\ \ 




ON V 




■^\«v \ 




•\l 


~A^ 


\ X 


f\ 


9 "ft 


6.97/REV 


\ l\ 












\ * 


PITCH LINK LOAD 


\ 


BELOW 2500 LB 


\ 






\ 




V 


' 


* 


NATURAL FREQUENCY 


GJ SCALING FACTOR 


3.0 


.25 


4.0 


.50 


5.2 


1.0 


7.0 

., „ ,_!_„, L_ 


3.3 

i L_ i 



0.3 
ADVANCE RATIO - (i 



0.4 



Figure 16. Control Load Endurance Limit Boundaries for Blades with Torsional 
Natural Frequencies of 3, 4, 5.2 and 7 Per Rev. 



Conclusions and Recommendations 

The results of the theory-test com- 
parison performed for the 6- foot-diameter 
model blades and the study of varying 
torsional properties for the full-scale 
CH47C size blades have lead to: 

1. The theory-test comparison with the 
6 -foot-diameter model data indicates that 
the aerolastic rotor analysis reasonably 
represents the large stall induced control 
loads, the, control load change with blade 
loading, and the load variation with 
changes in blade torsional properties. 
Therefore, the analytical study of the 
CH47C size blades should provide at least 
a qualitative evaluation of the control 
load variation. 

2. Changes in control system stiffness, 
pitch inertia, and blade torsional stiff- 
ness vary the large, stall-induced control 
loads. However, the control load change 
is not a simple function of torsional 3 
natural frequency as previously suspected, 
since torsional frequency changes, due to 
varying the blade torsional stiffness, 
produce control load changes larger than 

* It may not be possible for an actual 
rotor to experience blade divergence. 
Before large divergence loads result, 
there" is a significant increase in re- 
quire^ rotor power. Therefore, a real 
rotor may simply run out of power and be 
unable to attain a flight condition for 
which divergence would occur. 



124 



other methods of changing torsional fre- 
quency. 

3. A blade with a torsional natural 
frequency of 4 per rev represents a com- 
promise between significantly reducing 
stall flutter' loads, while allowing 
moderate increases in the advancing blade 
loads at high speeds. This compromise 1. 
provides the best endurance limit flight 
envelope up to an advance ratio of 0.45. 
Beyond this advance ratio it appears that 

a torsionally stiff blade will provide a 
better endurance limit flight envelope. 2. 

4. Additional work is required in the 
following areas. 

• A model test program is needed to 
validate the analytical results 

over a wide range of flight con- 3. 
ditions with remote collective and 
cyclic pitch to insure trimmed 
flight. 

• Theory improvements are needed to 
eliminate deficiencies discovered 4. 
in the theory-test comparison. 

• Continue analytical studies to in- 
vestigate mechanisms of the load 
generation, .maneuver and high- 



speed load trends and other means 
for expanding the endurance limit 
flight envelope. 



References 



F. J. Tarzanin, Jr., PREDICTION OF 
CONTROL LOADS DUE TO BLADE STALL, 
27th Annual National V/STOL Forum of 
the AHS, Preprint No. 513, May 1971. 

F. J. Tarzanin, Jr. and R. Gabel, 
BLADE TORSIONAL TUNING TO MANAGE 
ROTOR STALL FLUTTER, Presented at the 
A1AA 2nd Atmospheric Flight Mechanics 
Conference, AIAA Paper No. 72-958, 
September 1972. 

F. 0. Carta, L. M. Casellini, P. J. 
Arcidiacono, H. L. Elman, ANALYTICAL 
STUDY OF HELICOPTER ROTOR STALL FLUTTER, 
26th Annual Forum of the AHS, June 1970. 
the AHS, June 1970. 

F. J. Tarzanin, Jr. and J. Ranieri, 
INVESTIGATION OF TORSIONAL NATURAL 
FREQUENCY ON STALL- INDUCED DYNAMIC 
LOADING. Performed under contract 
DAAJ02-72-C-0092, USAAVLABS TR- 
(Not yet released) . 



125 



APPLICATION TO ROTARY WINGS OF A SIMPLIFIED AERODYNAMIC LIFTING 
SURFACE THEORY FOR UNSTEADY COMPRESSIBLE FLOW 

B. M. Rao* and W. P. Jones** 

Department of Aerospace Engineering 

Texas ASM University, College Station, Texas 



Abstract 

In a recent paper, Jones and Moore have deve- 
loped a simple numerical lifting surface technique 
for calculating the aerodynamic coefficients on 
oscillating wings in subsonic flight. The method 
is based on the use of the full lifting surface 
theory and is not restricted in any way as to fre- 
quency, mode of oscillation or aspect ratio when 
M < 1. In this study, this simple but general met- 
hod of predicting airloads is applied to helicopter 
rotor blades on a full three-dimensional basis. 
/The general theory is developed for a rotor blade 
at the <|j = tt/2 position where flutter is most lik- 
ely to occur. Calculations of aerodynamic coeffi- 
cients for use in flutter analysis are made for 
forward and hovering flight with low inflow for 
Mach numbers and 0.8 and frequency ratios p/H=l • 
and 4. The results are compared with values given 
by two-dimensional strip theory for a rigid rotor 
hinged at its root. The comparisons indicate the 
inadequacies of strip theory for airload predicti- 
on. One important conclusion drawn from this stu- 
dy is that the curved wake has a substantial effect 
on the chordwise load distribution. The pitching 
moment aerodynamic coefficients differ appreciably, 
from the results given by strip theory. 

Introduction 

In a survey paper, Ref. 1, Jones' et al. give 
a detailed account of significant developments in 
the field of unsteady aerodynamics of helicopter 
rotor blades. One of the problem areas surveyed 
was that of blade flutter as it has been found that 
under certain operating conditions, rotor blades 
can flutter in both hovering and forward flight. 
This phenomenon has been investigated by several 
researchers in Refs. 2, 3, 4, and 5 and the results 
of their studies have improved our understanding 
of the problem. For the case of hovering flight, 
J. P. Jones in Ref. 2 applied a method developed 
by W. P. Jones in Ref. 6 to derive the approximate 
aerodynamic coefficients for an oscillating single 
rotor blade for use in his flutter analysis. He 
approximated the actual flow conditions by neglect- 
ing curvature effects and assuming a simple two- 
dimensional mathematical model cosisting of a ref- 
erence blade and an infinite number of wakes lying 
beneath the reference blade extending from -°° to °°. 
He considered flapping and pitching motions and com- 
pared his results with those obtained experimenta- 
lly by Daughaday and Kline in Ref. 3. On the basis 

Presented at the AHS/NASA-Ames Specialists' Meet- 
ing on Rotorcraft Dynamics, February 13-15, 1974. 
The funds for computation were provided by the 
U. S. Army Research Office, Durham. 

* Associate Professor 

** Distinguished Professor 



127 



of this work it was concluded that the wake is pri- 
marily responsible for some of the vibratory pheno- 
mena found on helicopters in practice. For low 
inflow conditions, Loewy in Ref. 4 used a similar 
mathematical model to that of J. P. Jones and inve- 
stigated the variation in the pitching moment damp- 
ing coefficient of a particular blade section as 
p/fl varied for specified positions of axis of osci- 
llation and a range of values of wake spacing. He 
found that the damping coefficient became negative 
whenever p/S was slightly greater than an integer 
for axis of oscillation forward of quarter-chord. 
Similarly he found that the damping coefficient 
for a flapping oscillation dropped sharply at inte- 
gral values of p/S2 but did not actually become neg- 
ative. Tinman and Van de Vooren in Ref. 5, on the 
other hand, assumed that there was no inflow thro- 
ugh the rotor disk and developed a theory for cal- 
culating the aerodynamic forces on a blade rotating 
through its own wake. Their results agree with 
those obtained in Refs. 2 and 4 in the limit when 
zero spacing between the wakes is assumed. All 
this theoretical work confirms the conclusion that 
the proximity of the wake is a contributing factor 
to rotor blade flutter. 

All the theoretical work described above is 
based on the assumption that the flow is incompre- 
ssible. However, with the advent of helicopters 
capable of flying with blade tip speeds ranging up 
to and in excess of the speed of sound, compressi- 
bility effects need to be taken into account when 
determining coefficients for use in flutter analy- 
sis. Jones and Rao in Ref . 7 were able to do this 
on the basis of two-dimensional theory and have 
derived coefficients for a range of Mach numbers, 
reduced frequencies, and wake spacing. Their ana- 
lysis is based on the use of Loewy's model, Ref. 4, 
of the helical wake and the application of a theory 
developed earlier by Jones in Ref. 8 for an oscilla- 
ting airfoil in compressible flow. The values of 
the coefficients given in Ref. 7 agree with those 
obtained in Refs. 2 and 4 for zero Mach number but 
differ appreciably when the Mach number is varied. 
Hammond in Ref. 9 also developed a theory for det- 
ermining compressibilty effects by using a differe- 
nt model of flow from that used in Ref. 7. In his 
model, the wake of the qth blade of a Q bladed 
rotor after n revolutions extends from -2ir(n+q/Q) 
to »; in Jones and Rao's model it extends from -°° 
to o°. His aerodynamic coefficients for several 
Mach numbers and inflow ratios are in general agr- 
eement with "the results of Jones and Rao in Ref. 7. 

While the aerodynamic derivatives predicted 
by two-dimensional strip theory are widely used in 
predicting the flutter speeds of helicopter rotor 
blades, the method does not allow for curvature 
and finite aspect ratio effects. For incompressi- 
ble flow, Ashley, Moser, and Dugundji in Ref. 10 
developed a three-dimensional model in which they 



modified Reissner's theory, Ref. 11, for oscillat- 
ing wings in rectilinear flow by including the free 
stream- velocity variations along the span. Their 
results indicate a negligible difference between 
two and three-dimensional solutions up to 95% of 
the span. Jones and Rao in Ref. 12 similarly stu- 
died tip vortex effects in compressible flow and 
they also concluded that such effects are negligi- 
ble except in regions close to the tip. In some 
of his earlier work, Miller in Refs. 13, 14, and 
15, developed a helical wake model in which the 
rotor wake was divided into a "near" wake and a 
"far" wake. The near wake included the portion 
attached to the blade that extend approximately 
one-quarter of a revolution from the blade trailing 
edge. The effects of the near wake include an in- 
duced chordwise variation in downwash and were for- 
mulated using an adoptation of Loewy's strip theory. 
The chordwise variation in the velocity over the 
airfoil induced by the far wake was neglected. 
Miller extended his model to study the forward fli- 
ght case and found that the nonuniform downwash 
induced at the rotor disk by the wake vortex system 
could account for the higher harmonic airloads en- 
countered on rotor blades in forward flight. He 
also showed that under certain conditions of low 
inflow and low speed transition flight the return- 
ing wake could be sucked up into the leading edge 
of the rotor which would account for some of the 
vibration and noise. Piziali in Ref. 16 has deve- 
loped a"n alternative numerical method in which the 
wake of a rotor blade is represented by discrete 
straight line shed and trailing vortex elements. 
He satisfied the chordwise boundary conditions, but 
the rotor blade was limited to one degree of free- 
dom in flapping. Sadler in Ref. 17, using a model 
similar to Piziali 's, developed a method for predi- 
cting the helicopter wake geometry at a "start up" 
configuration. He represented the wake by a fine 
mesh of transverse and trailing vortices starting 
with the first movement of the rotor blade genera- 
ting a bound vortex, and, to preserve zero total 
vorticity, a corresponding shed vortex in the wake. 
Integrating the mutual interference of the trailing 
and shed vortices upon each other over small inter- 
vals of time, Sadler was able to predict a wake 
geometry. Although his model showed fair agreement 
with the available experimental data for advance 
ratios above one-tenth, Sadler's method is limited 
due to the large computational time required tp 
represent the wake by a finite mesh. 



soning led them to represent the blade motion by a 
series of oscillatory pulses, where each disturba- 

TT TT 

nee occurs over the range, -r- - Aij). <$<■=■+ A$„. 

Corresponding to each burst of oscillation, packets 
of vorticity are assumed to be shed into the wake. 
With increasing forward speed, the spacing between 
the packets of vorticity also increases and it was 
found that the flutter speed became constant when 
y, the advance ratio, was above 0.2. The approach 
used in the present study differs from that adopted 
by Shipman and Wood in that continuous high freque- 
ncy small oscillations are assumed to be superimpo- 
sed on the normal periodic motion of the blade. 
The airloads and aerodynamic derivatives associated 
with the perturbed oscillation of the rotor blade 
can then be calculated by the method described in 
this paper. Since the rotor blade will first att- 
ain its critical speed for classical flutter at 
i|i = it/ 2, the aerodynamic derivatives corresponding 
to this value of ifi only have been calculated. The 
method takes finite aspect ratio and subsonic com- 
pressibility effects fully into account. Typical 
results for a rotor blade hinged at its root desc- 
ribing flapping and twisting oscillations are given 
for a range of Mach numbers and frequency values. 

Basic Equations 

In the development of the analysis of the 
Jones-Moore theory, Ref. 18, for oscillating wings 
in rectilinear flight, the space variables x, y, z, 
and t are replaced by X, Y, Z, and T, respectively, 

s0 that „„ ** *z ^ n ,,. 

x = iX, y = j-, z = — , t = — (1) 

where % is a convenient reference length, U is 

the uniform velocity, M is the Mach number and 

2 1/2 
8 = (1-M ) . The velocity potential of the flow 

around a surface oscillating at a frequency p 

can then be expressed as . . . 

<j>(x,y,z,t) = U»(X,Y,Z)e li " + Wi; (2) 

2 2 
where co = p«7U, X = M co/B . The function * may be 

regarded as a modified velocity potential. Fur- 
thermore, it can be shown that it satisfies the 
wave equation 



2 2 2 

a $ O. , O. 

2 2 2 

9X 8Y Z 3Z 



+ K $ = 



(3) 



where k = Mu/g 



Though many forms of flutter can occur on ro- 
tor blades, attention in this report is concentra- 
ted on the determination of appropriate aerodynamic 
coefficients for use in the analysis of blade flu- 
tter of the classical bending- torsion type. Ship- 
man and Wood in Ref. 20 have considered this prob- 
lem but they did not take compressibilty and fini- 
te span effects into account. The two-dimensional 
mathematical model used is similar to that employ- 
ed by other authors except that they assumed that 
flutter would first occur when the relative velo- 
city over the rotor blade reaches its critical va- 
lue when 4> = ir/2. For greater or lower values of 
ij), the relative speed would be reduced below the 
critical speed for flutter and any incipient grow- 
ing flutter oscillation would be damped. This rea- 



Since in this problem the motion of the sur- 
face is assumed to be prescribed, the downwash 
velocity at any point on it must be the same as 
the downwash induced by the velocity (or doublet) 
distribution over the surface and its wake. This 
condition must be satisfied in order to ensure 
tangential flow over the surface at all points. 
It is also assumed that the rotor blade is a thin 
surface oscillating about its equilibrium position 

in the plane z = 0. If i; = £'e pt defines the 
downwash displacement at any point (x,y) at time 
t, this boundary condition requires that the down- 
ward velocity and 8<f>/3z must be equal. In the 
transformed coordinates, this implies that 



128 



3Z 



e -i(XX + uT) 



(4) 



where w = -r~ + Dr* Is known. 

at dX 

A further condition that must be imposed on 
any solution is that it leads to zero pressure dif- 
ference across the wake created by the oscillating 
surface. From the general equations of flow it can 
be established that the local lift JE(x,y,t) at any 
point is given by 

j.(x,y,t) = p(||+uf|) (5) 

where k » $^ - $ „ , the discontinuity in the veloci- 

(5) , it immediately follows 



•u 'tf 

ty potential. From Eq. 
that on the surface 



S,(x,y,t) = pU (iyK + — ) e 



(6) 



where v = co/@ and K = 
ivK + f =0 



This yields 



(7) 

everywhere in the wake since the lift must then be 
zero. From Eq. (7), it can be deduced that at any 
point in the wake 



K(X,Y) = K(X t ,Y) e" iv(X-X t ) 



(8) 



where X 

edge of the section at Y 



X denotes the position of the trailing 



As shown in Ref . 19, the solution of Eq. (3) 
may then be derived from the integral equation 



-iKC 



4irW(X ,Y ,0) = // KCX.Yy^F-r- 

p p z^o az K 



-)dXdY 



(9) 



where W is the modified downwash at the point X ,Y 

P P 
given by Eq. (4), K has to take the form specified 
by Eq. (8) at points in the wake and 

5 = [(X-X ) 2 + (Y-Y ) 2 + Z 2 ] 1/2 . 

The double integral in Eq. (9) must be taken 
over the area of the oscillating surface and its 
wake. It should be remembered, however, that K = 
along the pleading edge and the sides of the area of 
integration. 

In the numerical technique developed in Ref. 
18 for calculating the airloads on oscillating 
wings in rectilinear flight, the wing is divided 
into a number of conveniently shaped boxes and K 
is assumed to be constant over each box. The wake, 
on the other hand, is divided into a number of cho- 
rdwise strips and K over each strip is defined by 
Eq. (8). The contribution of the wake to the down- 
wash W(X ,Y ) is then derived by direct numerical 

P P 
integration. 

The application of the method outlined above 
to determine the airloads on rotor blades presents 
certain difficulties, the principal one being that 
the flow velocity over the rotor blades is not con- 
stant as assumed in the derivation of Eqs. (3) and 
(8) for wings in straight flight. To overcome this 
difficulty, it is assumed that the rotor blade can 
be represented by a number of spanwise segments 



over every one of which the flow is taken to have 
its average value and appropriate Mach number. On 
this basis the above analysis can be modified for 
application to rotor blades as outlined in the 
next section. 

Rotor Blade Theory 

In the present analysis, the rotor blade is 
taken to be fixed at the ifi «■ ir/2 position and its 
helical wake is assumed to extend rearwards as 
indicated in Fig. 1. Normally, one would expect 
the vorticity shed by the perturbed blade to be 
carried downstream by the distorted wake of the 
loaded rotor blade. However, in the present preli- 
minary study, uniform inflow is assumed and any 
distortion of the wake due to blade-tip vortex 
interference is ignored. The aerodynamic coeffic- 
ients corresponding to any prescribed motion can . 
then be calculated for forward and hovering flight 
by the method described below. 

a) Forward Flight (Rotor Blade at $ = tt/2) 

Let R denote the tip radius of the blade and 
assume x = Rx' , y = Ry' , and z = Rz 1 . For forward 
flight with velocity V, the relative local velocity 
at section y will be denoted by U(=V+«Ry') and U' 
(=p+y'), where fi is the angular rotation and u(=W 
S2R) is the advance ratio. It then follows that at 
the section y' , the downwash w(x',y') is given by 

(10) 



wCx'.y*) = flR(i|c' + II'IC) e ipt 



P"P 



pax 1 



where 5 = R?'e ™ is the displacement of the blade 
at the point, (x',y'). When the blade is describ- 
ing flapping and twisting motions, 5' may be expr- 
essed as 

C' = Y'f(y') + x'a'F(y') (11) 

where y' and <*' ar e the amplitudes at the reference 
section and f (y') and F(y') are the modes of flap- 
ping and twisting oscillations, respectively. If 
the blade is assumed to be rigid and hinged at the 
root, f(y') = y', and F(y') = 1 in the above equa- 
tion. For convenience, the reference section is 
taken to be at the tip but, in actual flutter cal- 
culations, the section at 0.8R would be a better 
choice. 

To obtain the distribution of K corresponding 
to the motion prescribed by Eq. (10), Eq. (9) is 
first expressed in terms of the original variables 

and K is replaced by Rk'e p . It may then be 
written as 

/ t 1 p X p ..,1 f ~2 -iic'r' 
4irw'e F v _ ,,, F ,, , ... -iX'x'3 ,e 



= //k'Gc'.y^e 
^ P dx'dy ' 



3z 



,2^-r^ 



(12) 

where k' = -&— , X' = Me' , w = w'e lpt , g 2 = 1-M 2 , 
6 fiH' 

and r' = [(x'-x^) 2 + B^y'-y^) 2 + e 2 z' 2 ] 1/2 . 

The above equation can be used to obtain the 
solution to the problem of determining the flow 
over a rotor blade with a rectilinear wake. Since 



129 



the wake can withstand no lift, the condition, 

8k &k 

t— + Ur~ = 0, must be satisfied. For a rectilinear 

dt oX 

wake, this yields 



k'(x',y') 



-i- 



p(x'-x') 



k[(y')e 



S2U' 



(13) 



However, if the wake originating from a blade strip 
is assumed to be curved 

p(s'-s') 

-X ■■ - , 

k'(s',y') -.- k^(y')e **' , (14) 

where q' = (p 2 + y' 2 + 2viy'sin *) 1/2 , s = Rs' is 
the distance along the vortex path and y' specifies 
the spanwise location of the blade strip. 

For computational purposes, Eq. (12) may be 
conveniently expressed as 



4ttW(x' y*) = //K(x',y') f dx'dy', 
P P 



(15) 



where W(x' ,y') = 



w(x',y') -iA'x' 



P P 



P P 



K(x'.y') = k'(x',y') e" U ' X ' 
„2 -iic'r' „ ■ 

' e r-> = -e 2 £ 



G = 



3z 



,2 



(- 



and 
iic'r* 
— t— )[ (1+iic'r*) 



2 2 

(i_2L«l.) + k*Vz' 2 ]. 

r' Z 
It should be noted that in the wake 

v'(s'-s') 

-* r-^- 

K(x',y') - K t (y«) e q 

where v' = -*-=■• 

fig 



(16) 



b) Hovering Flight: 

For the simplest case of hovering flight, 
u = and s' = y'8. Hence Eqs. (10), (11), (12), 
(13) , and (15) can be simply modified by replacing 
•u with zero. Eqs. (14) and (16) then become 



-i|(6-6.) 
k''(6,y') - k^(y') e H •* 

««<i - - -±v'(e-ej 

K(6,y') - K^(y') e fc , 



(17) 



(18) 



where y* defines the location of the blade strip 
from which the wake originates. 

Method Of Solution 

The schematic diagram of the oscillating rotor 
blade is shown in Fig. 1. Eqs. (15) and (16) are 
combined and expressed as 



4irW(x^,yp 



dx'dy' 



-iic'r' 

/ /K(x',y')(3( r 
blade r 

surface 



, -Xl-HLie'r') 

|J s 



s 



-i- 



v'(s'-sp 



/ / K t (y')e 
wake 



3< a 



-IK'r; 



..3 



2 2 
[ (l+i K 'r*)(l- ^_5l_) + K 'Vz' 2 ] ds'dn', (19) 

w 

where r' = [ (x'-x') 2 + B 2 (y'-y') 2 ] 1/2 
s p p 

r; - I (x'-x^) 2 + B 2 (y*-y p ) 2 + B 2 z' 2 ] 1/2 

ds* - (dx ,2 +dy ,2 ) 1/2 
and dn' is perpendicular to ds' and approximately 
equal to dy' on blade. The rotor blade is divided 
into a number of rectangular boxes (M x N) on which 
the doublet strengths are assumed to be constant as 
in Ref . 18. Based on this assumption, Eq. (19) can 
be expressed as 

4«W - I 2 S K . + S T K (20) 
** i=l j=l id ij j=l j tJ 
In Eq. (20) S.. and T. may be interpreted as the 

aerodynamic influence coefficients and the actual 
expressions are given later in this section. S. . 

T. are the downwash velocities induced at the box 

mn due to the unit strength doublets located^ at the 
box ij and the wake strip j respectively. K. . and 

K . are the doublet strengths at the box ij and the 

trailing edge of the wake strip j , respectively. 
With the use of the wake boundary condition, K^ 
can be expressed as 

u! v'(x' -x' ) 
j A „ 1 A Hj. 3 (n) 



tj 



./[e 



+ 2i- 



K tj - V 

Eqs. (20) and (21), then yield 



u: 



AirW mr , -2 S A K 
i=l 1=1 



where A 



ij °ld 



for i ^ M 

v*(x; 



t. rv 



and Aj 



S ij + T j /[e 



D J 



(22) 



v'(x' -<,) 



+ 21- 



t rv 



u j 



ij 

for i - M. _ 

For a given mode shape (W 's are known), Eq. (22) 

represents a system of M x N linear algebraic equa- 
tions, the solution of which yields the values for 

K 's. M and N denote the total number of chord- 
mn 

wise and spanwise stations respectively. 

Once the appropriate K distribution has been 
found, it is then relatively easy to determine the 
aerodynamic forces per unit span acting on the rotor 

blade. If, in Eq. (5), k - Rk'e ipt = RKe ipt e iX ' X ' , 

it then follows that the local lift L(y) - L'(y) 

e p and the nose-up pitching moment, M(y) - M 1 (y) 

e , referred to the mid-point of the chordwise 

section at y are given by 

2z' 

¥£& - (L.+iL.H— £) + (L,+iL,)a' 

p(U R ) 2 (c/2) 12c 34 



130 



■ x t 

<rt>(ir><i£ / k'dx'+U'k') 



(23) 



M'(y) 



2z! 



p (0,^/2) 



= CMj+lM 2 ) (-^-) + (M 3 +±M 4 )a* 



X t X t 

= ( a )( |R ) 2 [u , ( f k , dx , _ k . , _ ± £ / k ' x ' dx '] 

R X I X 2 

Ji * A (24) 

where c is the local chord, D_ is a reference velo- 
city, z' and a' are the local amplitudes of the 
flapping and twisting motions respectively, and L 1 , 
L_, M_, M„, and L„, L,, M-, M,, are the in phase 
and out of phase airload coefficients, respectively. 

Expressions for Aerodynamic Influence Coefficients 

Forward Flight (Rotor Blade at i[> = it/2) 

The influence coefficients are calculated by 
the method outlined in Ref , 18. For a box not con- 
taining the collocation point 



y'+d 2 x'+d -i K 'r 



1^.1 

s 



S,, = - / / (- 5— )e(l+i K 'r')dx'dy' (25) 

d y'f A 2 x i- d i n 

where ic* = -f— , r' - I(x*-x') 2 + g 2 (y'-y') 2 ] 1/2 , 
(S 2 JHJ! s m n 

d = Ax'/2, d„ = Ay '/2, and Ax' and Ay' are the 

chordwise and spanwise spacings of the rectangular 
grid on the surface of the rotor blade. When the 
collocation is inside the box considered, the value 
of-S. must be calculated by the method of Ref. 18. 

For the curved wake 



n^ 2 
T, = - / / e 
' ^ d 2 S t. 



v'(s'-s') . . , , 

. t -iK'r' 

-i 1 _ w 



q B(- r - )[(l+iK'r;) 

r' 

w 



(26) 



2 2 
(1- 3g *' ■ ) + k ,2 B 2 z' 2 ] ds'dn* 

r ,Z 

w 

where r' = [(x*-x') 2 + f3 2 (y'-y') 2 + 6 2 z' 2 ] 1/2 , 
w m n 

x' = u9 + y' sin 9, y' = y! cos 9, z' = d'9/2-ir, 

ds' = (dx' 2 +dy' 2 ) 1/2 = (u 2 +y' 2 +2uy' cos 9) 1/2 d9, 

and d(=Rd'), the downward displacement of the wake 
per revolution, is assumed to be small. T. 's are 

evaluated numerically at the j ' th spanwise strip by 
taking small increments of 9 and n. 

Hovering Flight (Low Inflows) 

For hovering flight, u = 0, s' = y'9, and 

ds' - y!d9. The expression for S .. , Eq. (25), can 

be simply modified by replacing u with zero and the 
wake integration for the j ' th strip 



n j«2- -iV(e-v - lK ' r w 

/ / e C 6 (- r - )[(l+i K 'r') 

n'-d, .9,. r' 3 w 

j 2 t w 



2 2 
(1- 3g *' ) + k'Vz' 2 ] y! d9dn' 

r' •• 

w 



(27) 



where x' = y' sin 9, y' = y' cos 9, and z 1 - d'9/2ir. 

The effect of the helical wake in hovering 
flight is estimated by two different methods. In 
the first method, a Helical Wake Model is used and 
the actual helical path is taken in evaluating the 
T. coefficients. In the second method, a Circular 

Wake Model is employed and the helical wake is re- 
placed by its near wake, which is assumed to extend 
over 9 £ 8 £ ir/2, and a number of regularly spaced 

circular disks of vorticity below the reference 
plane. The formula for k' for the n'th disk at 
z' = nd' is taken to be simply 

-i£[(6-9.)+2nir] 
k'(6,y',nd') = k£<y')e B ,. (28> 

the actual spacing between consecutive disks being 
Rd\ 

Results and Discussion 

A rectangular rotor blade of R/c = 10 was cho- 
sen and the. blade Was assumed to extend from 0.1R 
to R. For the computation of the airload coefficie- 
nts, a grid of thirty six rectangular boxes consist- 
ing of six chordwise and six spanwise stations were 
used. The convergence of the results was tested by 
taking grid sizes of 6x8 (chordwise x spanwise) and 
8x6. Rigid mode shapes for flapping and twisting 
oscillations are assumed so that 

? = yy' + ax' 
w'(x'.y') = £5R[a*(u+y,I+i£0 + Y'if(v+y^)] 



P"P 



p S p 



S v 



p 

2z! 



^p^x-ir^ 



(29) 



It should be noted that the above equation is valid 
only for the blade position at $ = it/2. For hover- 
ing flight, u = in Eq. (29). 

The airload coefficients for Mach numbers 
0.8 for several values of *- and wake spacing of two 

chords were obtained with reference to the blade's 
quater-chord axis. In Figs. (2) thru (5), selected 
airload coefficients for slow forward flight (u = 
0.1) are compared with the results obtained for 
hovering flight using both a helical wake model and 
two-dimensional strip theory. For this particular 
comparison, the reference velocity in Eqs. (23) and 
(24) was taken as the relative local velocity, U, 
and the tip Mach number was 0.8. From these plots, 
one can conclude that strip theory predicts substa- 
ntially larger values for the airload coefficients. 
One of the most important observations one can make 
is that the curved wake changes the chordwise load 
distribution in such a way that the center of pre- 
ssure shifts forward of the quarter-chord axis 
position (see Fig. 4). 



131 



Figs. (6) and (7) compare the results by seve- 
ral mathematical models used for the hovering fli- 
ght case. The airload coefficients are referred 
to the tip velocity (fiR) and this choice was made 
to indicate the trends of spanwise load distribut- 
ion. The Circular Wake model representation resu- 
lts in a substantial saving in computational time. 
For example, to obtain the airload coefficients for 
one set of geometric and flight conditions using 
6x6 grid on the blade, the Circular Wake model took 
only 1.5 minutes of computing time on IBM 360/65 
while the Helical Wake model took 2.5 minutes. 
Although the Circular Wake model seems to indicate 
the general trends of the airload coefficients, one 
should use the full helical wake to compute the 
airload coefficients accurately. 

Some typical results for hovering flight using 
the Circular Wake representation, compared with the 
results of two-dimensional strip theory, Ref. 7, 
are shown in Figs. (8) thru (11). The results for 
the curved wake are in good agreement with the re- 
sults for strip theory for the inner blade sections; 
however, the agreement is poor towards the tip. 
Figs. (12) and (13) show the variation with axis 
position of M, , conveniently referred to as pitching 
moment damping airload coefficient, at spanwise 
stations of 0.475R and 0.925R, respectively. From 
these results, one can conclude that the agreement 
between the curved wake results and strip theory 
is good near the quarter-chord position for M = 
and M = 0.8 but it becomes very poor as the axis 
is moved towards the trailing edge. 

References 



1. 



6. 
7. 

8. 
9. 



Jones, W. P., McCrosky, W. J., and Costes, J. 
J. , "Unsteady Aerodynamics of Helicopter Rotor 
Blades," NATO AGARD Report No. 595, April 1972. 
Jones , J . P . , "The Influence of the Wake on 
the Flutter and Vibration of Rotor Blades," 
Aeronautical Quarterly , Vol. IX, August 1958. 
Daughaday, H. and Kline, J., "An Investigation 
of the Effect of Virtual Delta-Three Angle and 
Blade Flexibility on Rotor Blade Flutter," 
Cornell Aeronautical Laboratory Report, SB-86 
2-5-2, August 1954. 

Loewy, R. G., "A Two-Dimensional Approximation 
to the Unsteady Aerodynamics of Rotary Wings," 
Journal of the Aeronautical Sciences , Vol. 24, 
No. 2, February 1957, pp. 81-92. 
Timman, R. and Van de Vooren, A. I. , "Flutter 
of a Helicopter Rotor Ratating in its Own 
Wake , " Journal of the Aeronautical Sciences , 
Vol. 24, No. 9, September 1957, pp. 694-702. 
Jones, W. P., "Aerodynamic Forces on Wings in 
Non-Uniform Motion," R&M No. 2117, 1945, Bri- 
tish Aeronautical Research Council. 
Jones, W. P. and Rao, B. M. , "Compressibility 
Effects on Oscillating Rotor BladeB in Hover- 
ing Flight," AIAA Journal , Vol. 8, No. 2, 
February 1970, pp. 321-329. 
Jones, W. P., "The Oscillating Airfoil in 
Subsonic Flow," R&M No. 2921, 1956, British 
Aeronautical Research Council. 
Hammond, C. E., "Compressibility Effects in 
Helicopter Rotor Blade Flutter," GITAER Report 
69-4, December 1969, Georgia Institute of 



Technology, School of Aerospace Engineering. 

10. Ashley, H. , Moser, H. H. , and Dugundji, J., 
"Investigation of Rotor Response to Vibratory 
Aerodynamic Inputs, Part III, Three-Dimensional 
Effects on Unsteady Flow Through a Helicopter 
Rotor," WADC TR 58-87, October 1958, AD203392, 
U. S. Air Force Air Research and Development 
Command. 

11. Reissner, E. , "Effects of Finite Span on the 
Airload Distributions for Oscillating Wings, 
Part I - Aerodynamic Theory of Oscillating 
Wings of Finite Span," NACA Technical Note 
No. 1194, 1947. 

12. Jones, W. P. and Rao, B. M. , "Tip Vortex Effe- 
cts on Oscillating Rotor Blades in Hovering 
Flight," AIAA Journal , Vol. 9, No. 1, January 
1971, pp. 106-113. 

13. Miller, R. H. , "On the Computation of Airloads 
Acting on Rotor Blades in Forward Flight," 
Journal of the American Helicopter Society , 
Vol. 7, No. 2, April 1962, pp. 55-66. 

14. Miller, R. H. , "Unsteady Airloads on Helicop- 
ter Rotor Blades," Journal of the Royal Aero- 
nautical Society , Vol. 86, No. 640, April 1964, 
pp. 217-229. 

15. Miller, R. H. , "Rotor Blade Harmonic Air 
Loading," AIAA Journal , Vol. 2, No. 7, July 
1964, pp. 1254-1269. 

16. Piziali, R. A., "A Method for Predicting the 
Aerodynamic Loads and Dynamic Response of 
Rotor Blades," USAAV-LABS Technical Report 
65-74, January 1966, AD 628583. 

17. Sadler, S. G. , "A Method for Predicting Heli- 
copter Wake Geometry, Wake Induced Flow and 
Wake Effects on Blade Airloads," presented at 
the 27th Annual National V/ST0L Forum of the 
American Helicopter Society, Washington, 

D. C, May 1972. 

18. Jones, W. P. and Moore, J. A., "Simplified 
Aerodynamic Theory of Oscillating Thin Surfa- 
ces in Subsonic Flow," AIAA Journal , Vol. 11, 
No. 9, September 1973, pp. 1305-1309. 

19. Jones, W. P., "Oscillating Wings in Compressi- 
ble Subsonic Flow," R&M No. 2885, October 19 
55, British Aeronautical Research Council. 

20. Shipman, K. W. and Wood, E. R. , "A Two-Dimen- 
sional Theory for Rotor Blade in Forward Fli- 
ght," Journal of Aircraft , Vol. 8, No. 12, 
December 1971, pp. 1008-1015. 



132 











1 












K&- 


"A 




tai.y'i) — | 






\ 




\ 














i 












\J» 








ii 


















, / 


V 






III 












/ jth wake/ 
\/ strip / 












S 1 













*■ y ,n,l 



♦ 

x',m,i 



Pig. 1 Schematic Diagram of Rotor Blade and 
its Wake. 



4.5 



4.0 



3.5 



3.0 



2.5 



2.0 



FORWARD FLIGHT (p=0.l) 

HOVERING FLIGHT 

STRIP THEORY _ 



0.10 




M'OB.jp I, Ur'U 



FORWARD FLIGHT (^-0.1) 

HOVERING FLIGHT 

STRIP THEORY 




Fig. 3 Spanwise Variation of L » 



2.0 
1.5 


-^ y^-' \ 

/ M = 0.8, -jj =1 , U R =U ' ( 


1.0 

0.5 




/ FORWARD FLIGHT (^ = 0.1) | 

/ -— — HOVERING FLIGHT 1 
STRIP THEORY j 

— 

-~UaJ_. 1 1 1 



0.10 



1.0 



0.8 



0.25 0.40 0.55 0.70 0.85 1.00 

R 
Fig. 4 Spanwise Variation of M,. 



0.6 



-M 4 



Fig. 2 Spanwise Variation of Lj. 



0.4 



0.2 



\ 




\\ M-0.8, "f.UFcU 




-\ FORWARD FLIGHT (^ 

\\ HOVERING FLIGHT 


■0.1) 


\\ STRIP THEORY 

\ ^ 






\j*«O.I J 


:\ 


1 


1 1 



133 



0.1 0.25 0.40 0.55 0.70 0.85 I.C 

1 
R 

Fig. 5 Spanwise Variation of M. . 



0.25 



0.20 





M = 0.8, ■£ = !, U„=flR 




HELICAL WAKE 




CIRCULAR WAKE y 

STRIP THEORY y 










S 






s 








\ 
I 
1 




^^ 


1 

1 
t 
1 
t 

1 



0.1 0.25 0.40 0.55 0.70 0.85 1.00 

x 

R 

Fig. 6 Spanwise Variation of L. for 
Hovering Flight. 



0.7 



U H =flR 



-CIRCULAR WAKE MODEL / 
-STRIP THEORY / , 

/ / 




Fig. 8 Spanwise Variation of L 2 for 
Hovering Flight. 



0.25 


M=0£, jy»l, U R = flR 


1 








HELICAL WAKE / 








CIRCULAR WAKE • 

STRIP THEORY / 






0.20 








/ 








/ 






M. 0.15 
4 


— / S" 


— 


\ 




/ ,' 




\ 
\ 




/ / 




0.10 


_ S /V~ 






■' >' / 












0.05 


S — -'"L-/ 












o 





0.1 0.25 0.40 0.55 0.70 0.85 1.0 



y 

"R~ 

Fig. 7 Spanwise Variation of M, for 
Hovering Flight. 



3.5 



3.0 



CIRCULAR WAKE MODEL / 

STRIP THEORY / 




01 025 0.40 0.55 0.70 0.85 1.0 



y 

"R" 



Fig. 9 Spanwise Variation of L, for 
Hovering Flight. 



134 



I 20 



0.8 



Ur'OR 

— CIRCULAR WAKE MODEL 
STRIP THEORY 




Fig. 10 Spanwise Variation of M 3 for 
Hovering Flight. 



|-=0.475, U„=GR 



CIRCULAR WAKE MODEL 

-— STRIP THEORY 



-M„ 




Fig. 12 Variation of M^ With Reference Axis 
Position for Hovering Flight. 



-M, 




Fig. 11 Spanwise Variation of M, for 
Hovering Flight. 



2.5 



2.0 



-M„ 



-i =0.925, U R =»R 



CIRCULAR WAKE MODEL 

STRIP THEORY / 

/ 
/ 
/ 
/ 
/ 
/ 




Fig. 13 Variation of M. With Reference Axis 
Position for Hovering Flight. 



135 



ROTOR AEROELASTIC STABILITY 
COUPLED WITH HELICOPTER BODY MOTION 

Wen-Liu Miao 

Boeing Vertol Company 

Philadelphia, Pennsylvania 

Helmut B. Huber 

Messersehmitt-Boelkow-Blohm Gmbh 

Ottobrun-Munieh 

Federal Republic of Germany 



Abstract 



ro 



A 5. 5-foot-diameter, soft- in-plane, hingeless- 



rdt-or system was tested on a gimbal which allowed the 
helicopter rigid-body pitch and roll motions. With this 
model, coupled rotor /airframe aeroelastic stability 
boundaries were explored and the modal damping ratios 
were measured. The time histories were correlated 
with analysis with excellent agreement. 

The effects of forward speed and some rotor de- 
sign parameters on the coupled rotor/airframe stability 
were explored both by model and analysis. Some phys- 
ical insights into the coupled stability phenomenon were 
suggested. 

Introduction 

The coupled rotor-airframe aeroelastic stability 
phenomenon of air resonance has received considerable 
attention in recent years. A scaled model of the BO-105 
helicopter was built and tested to explore this phenom- 
enon and its sensitivity to design parameters. 1 An ex- 
tensive analytical study was performed and correlated 
with BO-105 flight test data. 2 

To further explore this coupled stability phenom- 
enon, a large scale model having different resonance 
characteristics than the BO-105 was built and tested. 
Parameters that were influential to the stability 1 ' 2 
were incorporated into the model and their effects were 
examined. An improved test technique enabled the de- 
termination of modal damping ratio at every test point, 
providing better data for correlation and better assess- 
ment of stability. 

Description of Model 

The model, shown in Figure 1, consisted of a 
Froude-scaled model rotor mounted on a rigid fuselage, 
■which in turn was mounted on a two-axis gimbal having 
* 10 degrees travel in pitch and roll. The model had a 
5. 5-foot-diameter, soft-in-plane, hingeless rotor with 
pertinent hub parameters such as precone, sweep, and 
control system stiffness being variables to enable in- 
vestigation of their effects on coupled rotor-airframe 
stability. 

A proportional (closed-loop) control system 
equipped with a cyclic stick provided lateral and longi- 
tudinal control to fly the model in the pitch and roll de- 
grees of freedom. In addition, a shaker system was 
installed in the longitudinal and lateral cyclic system 

Presented at the AHS/NASA-Ames Specialists Meeting 
on Rotorcraft Dynamics, February 13-15, 1974. 




Figure 1. Dynamic Model Helicopter With 5.5-Foot- 
Diameter Single Rotor 



(see Figure 2) to allow excitation of the model at the 
desired frequency. This enabled the measurement of 
the modal damping ratios at each test point. The meas- 
ured modal damping permitted the precise determina- 
tion of the stability boundaries and also showed the 
sxtent of stability when the model was stable. 




Figure 2. Details of Model Rotor Hub and 
Swashplate 



137 



The stability and control augmentation system was 
based on position feedback. Position potentiometers on : 
the helicopter gimbal axes provided position feedback 
signals which were amplified, filtered, and fed into the 
cyclic actuators for automatic stabilization of the model. 
The filter was designed to block any feedback at a fre- 
quency of ft-u)£ and thus eliminated any control inputs 
that would tend to interact with the air-resonance mode. 

Collective pitch was set by means of an open-loop 
control and a pitch-angle indicator. Other controls pro- 
vided for the operator included mounting-pylon pitch 
attitude, stick trim, and quick-acting and slow-acting, 
self-centering snubbers to lock out the pitch and roll 
degrees of freedom. The horizontal stabilizer was 
manually trimmable and rotor speed was controlled by 
the wind tunnel operator. 

Signals from the blade flap, torsion, and chord 
strain gages, along with body pitch and roll motion, 
cyclic stick position, and l/rev, were recorded on os- 
cillograph as well as on multiplex tape recorder. One 
of the chord-bending traces was filtered to display the 
chord bending at the critical lag natural frequency to 
allow quick determination of modal damping on line. 
Most of the testing was performed in the wind tunnel at 
the University of Maryland. 

Test Technique 

As discussed in References 1 and 2, the air- 
resonance mode stability is determined by the blade 
collective pitch as well as the rotor speed. Therefore, 
for every airspeed, a comprehensive variation of rpm 
and collective pitch was conducted. 



SET UP 
TEST CONDITION 



RECORD 
REASONABLE 
DIVERGENCE 
OF MODEL 
MOTION 




SET UP NEW 
CONDITION 



RECORD 

CONVERGENCE 
OF MODEL 
MOTIONS 



STABILIZE 

MODEL BV: DROP RPM, 

DROP COLLECTIVE, OR SNUBBING 



ANALYZE DATA 
OBTAIN MODAL 
DAMPING RATIO 





"°* 



V 



COLLECTIVE » 
PITCH I 




Figure 3 shows the test flow of events for each 
data point taken. After the test conditions had been set 
up (rpm, tunnel speed, and collective pitch), the model 
was trimmed and was held at the trim attitude with the 
stability and control augmentation system (SCAS). The 
shaker and the tracking filter frequ< ■ were set to 
0-w e and oj ? respectively, with the absolute magni 
tudes dependent on the rotor speed. Both the multiplex 
tape recorder and the CEC recorders were turned on to/ 
record the steady-state response of the model. The 
swashplate was then oscillated in the lateral control 
direction. After the termination of the excitation, re- 
cording was continued until steady- state conditions weje 
again reached, when practical. The decay of the filte/ed, 
in-plane, bending-moment trace was reduced to obtai$ 
the modal damping ratio. 

SYMBOL CONDITION 



O 

d 



STABLE 

MARGINAL 

UNSTABLE 



Eh 
W 



200 








o 






o 


o 












o 


o 


o 


o 


o 


o 




160 


- 


o 


o 


o 
o 


o 
o 


o 
o 


o 
o 


o 
o 


o 
o 






o 







o 


o 


o 


o 


ooo 






o 


o 


\0 


o 


o 





o 


ooo 


120 


§! 


»o 

lo 


o 

o 


o 


o 


o 
o 


o 
o 




o 


odd 
od 




ooo 


o 


o 


o 




o 


o 


o» 




ooo 


o 


o 


o 


o 




o 


ood 


80 


~o 




o 


o 


o 


o. 


o^ 




ooo 




o 




o 


o 


o 


o 


o 


o 


"ooe ig 




o 




o 


o 


o 


o 


o 


o 


ooo 


40 


o 
o 




o 
o 


o 




o 
o 


o 
o 


o 
o 


o 
o 


ooo 









o 


o 


o 


o 


o 










o 







o 







o 


o 








- o 







o 




o 


o 


o 


o 




o 




o 







o 


o 


o 


o 




o 















o 


o 


o 


-40 


o 

~ 




o 




o 
o 






o 




o 




o 


o 
o 









o 


o 




o 





o 


o 











1 


o 




o 
_1 


o 


o 

,„l 


o 

1 



60 80 
ROTOR SPEED 



100 



120 



140 



NORMAL ROTOR SPEED 



PERCENT 



Figure 3. Plow Diagram of Test Technique 



Figure 4. Typical Map of Test Points in Hover 



Test Results 

Figure 4 shows a typical map of test points taken 
at a constant tunnel speed, in this case in hover. Two 



138 



stability boundaries were present: one at about 70 per- 
cent of normal rotor speed and 120 percent of normal 
collective pitch and another at about 135 percent rpm 
and 100 percent collective. Examination of the coupled 
frequency variation with rotor speed while holding con- 
stant thrust, Figure 5, reveals that the low-rpm bound- 
ary corresponds to the resonance with the body-pitch- 
predominant mode and the high-rpm boundary with the 
body-roll-predominant mode. 



1.0 



















jgg 


gs^ 














0M1NANT , 










R 
M 


M>E _, 





*G 










— \/PI 
' — <fM0 


TCH-PREDOMINANT 
DE 

1 









80 90 
ROTOR SPEED 



NORMAL ROTOR SPEED 



100 110 
- PERCENT 



Figure 5. Coupled Resonance Characteristics 



Figures 6, 7, and 8 show the time histories of 
three hover air-resonance points which are at constant 
collective pitch of 133 percent 6NOR 0-& hover collec- 
tive at Njjor) w i*h rotor speeds of 100 percent, 72 
percent, and 67 percent NNOR respectively. At NNOR 
the chord bending decayed after the excitation termi- 
nated, at a rate of approximately 1 percent of critical 
damping, and the body participation was barely detect- 
able. Approaching the stability boundary at 72 percent 
rpm, the chord bending took longer to decay compared 
to the 100 percent rpm case. Body participation was 
quite pronounced in both pitch and roll. While the fil- 
tered chord-bending gage in the rotating system was 
indicating at the blade lag natural frequency, ca^ , the 
body pitch and roll motions responding in the same air- 
resonance mode were at the fixed-system frequency of 
8-id. . It is of interest that these Q-us^ body motions 
are superimposed on some very low-frequency, flying- 
quality-type motions. 



At 67 percent rpm, Figure 8, the air-resonance 
mode started to diverge after being excited; when the 
body was snubbed, the blade motion decayed and re- 
turned to the l/rev forced response. 



The response characteristics described here held 
true for all airspeeds tested up to a scaled test speed 
limit of 225 knots. 



HOVER, 100% N Nnn , 133% 6 Mr)p , RON NO. 8 



"NOR 



NOR' 



BLADE CHORD 



/V\/1%\/\a/W 



\W\AWVWvW\aa 



BLADE CHORD FILTERED 



-MA/VW\/W\/WWWAV.v u ^,^^ 



"\y\y\jr^^ 



-H u c|«- 



BODY PITCH 




BODY ROLL 




LATERAL EXCITATION 



1/REV 

I I I I I I I I I 1 I I I I I I I I > I M I I I I I 1 I I I I I I II 1 I I i I 1 I I I I 1 I I I I I I 1 I I 1 I 1 



Figure 6. Response Time Histories in Hover at %qr 

139 



BLADE CHORD 



HOVER, 72% N N0 R, 133% e NOR, - RUN N0 - 13 




1/REV 
IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIMIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIMIIIIIIIIIIIIIIIMIII 

Figure 7. Response Time Histories in Hover at 72 
Percent N N0R 



Analytical Model 

To treat the dynamically and aerodynamically 
coupled rotor-airframe air-resonance problem, the 
analytical model shown in Figure 9 is used. In this 
model, the elastic cantilevered blade is represented by 
a spring-restrained, hinged rigid blade. Three hinges 
are used to simulate the first flap, first lag, and first 
torsion modes, in that order from inboard to outboard. 
In addition, a pitch degree of freedom is provided in- 
board of the flap hinge to facilitate the simulation of any 
torsional stiffness distribution relative to the flap and 
lag hinges. The blade model includes built-in pitch axis 
precone, blade sweep, kinematic pitch-flap and pitch- 
lag coupling, and a variable chordwise center-of-gravity 
distribution over the blade span. 



The airframe has five rigid-body freedoms: 
longitudinal, lateral, vertical, pitch, and roll; and two 
flexible freedoms: pylon pitch and pylon roll. The 
equations of motion are nonlinear and are solved by a 
numerical time-history solution technique. The blade 
: degrees of freedom are calculated for each individual 
blade. 



To evaluate the aeroelastic stability, the aircraft 
can be perturbed from the trimmed state. For air- 
resonance investigations this is usually done by oscil- 
latory stick excitations, which Can be simulated in any 
frequency. The time history of each degree of freedom 
is then subjected to an oscillation analysis program to 
obtain the frequencies, amplitudes, phases, and 
damping coefficients. A more detailed discussion of 
this analytical procedure can be found in Reference 2. 



The aerodynamic model is based on current blade- 
element theory and can handle all hover, fo I .light, 
and maneuver flight conditions. It uses two-dimensional 
airfoil data with stall, reverse-flow, and compressibil- 
ity effects. 



Using a linear lift-curve slope, this coupled 
analysis in hover can be reduced to a set of second- 
order differential equations with constant coefficients 
by applying the quasi-normal coordinate transformation 



140 



HOVER, 67% N N0R# 133% 8j, 0Rf RUN NO. 12 



BLADE CHORD 




1/REV 

iiiiuiiiiiiiii iii iii i i ii i iiii i i iii i i i iiiiii ii i ii ii iiii i i iii iiii i i i ii ii i ii ii i iii i ii ii i i i iiiii i ii i iii i i i i i iiii i iiiii i i i iii i iiiniiiii ii iiiuiiiii 

Figure 8. Response Time Histories in Hover at 67 
Percent Njjqr 



for the rotating eoordinates3. This enables the closed- 
form solution. The eigenvalues and eigenvectors thus 
obtained yield the information on frequencies, damping, 
and mode shapes. 



PITCH FLAP LAG TORSION 

«. \ \ c, ^ 




'LONGITUDINAL VERTICAL 



PITCH 



LATERAL 



Figure 9. Coupled Rotor-Fuselage Analytical Model 



Correlation 

Rotor Thrust 

Figure 10 shows the air-resonance mode modal 
damping ratio variations with thrust at NnoR in hover. 
The agreement between test and analysis is quite good. 
The propitious trend with increasing collective pitch is 
due partly to the increase of aerodynamic damping, but 
is mainly a result of the favorable piteh-flap-lag 
coupling. A typical blade elastic coupling is shown in 
Figure 18 where the blade flap, lag, and pitch torsion 
responses to a cyclic-pitch input are indicated. The 
type of elastic coupling of this rotor system is discussed 
in more detail in a subsequent section. 



Rotor Speed 

Shown in Figure 11 are the test correlations of 
the air-resonance mode damping variation with rotor 
speed at constant collective pitch (133 percent SjJOR) 
in hover. The analytical results are in good agreement 
with test points over the whole rotor speed range. The 
stability boundary corresponding to the resonance with 
the body-pitch-predominant mode at low rpm is pre- 
dicted well by theory. The somewhat higher level of 
damping of the test points might indicate that the struc- 
tural damping of the real model blade is higher than the 
0. 5 percent damping assumed in the analysis. 



141 



6.0 




HOVER 

COLLECTIVE PITCH = 133% 9 



ROTOR SPEED 



= 72% N, 



EST] 



NOR 
[ NOR 



LATERAL EXCITATION 



BLADE CHORDWISE MOMENT 



50 100 

COLLECTIVE PITCH 
HOVER lg COLLECTIVE PITCH 



150 200 
~ PERCENT 




Figure 10. Effect of Thrust on Air-Resonance 
Stability 



HOVER 

COLLECTIVE PITCH = 133% 6 N0R 
IN-PLANE DAMPING . = 0.5% CRITICAL 
CONTROL STIFFNESS = 642 IN. -LB/RAD 



O H 
SSfrl 
H H 

|U 

Q H 

a « 

S Ai 



-1 









A 


A 

i 


i 


A 












3< 

a 
< 

s 


\ 






Ul 

-1 
m 

l{ 

3 


7 


A 






A TEST 
— ANALS 


POINTS 
SIS 



50 60 70 80 

ROTOR SPEED 
NORMAL ROTOR SPEED 



90 



PERCENT 



100 



110 



Figure 11. Effect of Rotor Speed on Air-Resonance 
Stability : 



| A N A L Y S I S| 

LATERAL EXCITATION 



[ -W\A 



BLADE IN-PLANE MOTION 




Hhr. 



BODY ROLL 




AAA/W 



k- 



n-a) r 



•TIME 



ONE-PER-REV MARK 

Figure 12. Correlation of Test and Analysis of 
Time Histories in Hover 



The good agreement of Figure 11 is merely a 
reflection of the excellent correlation between test and 
analysis in the time-history waveform of blade and body 
motions. One example is shown in Figure 12. For this 
case the oscillation analysis program yields a damping 
coefficient of 0. 39 percent at blade lag natural frequency 
for the rotating blade. 



Forward Flight 

The test trend of air-resonance mode damping 
with airspeed is also verified by analysis in Figure 13. 
Test points shown in this diagram were obtained with 
constant collective pitch, so that they do not correspond 
to a lg-thrustAevel-flight condition. The analysis was 



142 



performed under the same colleetive/shaft-angle set- 
tings to get an exact simulation of the test conditions. 
At 150 knots, the collective pitch is slightly reduced, 
from 133 percent to 111 percent, which produces a 
sharp decrease in rotor thrust. Therefore the air- 
resonance mode is less stable than for a normal lg- 
thrust condition. 

COLLECTIVE PITCH = 133% 6 N0R (111% 9 N0R ) 
ROTOR SPEED = 100% % 0R 
IN-PLANE DAMPING =0.5 PERCENT CRITICAL 
CONTROL STIFFNESS = 642 IN. -LB/RAD 
31 1 1 1 <—r 



T E S T 



80 KNOTS FORWARD FLIGHT 

COLLECTIVE PITCH = 133% 6 N0 R 
ROTOR SPEED =100% NjjOR 
SHAFT TILT ANGLE = -4 DEGREES 




[ 



LATERAL EXCITATION 

^AAAAAA^ — 



BLADE CHORDWISE MOMENT 



NOR 




BODY ROLL 



80 120 160 

VELOCITY - KNOTS 

Figure 13. Effect of Forward Speed on Air- 
Resonance Stability at Constant 
Collective Pitch 



Theory' shows some influence of cyclic control on 
air-resonance stability at high speed. As longitudinal 
cyclic also controls rotor thrust in forward flight, this 
variation of air-resonance stability comes solely from 
the change in rotor thrust. With thrust held constant the 
stability is insensitive to steady 1/rev cyclic-pitch vari- 
ation. This is shown in a later section. 

In Figure 14 one example of a typical time history 
at a scaled airspeed of 80 knots is compared between 
test and analysis. When one considers the complex fre- 
quency modulations during this excited air-resonance 
case, the correlation can be said to be excellent. This 
should indicate that theory allows a definitive and reli- 
able view of a helicopter's stability characteristics. 

Additional test results of air-resonance stability 
in forward flight are illustrated in Figure 15. This 
trend, which was obtained for a lg/level-flight condition, 
follows the rotor power curve quite well. As shown in 
Figure 10, for a moderate range of thrust variation, say 
around lg, the air-resonance mode becomes more 
stable with increasing thrust and less stable with de- 
creasing thrust. The forward-speed trend here simply 
reflects this thrust (and aerodynamic coning angle) de- 
pendency. This trend, which shows that the air- 
resonance mode stability improves significantly at high 
forward speeds, is also apparent in the BO-105 flight 
test data2. 




ANALYSIS 



LATERAL EXCITATION 

FiAAAA/W^- 



BLADE IN-PLANE MOTION 




ONE -PER-REV MARK 

'iiifiiiiiiiiiiiiiiiiiiiiiiiii 



TIME 



llllllllllllllll 



Figure 14. Correlation of Test and Analysis of 
Time Histories in Forward Flight 



143 




80 120 160 

VELOCITY - KNOTS 



200 



Figure 15. Effect of Forward Speed on Air- 
Resonance Stability in lg Level 
Flight 



Physics of Air Resonance 



General 



The mechanism and the stability characteristics 
of air resonance have been well described in numerous 
papers. !> 2 > 4 > 5, 6 n suffices to say here that the soft- 
in-plane hingeless -rotor system derives its inherent 
stability mainly from the powerful flap damping. While 
rotors with untwisted blades may have substantial reduc- 
tion in the flap damping near zero thrust, the damping 
available remains essentially unchanged for blades with 
nominal twist. Figure 16 shows the test data for various 
blades of different twist. Above a thrust coefficient of 
0. 005, the twisted blade and the untwisted blade both 
have the same thrust-per-collective slope. While the 
untwisted blade has a drastic reduction in slope with re- 
duction in thrust in both theory and test, that of the 
twisted blade remains the same. 



SYMBOL 


TEST 


AIRFOIL 


°t 


A 


BTS 6-FT ROTOR 


V23010 


-7° 


D 


RTS 6-FT ROTOR 


VR7 


-9° 


tf 


RTS 6-FT ROTOR 


VR7/8 


-9° 


14-FT ROTOR 


V23010/13006 


-10.5" 


a 


UHM COMPOSITE 


V23O10 


-10.5° 


o 


UHM 6-FT ROTOR 


VR5 


-14° 


• 


MBB TIEDOWN 


0012 


0° 


▲ 


AMRDL MODEL 


23012 


0° 



g 0.03 



° 0.02 



« 0.01 





<J =0 


061 






E/1 


ft 












ill 
I ff 








TEST DATA 
AT () t = -9° 


>. cjj 


7j» 

7^~~TEST DATA 
AT t = 0° 










V 












j/% 


F THEORY AT 


t =0° 




TEST DA 
AT 0t = 


rA 


2Ba? 




































// 
// 
»/ 
/ 















-4-2 2 4 6 8 10 
COLLECTIVE PITCH, 6.75 - DEGREES 

Figure 16. Effect of Blade Twist on Thrust 
Coefficient 



Let us examine the coupling terms that are inher- 
ent in the hingeless rotor system with an equivalent 
hinge sequence of pitch-flap-lag from inboard to out- 
board. One term that stands out is the perturbation 
pitch moment produced by the induced drag (steady 
force) acting through a moment arm of vertical-flapping 
displacement (perturbation deflection) . This flap-pitch 
coupling term due to the induced drag has the Sense of 
flap up/pitch noseup. Figure 17 compares the air- 
resonance mode damping of the same rotor system with 
this particular coupling term suppressed. With the 
induced-drag term suppressed, the air-resonance mode 
does not become unstable at high collective where the 
induced drag dominates. 



By the same consideration, the air-resonance 
mode should become more stable in descent since during 
descent, the induced drag acts toward the leading edge 
producing a flap-pitch coupling of flap-up/pitch-nose- 
down sense, which is stabilizing. 



HCTIOR SPEED - lOOt 1% 0E 



r\ 




HftNEUVER q LEVEL 



Figure 17. Stability Characteristics With Sup- 
pression of Flap-Pitch Coupling Term 
Due to Induced Drag 



144 



Pitch- Flap- Lag Coupling Characteristics 

For a complete understanding of the elastic- 
coupling characteristics of a hingeless rotor with a 
pitch-flap-lag sequence of hinges , all blade motions 
must be considered together. For this purpose it is in- 
structive to analyze a simple cyclic-pitch case in hover. 
In Figure 18 the elastic flap, lead-lag, and pitch mo- 
tions are shown over one rotor revolution. It can be 
seen clearly that the flap and lag motions are accom- 
panied by an elastic pitch torsion, the resultant coupling 
being in the sense of flap up/lead forward/pitch nose- 
down. For a clear understanding this complex coupling 
can be divided into two distinct coupling phenomena: the 
one equivalent to a negative pitch-flap coupling (flap up/ 
pitch nosedown) , the other equivalent to a positive pitch- 
lag coupling (lead forward/pitch nosedown). The cou- 
pling factors are 0. 4 degree pitch per degree flap and 
0. 6 degree pitch per degree lag. 



flexibility. Some of these design rules have already been 
applied to this model rotor design (low precone, aft 
sweep, soft control systems). 

Parametric Sensitivities 

The following paragraphs describe the air- 
resonance mode stability sensitivities obtained from the 
model test. 

Climb and Descent 

Figure 19 shows the sensitivity with lg climb and 
descent at a scaled airspeed of 80 knots. With normal 
control system stiffness (90 in. -lb/rad) , descent sta- 
bilizes the mode as discussed in the previous section; 
conversely, climb has a destabilizing effect. 






O 
H 
Q 



3 
O 
H 

n 
o 
w 

Hi 

h 

pa 
a 

w 



PQ 



PRECONE * DEGREES 

SWEEP =2.5 DEGREES AFT 

CONTROL STIFFNESS =90 IN. -LB/RAD 




360 



AZIMUTH ANGLE 



DEGREES 



!** 2 



J 1 

V 


1 — 

= 80 KNOTS 


. , 

SYMBOL 


CONTROL STIFFNESS 


a 


= %OR 


D 


90 IN. -LB/RAD 




O 


642 IN. -LB/RAD 
















_VV_ 


















H 






















i-l 






















a 






















f 






















Ul 





-20 -10 * 10 

t CLIMB | DESCENT | 

ROTOR ANGLE OF ATTACK - DEGREES 



30 



Figure 19. 



Effects of Climb, Descent, and Control 
System Stiffness on Air-Resonance 
Stability 



Figure 18. Blade Elastic Coupling 



Control System Stiffness 

Also shown in Figure 19 are the test data obtained 
with the control system stiffness seven times stiffer 
than normal. The effect of climb and descent almost 
disappeared. Since the stability is affected by the pitch- 
flap-lag coupling, a stiff control system minimizes the 
coupling effect, be it favorable or unfavorable. 

Precone 

Precone of the pitch axis directly alters the pitch- 
flap-lag coupling. The beneficial effect of lower precone 
has been evaluated many times. 1.2,7 Figure 20 shows 
the test confirmation of the favorable effect of the low 
precone. 



Besides the well-known stabilizing effect of pitch- 
flap coupling, the pitch-lag part of the total coupling is 
of utmost importance for the in-plane motions of the 
blade. Positive pitch-lag coupling (decrease of pitch as 
the blade leads forward, increase of pitch as the blade 
lags back) has a highly stabilizing effect on the lead-lag 
oscillations. Recent investigations!. 2 have shown that 
these coupling characteristics can be influenced by sev- 
eral hub and blade parameters, for example, by feather- 
ing axis precone, blade sweep, and control system 



Cyclic Trim 

An evaluation of the cyclic trim on the air- 
resonance stability was accomplished by varying the 
angle of incidence of the tail. The tail incidence angle 
was varied from 2 degrees through 45 degrees. As 
shown in Figure 21, the stability is insensitive to the 
range of cyclic-trim variation at constant thrust. This 
suggests that the steady 1/rev cyclic-pitch variation in 
forward flight can be ignored with respect to the air 
resonance. 



145 



H 

zu 

M 

as 

w 

nu 

K 
J W 
<C 0. 

a 



3.0 



2.0 



1.0 



V = 
N N0R = 
SYMBOL 
A 

O 



150 KNOTS 
100 PERCENT 

PRECONE 
DEGREES — 
1.5 DEGREES 
3.25 DEGREES 




200 



COLLECTIVE PITCH 
HOVER lg COLLECTIVE PITCH 





SXMBOt 


THRUST 


V = 150 KNOTS 


lg THRUST AT HOVER 


)R = 100 PERCENT 








o 


92 PERCENT 




A 


116 PERCENT 




a 


139 PERCENT 




^ 


162 PERCENT 



2.0 



- PERCENT 



M 

H 

os< 

ZH 
OjU 
§H 

ass 

H 
< « 

a a 

Oft 

E 




10 20 30 

STABILIZER ANGLE 



Figure 20. Effect of Blade Precone on Air- 
Resonance Stability 



Figure 21. Effect of the Incidence Angle of the 
Horizontal Tail on Air-Resonance 
Stability 



Conclusions 

1. The air-resonance mode stability is sensitive 1. 
to collective pitch (thrust). 

2. Air-resonance mode stability is also sensitive 
to climb and descent; that is, descent is stabilizing while . 
climb is destabilizing. 

3. The prime coupling term in the rotor system 2 - 
which causes the degradation of stability at high thrust 

is the induced drag. This coupling also provides the 
trend versus climb and descent. 

4. Air-resonance mode stability in lg level flight „ 
shows the rotor-power-curve trend with highly stable 
characteristics at high speed. 

5. The elastic-coupling behavior of the model 

rotor with normal control system stiffness is charac- 4, 

terized by a pitch-flap coupling (0. 4 degree pitch per 
degree flap) and a pitch-lag coupling (0.6 degree pitch 
per degree lag). 

6. High control system stiffness minimizes the 
flap-pitch coupling effectiveness and reduces the sensi- 5. 
tivity of the air-resonance stability to design parameters 
which are otherwise influential. 

7 . Less precone is stabilizing for a soft-in-plane 
hingeless-rotor system with an equivalent hinge sequence 

of pitch-flap-lag from inboard to outboard. 6. 

8. Variation in cyclic trim does not affect air- 
resonance stability. 

9. The testing technique to define air-resonance 7 * 
modal damping discretely at many operational conditions 
proved highly successful. Use of these methods to define 
modal damping, rather than defining only the boundaries, 
allows for a more definitive view of an aircraft's stability 
characteristics. 



References 

Burkam, J.E; , and Miao, W. , EXPLORATION OF 
AEROELASTIC STABILITY BOUNDARIES WITH A 
SOFT-IN-PLANE HINGELESS-ROTOR MODEL, 
Preprint No. 610, 28th Annual National Forum of 
the American Helicopter Society, Washington, D. C. , 
May 1972. 

Huber, H.B. , EFFECT OF TORSION-FLAP-LAG 
COUPLING ON' HINGELESS ROTOR STABILITY, 
Preprint No. 731, 29th Annual National Forum of 
the American Helicopter Society, Washington, D. C. , 
May 1973. 

Gabel, R. , and Capurso, V. , EXACT MECHANI- 
CAL INSTABILITY BOUNDARIES AS DETERMINED 
FROM THE COLEMAN EQUATION, Journal of the 
American Helicopter Society, January 1962. 

Lytwyn, R.T., Miao, W. , and Woitsch, W. , AIR- 
BORNE AND GROUND RESONANCE OF HINGE- 
LESS ROTORS, Preprint No. 414, 26th Annual 
National Forum of the American Helicopter Society, 
Washington, D. C, June 1970. 

Donham, R.E., Cardinale, S.V. , and Sachs, I.B., 
GROUND AND AIR RESONANCE CHARACTERIS- 
TICS OF A SOFT INPLANE RIGID ROTOR 
SYSTEM, Journal of the American Helicopter 
Society, October 1969. 

Woitsch, W. , and Weiss, H. , DYNAMIC BEHAVIOR 
OF A HINGELESS FIBERGLASS ROTOR, AIAA/ 
AHS VTOL Research, Design, and Operations 
Meeting, Atlanta, Georgia, February 1969. 

Hodges, D.H. , and Ormiston, R.A., STABILITY 
OF ELASTIC BENDING AND TORSION OF UNI- 
FORM CANTILEVERED ROTOR BLADES IN 
HOVER, AIAA/ASME/SAE 14th Structures, Struc- 
tural Dynamics , and Materials Conference, 
Williamsburg, Virginia, March 1973. 



146 



AH APPLICATION OF FLOGPET THEORY TO PREDICTION OF MECHANICAL INSTABILITY 

C. E. Hammond 

Langley Directorate 

U.S. Army Air Mobility R&D Laboratory 

NASA Langley Research Center 

Hampton, Virginia 



Abstract 

The problem of helicopter mechanical insta- 
bility is considered for the case where one blade 
damper is inoperative. It is shown that if the hub 
is considered to be nonisotropic the equations of 
motion have periodic coefficients which cannot be 
eliminated. However, if the hub is isotropic the 
equations can be transformed to a rotating frame 
of reference and the periodic coefficients elimi- 
nated. The Floquet Transition Matrix method is 
shown to be an effective way of dealing with the 
nonisotropic hub and nonisotropic rotor situation. 
Time history calculations are examined and shown 
to be inferior to the Floquet technique for deter- 
mining system stability. A smearing technique 
used in the past for treating the one damper inop- 
erative case is examined and shown to yield uncon- 
servative results. It is shown that instabilities 
which occur when one blade damper is inoperative 
may consist of nearly pure blade motion or they 
may be similar to the classical mechanical 
instability. 

Notation 



lag damping rate 

effective hub damping in x-direction 
effective hub damping in y-direction 
lag hinge offset 



e 

X b 

*i 

*x 

^b 
N 



second mass moment of blade about lag 
hinge 

lag spring rate 

effective hub stiffness in x-direction 

effective hub stiffness in y-direetion 

blade mass 

effective hub mass in x-direction 

effective hub mass in y-direction 

number of blades in rotor 

characteristic exponent corresponding to 
j th eigenvalue of the Floquet Transition 
Matrix 



x, y 



x c> v c 



*Q» 


^h 


*1> 


n 


e± 




% 




ii 





*i 



[A(t)] 

CD(t)] 
[Q] 

[0(t)] 
fs(t)| 



force acting on hub in x-direction 

force acting on hub in y-direction 

first mass moment of blade about lag 
hinge 

period of the periodic coefficients, 
T = &c/n 

coordinates of hub in rotating reference 
frame 

coordinates of rotor center of mass in 
fixed reference frame 

coordinates of hub in fixed reference 
frame 

coordinates of elemental blade mass dm 
in fixed reference frame 

lag deflection of i th blade 

defined by Equations (l8) 

defined by Equations (7) 

j th eigenvalue of the Floquet Transition 
Matrix 

defined by Equations (18) 

defined by Equations (7) 

distance from lag hinge to elemental 
blade mass dm 

azimuthal location of ith blade 

rotor speed 

defined by Equations (l8) 

defined by Equations (7) 

characteristic matrix, periodic with 
period T 

state matrix, periodic with period T 

Floquet Transition Matrix 

state transition matrix 

state vector 



Presented at the AHS/NASA Ames Specialists' Meet- 
ing on Rotorcraft Dynamics, February 15-15, 197 1 t-. 



L-9l£lv 



147 



The problem of mechanical instability of 
helicopters on the ground has been recognized and 
understood for many years. The analysis by Coleman 
and Feingold- 1 - has become the standard reference on 
this phenomenon although it was not published until 
many years after the first incidents of mechanical 
instability, or ground resonance as it is commonly 
known, were encountered on the early autogyros. 
The mechanical instability phenomenon is most com- 
monly associated with helicopters having articu- 
lated rotors; however, helicopters using the soft- 
inplane hingeless rotors which have became popular 
in recent years are also susceptible to this 
problem. Machines employing these soft-inplane 
hingeless rotors are also known to experience a 
similar problem, commonly known as air resonance, 
which occurs in flight rather than on the ground. 
The air resonance problem has received much atten- 
tion in recent years (see, e.g., Refs. 2 and 3). 

From the analysis of Reference 1 and others 
it is known that the ground resonance problem is 
due primarily to a coupling of the blade inplane 
motion with the rigid body degrees of freedom of 
the helicopter on its landing gear. These analyses 
have shown that with the proper selection of blade 
lag dampers and landing gear characteristics the 
problem of mechanical instability can be eliminated 
within the operating rotor speed range. All of the 
mechanical instability analyses conducted to date 
have one assumption in common - all blades are 
assumed to have identical properties. This is a 
reasonable assumption under ordinary circumstances j 
however, the U.S. Army has a requirement on new 
helicopters which invalidates this assumption. The 
requirement is that the helicopter be free from 
ground resonance with one blade damper inoperative. 
As will be shown later, this one blade damper inop- 
erative requirement has a serious impact on the 
classical method of analyzing a helicopter for 
mechanical instability. Further, there is at pres- 
ent no published method available for treating the 
case where each of the blades is permitted to have 
different properties. Thus the designer is faced 
with the dilemma of trying to satisfy the require- 
ment with an analysis method in which one of the 
basic assumptions is severely violated. 

Two methods have been used to circumvent this 
difficulty. The first of these involves a physical 
approximation so that the classical analysis 
becomes applicable. In this approach all blades 
are still assumed to have identical lag dampers 
even when one blade damper is removed, but the 
value of each of the dampers is reduced by the 
amount ci/W where N is the number of blades 
and c^ is the original lag damper rate. As can 
be seen, with this approach a system is analyzed 
which is quite different from the actual situation 
of a rotor with no damping on one blade. The sec- 
ond method which has been used is to reformulate 
the equations of motion allowing for differing 
blade characteristics and to obtain the stability 
characteristics of the system using a time history 
integration of the equations. This second approach 
has the drawback that interpretation of stability 
characteristics from time history calculations is 
often difficult and open to question. The method 



will yield correct results, however, provided the 
equations are integrated over a sufficiently long 
time period. 

The purpose of this paper is to present a 
method of obtaining the mechanical stability char- 
acteristics directly for a helicopter operating on 
the ground with one blade damper inoperative. As 
will be shown later, the equations governing the 
motion of this system have periodic coefficients. 
This fact suggests the use of Floquet theory as the 
means for determining the stability characteristics 
of the system. In the following, the one-damper- 
inoperative problem is formulated and the resulting 
equations are solved using the Floquet Transition, 
Matrix method described by Peters and Hohenemser. 
Results obtained using this method are compared 
with results obtained from the two previously used 
methods and recommendations are made concerning 
the future use of the three methods described. 

Equations of Motion 

The equations of motion for the mechanical 
instability problem will be formulated using an 
Eulerian approach. It will be assumed, as is done 
in Reference 1, that the helicopter on its landing 
gear can be represented by effective parameters 
applied at the rotor hub. It will be further 
assumed that only inplane motions of the hub and 
blades are important in determining the ground 
resonance characteristics of the helicopter. Thus 
the degrees of freedom to be considered consist of 
two inplane hub degrees of freedom and a lead- lag 
degree of freedom for each blade in the rotor. The 
mathematical model to be used in the analysis is 
shown in Figure 1. Note that in the figure only a 
typical blade is shown. The analysis will be 
formulated for a rotor having IT blades, and each 
blade is assumed to have a rotational spring and 
damper which act about the lag hinge. 

The blade equations are developed by summing 
moments about the lag hinge. The coordinates of 
the elemental mass dm in the fixed system are 



x. = x, + e cos +. + p cos(+. + £.) 
y i = y h + e sin 1/ ± + p sinC^ + t, ± ) 



(1) 



where 



t ± = At + 2fl(i - l)/N i = 1,2,..., H 

These expressions can be differentiated twice with 
respect to time to yield the accelerations exper- 
ienced by the differential mass 

x 1 = \ - en 2 cos t ± - P (n + C i ) 2 cos(t i + q)j 

- pl ± sinC^ + C ± ) 
y x = y h - efl 2 sin t ± - p(fl + ^^sin^ + t, ± )\ 



+ (j[ t cosC^ + S ± ) 



Using D'Alembert's principle the Bummation of 
moments about the lag hinge can be written as 



(2) 



148 



Jp sin (tj + 5 i )x jL 3m - fp cos (+ i + ^y^cbn 



k.£, - Q.X. = 
iM i b x 



i = 1,2,. ..,U (3) 



where the integrals are evaluated over the length 
of the blade. Introducing the expressions for 
5c. and y. and defining the following 



%- So 



dm 



dm 



(4) 



the blade equations become 

1^ + ea\ sin q - S^ sin(+ i + ^ 

- y h cosC^ + 5i)] + \^ + \i % = o (5) 

i - 1,2,..., N 

If small displacements are now assumed the blade 
equations may. be linearized to obtain 

k + \k + (a> °i + ^o^i = (%/ e > [2 h sin +i 



- y^ cos t.^ 



i=l,2,...,N (6) 



where the following parameters have been introduced 

1i = CA 

Under the assumptions stated earlier the hub 
equations of motion can be written directly as 



m x, + c x, + k x, =P 
x n x n x n x 



Vh + C A + Vh = P y 



(8) 



where the coefficients on the left side of these 
equations are the effective hub properties in the 
x- and y-directions, respectively. The determina- 
tion of these properties depends on an extensive 
knowledge of the helicopter inertial character- 
istics and the stiffness, damping, and geometrical 
characteristics of the landing gear system. These 
properties may be determined either by ground shake 
tests of the helicopter, as suggested in Refer- 
ence 1, or by direct calculations. The right-hand 
side of the above equations are the forces acting 
on the hub due to the fact that the rotor is 
experiencing accelerations in the x- and y- 
directions. If the accelerations of the rotor 
center of mass are x and f c , respectively, the 
P x and Py are given by 



P., 



p y = "^c 



(9) 



The equations as written also indicate that in the 
absence of the rotor the hub degrees of freedom are 



uncoupled. This is an approximation, but it is 

an assumption made in Reference 1 and one generally 

used in helicopter mechanical stability analyses. 

If all blades in the rotor are assumed to have 
the same mass distribution, the coordinates for the 
total rotor center of mass may be written as 



= *h + I E x i 



1=1 c 



(10) 



y c = ^h + | E y i 



1=1 



where x^ and y^ are the coordinates of the 
individual blade center of mass, measured with 
respect to the hub. If the center of mass of the 
ith blade is a radial distance p c from the lag 
hinge 



x i c = e cos \ + P c eos^ + t. % ) 

y ± = e sin ^ + P c sin^ + ^) 

Making the observation that, for H > 1 

N N 

£ cos t k = E sin t k 



(11) 



k=l 



k=l 



the rotor center of mass coordinates become 

H 

x c = *n - ( p c / N) £ k sin *i 

N 

y c = y h + (p c /n) £ t, ± cos ^ 



(12) 



These expressions may now be differentiated twice 
with respect to time and the forces P x and P y 
obtained as 

P x = - Bn b S h +8 b ^k- a \ )Bin V 2 < cos *i 

J. r- 2 

P y = -H^h-^ SPi-^i^ 8 +i- 2Q 5i Sin *i 

(15) 

The hub equations of motion thus become 

(m x + Hrn^ + c^ + k^ = 



h r.. p 

s^ £ M x - n^)sin t ± + 20^ cos t. 
1=1 L 

(a y + ^)r h + y h + y h = 



(i»0 



-% E (?i - ^iJcos ^ - 2fl£ ± sin ^ 

The equations of motion for the system thus con- 
sist of (N + 2) coupled second-order differential 



149 



equations with the coupling terms having periodic 
coefficients. The periodic coefficients arise 
because the hlade equations are written in a rotat- 
ing reference system whereas the hub equations are 
in a fixed system. As is shown in the Appendix, 
if all the blades have identical lag springs and 
lag dampers, the periodic coefficients may he 
eliminated through the use of multiblade coordi- 
nates. The effect of these coordinates is to 
transform the hlade equations from the rotating to 
the fixed system of reference. The resulting con- 
stant coefficient system of equations is the set 
normally solved in the classical ground resonance 
analysis. As is shown, however, if the blades are 
allowed to have different lag springs and dampers, 
the periodic coefficients cannot be eliminated in 
the usual manner. 

An alternative does exist, however, for 
eliminating the periodic coefficients even when 
the blades are allowed to have differing character- 
istics. The alternative consists of transforming 
the hub equations into the rotating system of 
reference. In order to eliminate the periodic 
coefficients using this approach, the additional 
assumption must be made that the hub is isotropic. 
That is 



y + r^y + (a£ - fi 2 )y + 2f2x + Oi^x 

N r 
- - v l E KCj-flPCj)** f^Q-l)- 2Qi 3 sin §5(3-1) 

(17) 
where the following parameters have been introduced 



v h = V (m x + ^V 

\ - c x/ (m x + fc b ) 



(18) 



Introducing the rotating coordinates into the blade 
equations, Equations (6), results in 

2 2v 



l i + "A + << + ° < K i 



= (#•) 



(x - fl x - 2f2y)sin 21 (j - l) 

a 



- (y - fi 2 y + 2&) cos |L(j - l) 

A = 1,2, 



(19) 



.,N 



c => c 
x y 



This is the approach used in Reference 1 for treat- 
ing the two-bladed rotor which is another case 
where the periodic coefficients in the equations 
of motion cannot be eliminated by transforming the 
blade equations to the fixed system. 

The transformation from fixed to rotating 
coordinates is given by 



x = x, cos fit + y, sin Sit . 
y = -x, sin At + y. cos fit ' 



(15) 



Differentiating these expressions allows the fol- 
lowing identities to be established 

x. cos fit + y. sin fit =* x - fiy 

-x. sin fit + y. cos fit = y + & 

- 2- * 

x. cos fit + y, sin fit = x - fi x - 2f!y 



-x. sin fit + y. cos fit = y - fi y + 



2fix 



The hub equations in the rotating system are then 
obtained by appropriate combinations of the xjj 
and yjj equations, Equations (l4). The resulting 
equations are given below 

- - 2 2 

x + TjjX + (a£ - fi )x - 2fiy - fiiyr 

= v l E p j - fl 2 ^)sin %{ j - 1) + 2fl£j cos §£( j - 1)1 



Since modern helicopters do not in general 
have isotropic hubs, the above equations can only 
be used to approximate the effects of a noniso- 
tropic rotor. They are, however, easily solved for 
the stability characteristics of the system and 
thus they might be used to obtain a first approxi- 
mation to the mechanical stability boundary for a 
helicopter with one blade damper inoperative. 

Prom the foregoing discussion it can be seen 
that if either the rotor or the hub is isotropic, 
the mechanical stability characteristics of the 
system may be obtained using conventional tech- 
niques. If both the rotor and hub sire nonisotropic 
the equations of motion of the system contain 
periodic coefficients and thus the standard eigen- 
value techniques cannot be used to determine 
whether the system is stable or unstable. It is 
the purpose of this paper to demonstrate that 
Floquet theory can be used to analyze this general 
situation of a nonisotropic rotor coupled with a 
nonisotropic hub. 

Solution of the Equations 

If the periodic coefficients in the equations 
of motion are eliminated by assuming either an 
isotropic rotor or an isotropic hub, the stability 
of the system can be determined using standard 
eigenvalue techniques. The general case of a 
nonisotropic rotor coupled with a nonisotropic hub 
will be treated using Floquet techniques as 
described by Peters and Hohenemser,* and Hohenemser 
and Yin. 5 A brief description of the technique 
will be presented here for the sake of completeness. 

In state vector rotation the free motions of 
the system may be written as 



B = [D(t)] 8 



(20) 



150 



where the state variables for the problem being 
considered consist of 

» • » 
?!» i z > •••> 5 H * V V ^1' ^2' ■••' 5 H' V K 

and the equations which describe the motions of the 
system are Equations (6) and {lk) . The matrix 
[D(t)] is periodic with period T and for the 
mechanical stability problem T = 2rt/fl. 

Floquet's theorem states that the solution to 
the above system of equations has the form 



ill -CA(t)]U>«»)*| 



(21) 



where [A(t)] is the characteristic matrix and is 
also periodic with period T. The column of 
initial conditions jZ(o) is used in determining 
\a\ as 



a}- [A(0)]" 1 JZ(0)J 



(22) 



The matrix [A(o)], the modal damping A, and the 
modal frequency cu are determined from the Floquet 
Transition Matrix [Q] which is defined by the 
equation 



Z(T) - [Q] 2(0) 



(23) 



for all sets of initial conditions Z(o)J . It is 
shown in References k and 5 that the eigenvalues 
A* of the matrix [Q] can be used to determine Aj 



and cui since 



Aj - e ( V la3 J )T 



(210 



and the modal matrix of [Q] is just [A(o)]. The 
characteristic matrix [A(t)3 is then Bhown to be 
given by 

CA(t)] - [0(t)][A(O)][e- ( ^ +ia)) *l (25) 

where the state transition matrix [0(t)] is defined 
ty 



B(t)J = C0(t)]jZ(o) 



(26) 



The characteristic multipliers A., of the 
system are uniquely defined since the matrix [Q] 
is realj however, only the real parts of the 



characteristic exponents 
defined uniquely since 



"3 



Aj + to, 



^■?. (ta hl +lar sV 



(27) 



The imaginary part can only be determined within an 
integer multiple of 2jr/T. This indeterminacy of the 
cuj causes no particular difficulty if one is only 
interested in the stability of the system. However, 
if one is interested in understanding the mechanism 
involved in any instability which might be found, 
this indeterminacy can be quite troublesome. 

The Floquet Transition Matrix which is the 
basic element needed in the stability analysis is 
easily determined by a numerical integration of 
the equations of motion over one period T. If 
one desires to compute the characteristic functions 



[A(t)] the matrix [0(t)] is saved at each time 
point in the numerical integration to obtain [Q] 
For the calculations of this paper, the fourth 
order Runge-Kutta method with Gill coefficients' 
was used for the numerical integration. 



.6 



A comment is in order concerning the charac- 
teristic functions [A(t)]. The matrix [A(t)] is 
a complex valued matrix and is determined at as 
many time points as desired. The computation of 
these functions can be relatively expensive and 
intepretation can be difficult. The interpreta- 
tion is made easier by the procedure outlined in 
Reference 5 for converting the complex functions 
into real functions which may be plotted as func- 
tions of time. The scheme used is essentially the 
same as that used when it is desired to plot as a 
function of time the modes of a system having con- 
stant coefficients. That is, for a conjugate pair 
of characteristic exponents 

P j = h + iCD j 



P j = *j 



io>. 




the characteristic functions are also conjugate 
pairs. Thus the real modal function column for 
this conjugate pair of characteristic exponents 
will be given by 

{« (t)} - JA (t^ty*^* + |I d (t)|e^J- 1CU J 3t 

(28) 

A,(t)j is the jth column of [A(t)] and 
Is the complex conjugate of this column, 
he purpose in performing these manipulations is 
to be able to plot the modal functions to deter- 
mine the relative magnitudes and phases of the 
various degrees of freedom in each mode. A discus- 
sion of this technique as it applies to constant 
coefficient systems is given by Meirovitch.T In 
this paper the exp(Ajt) is omitted from the above 
equation since it is simply a constant which multi- 
plies each component of the mode and causes each 
component to damp at the same rate. Thus the 
plots of the characteristic functions which are 
presented later in the paper will appear to be 
neutrally damped. 

In making the calculations for this paper it 
was found that the output from the calculation of 
the modal functions became so voluminous and these 
calculations became so expensive that the modal 
functions were only computed for selected points. 
Generally a sweep of rotor speed was made and the 
results examined. If an unstable region was indi- 
cated the rotor speed corresponding to the maximum 
positive Aj was rerun and the modal functions 
calculated. 

Discussion of Results 

In order to demonstrate the application of 
the above-mentioned techniques and to obtain a 
general understanding of the effect of one blade 
damper inoperative on mechanical stability, a set 
of parameters were chosen. The parameters in the 



151 



mechanical stability analysis were chosen so as to 
be in the general range of interest for a single 
rotor helicopter and were such that the system was 
stable with all dampers functioning up to a rotor 
speed of ^00 rpm. The parameter values chosen for 
the calculations are shown in Table 1. 

The parameters presented in Table 1 correspond 
to an isotropic rotor and a nonisotropic hub. In 
the following discussion results are presented for 
the case of an isotropic hub coupled with a non- 
isotropic rotor and a nonisotropic hub coupled 
with an isotropic rotor as well as the case of 
interest which involves a nonisotropic hub coupled 
with a nonisotrpic rotor. When an isotropic hub 
is mentioned, this means that the hub parameters 
in both the x- and y-directions were assigned the 
values shown in Table 1 for the x-direction. An 
isotropic rotor implies that all dampers are 
operational and a nonisotropic rotor is meant to 
indicate that the lag damper has been removed from 
blade number 1. The analysis has been formulated 
in such a way that any number of blade lag dampers 
or lag springs may be removed to make the rotor 
nonisotropic. The results presented here, how- 
ever, only involve the removal of the lag damper 
from one blade. 

The ease of an isotropic hub was first run in 
an effort to become familiar with the nonisotropic 
rotor results before proceeding with the more 
complicated Floquet analysis. The isotropic hub 
permits the equations to be transformed into the 
rotating reference frame and results in a system 
of equations with constant coefficients, Equa- 
tions (l6), (l7)j and (l9)> even with a noniso- 
tropic rotor. 

Figure 2 shows the results of the calculations 
for the isotropic hub with all blade dampers work- 
ing. Note that since the equations were solved in 
the rotating system, the frequencies in the lower 
portion of Figure 2 are plotted in the rotating 
system. The numbers attached to the different 
modes in Figure 2 and in subsequent similar figures 
have no significance other than' to provide a label 
for the various modes. In Figure 2 the dashed 
lines represent the uncoupled hub modes. The 
uncoupled rotor modes follow along the curves 
labeled 1,2 which also represent, in the terminol- 
ogy of Beference 5, the rotor collective modes. 
Note that the uncoupled blade frequencies are zero 
for rotor speeds less than about 65 rpm. This is 
due to the fact that the blades are critically 
damped for these low rotor speeds. At the higher 
rotor speeds modes 3 and k are essentially rotor 
modes and modes 5 and 6 are essentially hub modes. 
At the lower speeds, however, due to the coupling 
between rotor and hub, mode h changes to a hub 
mode and mode 5 changes to a blade mode. Note 
from the damping plot that all the modes indicate 
stability over the entire rotor speed range. 

The results for one blade damper inoperative 
and an isotropic hub are plotted in Figure 3. 
Note that the removal of a blade damper has caused 
the appearance of a mode which was not present in 
Figure 2, namely the mode labeled 3 in Figure 3, 



and that this mode exhibits a mild instability . 
between 160 and 200 rpm. At rotor speeds below 
about 100 rpm this mode has a frequency which cor- 
responds to the uncoupled frequency of the blade 
which has no damper. At rotor speeds above 100 rpm 
this mode begins to deviate in frequency from the 
uncoupled frequency. Another interesting point is 
that mode 1 in Figure 3 is precisely the same as 
the collective modes of Figure 2, and in Figure 3 
there is only one such mode. Thus it appears that 
the unstable mode in Figure 3 has evolved from one 
of the two collective modes shown in Figure 2 
because of the removal of one of the blade dampers. 

A time history calculation was made for the 
point of maximum instability in Figure 3 which 
occurs at approximately 175 rpm- The results of 
the time history calculation are shown in Figure h. 
These results were obtained using the same inte- 
gration scheme as that used for generating the 
Floquet Transition Matrix. The top portion of the 
figure represents the individual blade lag motions 
whereas the lower portion represents the hub 
response in the x- and y-directions. Note from 
the figure that each of the degrees of freedom was 
given an initial displacement but the initial 
velocities were zero. The equations were inte- 
grated for 17 rotor revolutions. The figure indi- 
cates the blades which have lag dampers are .well 
damped, but the blade on which the damper is 
inoperative experiences large lag excursions. 
Also, the hub motions, although not large, do not 
appear to have a high degree of damping. From the 
time history one would conclude that the system is 
stable since the motions of the various degrees of 
freedom do not appear to be increasing in ampli- 
tude with increasing time. The eigenvalue analysis 
has shown, however, that an instability exists. 
The problem with the time history calculations is, 
of course, that the equations of motion have not 
been integrated over a sufficiently long time 
period for the initial conditions chosen. Herein 
lies the difficulty with using the time history 
approach for calculating the stability character- 
istics of systems. One can never be sure if a 
sufficiently long integration period has been 
used, and the choice of initial conditions which 
will minimize the integration time required is a 
trial and error process. It has been observed on 
an analog computer that for the ground resonance 
problem the choice of initial conditions has a 
strong bearing on the conclusion inferred from the 
time history traces. The time history integration 
is also much more time consuming on the digital 
computer than the eigenvalue analysis. The time 
to generate Figure k- which is for only one rotor 
speed was much greater than the time required to 
generate the eigenvalue results for all of Fig- 
ure 3- It is thus concluded that whenever it is 
at all possible the eigenvalue approach to sta- 
bility calculation is to be desired over the time 
history approach. 

Having examined the case of one blade damper 
inoperative an an isotropic hub, the next logical 
step is to examine the more realistic situation of 
a nonisotropic hub. Before examining the one 
damper inoperative situation it was first desired 



152 



to confirm that the system was stable with all 
dampers working. The modal damping and frequency 
of the -various modes with all dampers working and 
a nonisotropic hub are shown in Figure 5- As can 
be seen from the damping plot, all the modes are 
stable. In this ease the equations of motion are 
solved in the fixed frame of reference and hence 
the frequencies are plotted in this frame. The 
dashed lines on the frequency plot represent the 
uncoupled system: the horizontal dashed lines 
being the hub modes and the slanted dashed lines 
being the rotor modes. Note that because the rotor 
modes become critically damped at low rotor speeds 
the two uncoupled rotor frequencies come together 
before reaching the origin. The uncoupled rotor 
lines also represent the collective modes for the 
rotor. These modes are completely uncoupled from 
the other modes and hence are not included in the 
eigenvalue analysis of the nonisotropic hub 
coupled with an isotropic rotor. The damping for 
the collective modes is exactly the same as that 
shown for modes 1,2 in Figure 2. 

The validity of the Floquet analysis was 
verified by comparing results from this analysis 
with results from both the rotating system analysis 
(isotropic hub) and from the fixed system analysis 
(isotropic rotor). In each case the results from 
the Floquet analysis were identical to results 
from the other analyses. 

Having thus established the validity of the 
Floquet analysis, results were obtained for the 
nonisotropic hub and one blade damper inoperative. 
These results are shown in Figure 6. Note that 
these results are very much similar to those shown 
in Figure 5 except that, as was the case with the 
isotropic hub and one blade damper inoperative, 
there are additional modes introduced. Also indi- 
cated is a relatively strong instability between • 
210 and 305 rpm. The frequencies of the addi- 
tional modes which are introduced correspond, at 
low rotor speeds, to the frequencies of the 
uncoupled blade which has no damper. In the rotor 
speed range where the instability occurs, however, 
the frequency deviates from the uncoupled value as 
indicated by the mode labeled 3- In this range 
and at higher rotor speeds the mode labeled 5 is 
nearer the uncoupled blade frequency. It thus 
appears that for this case the instability is more 
a coupled rotor hub mode than a pure blade mode as 
was indicated for the isotropic hub. 

This conjecture is further strengthened by an 
examination of the modal functions. The modal 
functions for a rotor speed of 255 ^V^t which' is 
the point of maximum instability, are shown in 
Figure 7. The functions are plotted over a time 
period corresponding to one rotor revolution. Note 
from this figure that blade 1, the blade without a 
damper, has a significantly higher contribution to 
the mode than the other blades. Also from the 
plot of hub response it can be seen that the par- 
ticipation of the lateral hub degree of freedom, 
which has the higher of the uncoupled hub fre- 
quencies shown on Figure 6, is considerable. It 
is thus concluded from Figures 5 and 6 that the 



one damper inoperative situation can lead to a 
classical mechanical instability. 

Time history traces for this same condition 
are shown in Figure 8. These traces show the same 
general trends as observed in the case of the 
isotropic hub, that is, a large response of the 
blade having no damper and moderate responses from 
the other blades and the hub degrees of freedom. 
Again the time history traces are inconclusive 
regarding the stability of the system. 

One of the methods used in the past for 
treating the one blade damper inoperative case 
involves a smearing of the total blade damping. 
The reasoning for this approach is as follows. 
If the rotor has N blades then the total damping 
available in the rotor is Nc^ where ci is the 
damping on one blade. If one damper is removed, 
the total damping becomes (N - l)c^. Thus, using 
this approach, each blade in the rotor would be 
treated as if it had a lag damper equal to 
Ci(N - 1)/N. 

After an examination of the preceding one 
damper inoperative results it would be expected 
that this approach would lead to unconservative 
results. This is due to the fact that the insta- 
bilities encountered in the previous results 
involved large motions of the blade which had no 
damper. The smearing technique results in damp- 
ing, which is not greatly different from the 
original value, being applied to each blade and 
thus the true situation is not adequately modeled. 

To illustrate this method, the nonisotropic 
hub case was analyzed using the smearing approach. 
The results from these calculations are shown in 
Figure 9- Note that although mode 3 becomes 
lightly damped the system remains stable through- 
out the rotor speed range considered. The fact 
that mode 3 approaches instability is attributable 
to the fact that this mode was not heavily damped 
in the original calculations. A run of the iso- 
tropic hub ease, where all the modes were origi- 
nally well damped, indicated that the smearing 
technique resulted in well damped modes for one 
blade damper removed. The smearing technique is 
thus not recommended for treating the one blade 
damper inoperative situation since it leads to 
unconservative results. 

Since one way for eliminating the classical 
mechanical instability is to increase the blade 
damping, it was decided to attempt this approach 
on the instability indicated in Figure 6. The 
approach was to leave the damping identically zero 
on one blade and increase the damping on the 
remaining three blades. The results of this series 
of calculations are shown in Figure 10 where the 
region of instability is presented as a function 
of blade lag damping and rotor speed. As can be 
seen from the figure, increasing the blade damping 
on three of the blades has very little effect on 
the stability boundaries when one blade has zero 
damping. This result was somewhat expected since 
from the previous calculations it was observed 



153 



that the blade with zero damping responds more or 
less independently of the other blades in the 
rotor . 

During the increased damping calculations no 
attempt was made to determine whether or not the 
nature of the instability had changed. That is, 
whether the instability had changed from one 
involving both blade and hub motion to one con- 
sisting of primarily blade motion with only small 
amounts of hub motion, Further delving into pos- 
sible corrective actions for the instability which 
occurs with one blade damper inoperative was 
beyond the scope of this paper and thus more 
research is needed to determine how the instability 
may be eliminated. 

Conclusions 

There are several conclusions which may be 
inferred from the preceding results. First of all, 
the fact that a helicopter is free from mechanical 
instability with all blade d amp ers working does 
not guarantee that it will be free of instabilities 
with one blade damper inoperative. The instability 
encountered with one blade damper inoperative may 
be a blade mode instability or it may be the 
classical mechanical instability. 

The Floquet Transition Matrix method can be 
used effectively in examining the mechanical sta- 
bility characteristics of helicopters with one 
blade damper inoperative. When both the hub and 
rotor are considered to be nonisotropic, the equa- 
tions of motion contain periodic coefficients and 
the Floquet approach provides an efficient means 
for dealing with this situation. Since the 
Floquet approach yields the stability character- 
istics directly, it furnishes a more desirable 
approach to stability problems than time history 
calculations . 

Time history calculations can lead to erron- 
eous conclusions relative to the determination of 
system stability. The erroneous conclusions stem 
primarily from the fact that the time history 
calculations require considerable computer time 
and the tendency is to integrate over as short a 
time period. as possible. Thus, if the initial 
conditions are not chosen properly, the time 
history traces may still contain transients when 
the integration is terminated. The time history 
approach to stability problems is thus recommended 
only when no other recourse is available, and then 
several different combinations of initial condi- 
tions and integration periods should be examined 
before making a conclusion regarding stability. 

The smearing approach which has been used in 
the past for treating the one blade damper inop- 
erative situation leads to unconservative results. 
Therefore, this method is considered to be an 
unacceptable means for determining stability under 
these conditions. 



References 



OF SELF-EXCITED MECHANICAL 0SCILIAT3DUS OF 
HELICOPTER BQTOBS WITH HINGED BIADES, NACA 
Report 1351, 1958. 

2. Donham, R. E., Cardinale, S. V., and Sachs, 

I. B., GROUND AND AIR RESONANCE CHARACTERISTICS 
OF A SOFT IN-PLANE RIGID-ROTOR SYSTEM, Journal 
of the American Helicopter Society , Vol. 14, 
No. 4, October 1969, pp. 33- J H- 

3. Lytwyn, R. T., Miao, W., and Woitsch, W. , 
AIRBORNE AID GROUND RESONANCE OF HTNGELESS 
ROTORS, Journal of the American Helicopter 
Society, Vol. 16, No. 2, April 1971, PP- 2-9- 

k. Peters, D. A., and Hohenemser, K. H., APPLICA- 
TION OF THE FLOQUET TRANSITION MATRIX TO 
PROBLEMS OF LIFTING ROTOR STABILITY, Journal 
of the American Helicopter Society , Vol. 16, 
No. 2, April 1971, PP- 25-35- 

5. Hohenemser, K. H., and Yin/s. K., SOME 
APPLICATIONS OF THE METHOD OF MCJLTIBLADE 
COORDINATES, Journal of the American Helicopter 
Society , Vol. 17, No. 3, July 1972, pp. 3-12. 

6. Carnahan, B., Luther, H. A., and Wilkes, J. 0., 
Applied Numerical Methods , John Wiley & Sons, 
Inc., New York, 1969. 

7. Meirovitch, L., Analytical Methods in Vibra- 
tions, The Macmillan Company, New York, 1967, 
p. 411. 

Appendix 

If the rotor is considered to be isotropic 
the periodic coefficients appearing in the equa- 
tions of motion can be eliminated through the use 
of multiblade coordinates similar to those 
described in Reference 1. These coordinates 
essentially transform the blade degrees of freedom 
into a fixed reference frame. The transformations 
are given by 

N 



Si - E ?i sin *i 
i=l 

N 

l II = ^ ^i COS *i 

i=l 



(Al) 



Differentiating these expressions leads to 
the establishment of the following identities 

N 






i=i 



(A2) 



154 



£ L sin t = 6 - air - 2fl| 



i=l 



II 



£ '( ± cos t ± = in - 21 n + 2J2IJ 



i=l 



2t can be seen from these identities that the 
transformation is made by multiplying the blade 
equations, Equations (6), by either sin % or 
cos ilr^ and adding the equations. Crucial to this 
operation is the ability to remove the i\± and 
ccg. from the summations. This can only be done 
if all the blades have identical lag springs and 
lag dampers. If one or more of the blades have 
differing characteristics, the tjj and/ or og^ 
cannot be factored from the summation and hence 
the identities above cannot be applied. Thus, if 
one or more of the blades are permitted to have 
different lag springs or lag dampers, the periodic 
coefficients cannot be eliminated using the pro- 
cedure described in this Appendix. 

If Equations (6) are first multiplied by 
cos i|f^ and summed and then multiplied by sin f^ 
and summed, the following equations are obtained 
after introduction of the identities (A2) 



6 n + \ ! n 



2 2 (1 - v 2 ) - a^Ji^ + Sfilj. 



+ an^ = (v^/e) 
- jr h E cos 2 ti 



i=l 



% 2 sin +i cos +1 

. 1=1 



(A3) 



*t + I** 



i*! 



n 2 (i-v2)-<|| T - 2 «i IT 



Xfc E sin +J.. 

1=1 



"Mil = ( v o/ e) 

B - 1 

^ £ sin +i cos *i 
i=i j 



Making the following observations that for N > 2 

N 
^ sin f . cos \|r = 

i=l 

J, P i P 

X) cos^ i|r = X) sin + = N/2 

1=1 1 i=l X 



the equations become 



'II 



*i* 



II 



fl 2 (l 



- v 2 ) - c 2 ], 



Ijj + 20IJ. 



+ Sit) i l I = -(Hv^/2e)3r h 



*I + \h - [ fl2(l " V o } " 4J' 



(A*) 



Oil"! 



2J2| 



II 



^i 6 II » K/ 2e >*h 



These two equations describe the rotor motions in 
the fixed frame of reference. In terms of the 
variables described by Equations (Al) the hub 
equations, Equations (1*0, became 



(m x + Bn^ + c^ + kx^ = Sjj 
(m y + ""tX + Vh + Vk = " S ^H 



(A5) 



The stability of the rotor-hub system can now 
be determined using Equations {Ak) and (A5) which 
have constant coefficients. This set of equations 
or a set similar to it is the one normally used in 
helicopter mechanical stability analyses. 

As a final observation, note that if. the 
blade equations, Equations (6), are simply summed, 
the following equation 

«o + Vo + (£D °i + ^X - ° (A6 > 



is obtained, where 



6 o = Z 5 



(A7) 



1=1 



This equation represents the rotor collective mode 
and it may be observed that this equation is com- 
pletely decoupled from the hub degrees of freedom. 
Hence, the collective mode cannot influence the 
stability of the system and it is therefore not 
normally included in the mechanical stability 
analysis . 



155 



TABLE 1. PARAMETERS USED IN THE SAMPLE CALCULATIONS 



Number of blades 

Blade mass, m. 

Blade mass moment, S, 

Blade mass moment of inertia, X. 

Lag hinge offset, e 

Lag spring, k^ 

Lag damper, c. 

Hub mass, m 

Hub mass, m 

Hub spring, k 

Hub spring, k 

Hub damper, c 

Hub damper, c 



6.5 slugs (9^.9 kg) 

65. slug-ft (289. 1 kg-m) 

800.0 slug-ft 2 (1084.7 kg-m 2 ) 

1.0 ft (O.30W m) 

0.0 ft-Ib/rad (0.0 m-N/rad) 

3000.0 ft-lb-sec/rad (^67. 5 m-N-s/rad) 

550.0 slugs (8026.6 kg) 

225.0 slugs (3283.6 kg) 

85000.0 lb/ft (121KA81.8 N/m) 

85OOO.O Vo/ ft (I2i«)li8l.8 N/m) 

3500.0 Ib-sec/ft (5IO78.7 N-s/m) 

1750.0 lb- sec/ft (25539.3 N-s/m) 




Figure 1; Mathematical representation of the rotor 
and hub. 



200 
ROTOR SPEED, RPM 



3 4 

ROTOR SPEED, Hz 



Figure 2. Modal damping and frequencies for iso- 
tropic hub, all blade dampers working. Fre- 
quencies plotted in the rotating system. 



156 




MODAL 
FREQUENCY, 
radlsec 12 




200 

ROTOR SPEED, RPM 
J_ 



3 4 

ROTOR SPEED, Hz 



Figure 3- Modal damping and frequencies for iso- 
tropic hub, one blade damper inoperative. Fre- 
quencies plotted in the rotating system. 



1 

1 
2 
■3 
-4 
28 

24 
20 

MODAL 16 
FREQUENCY, 
rad/sec 12 



- 


/ / 




s>y 


s S 
s x 
/ / 


- 


/ 




V-2 


_ 


"V >*^ 






Z" 1 


-V- — :*S^ — 








- / 


1 1 






1 



100 







1 



200 
ROTOR SPEED, RPM 

_L_ 



300 



400 



3 4 

ROTOR SPEED. Hz 



Figure 5- Modal damping and frequencies for non- 
isotropic hub, all blade dampers working. Fre- 
quencies plotted in the fixed system. 



BLADE 
RESPONSE 




HUB 
RESPONSE 



1, 

.5 



-.5 

-1 
1 

.5 


-.5 
-1 



1 



BLADE 1 



J-± 



-^^t£Pr< 



J L 



J I I _J I L 



TIME, sec 

Figure k. Time history calculations for isotropic 
hub, one blade damper inoperative, fi = 175 rpm. 




24 
20 

MODAL 16 
FREQUENCY, 
rad/sec 12 



- 


6 A // 

4 y\r 


4?/ 


gr- 




£r 


- 


V-2 




" <& 


r» 






- /Z^ 


1 1 


1 1 



100 



200 
ROTOR SPEED, RPM 



300 



400 



3 4 

ROTOR SPEED, Hz 



Figure 6. Modal damping and frequencies for non- 
isotropic hub, one blade damper inoperative. 
Frequencies plotted in the fixed system. 



157 



BLADE 
RESPONSE 



HUB 
RESPONSE 



1.000 
.500 


-.500 

-1.000 

1.000 

.500 



-.500 
-1.000 



EIGENVALUE - 0.32508 +i 8.04396 



J I L 



J I I L 



_1 I 



I I I I I I L 



J I I 1 







.050 .100 .150 .200 .250 .300 

TIME, sec 



MODAL 16 

Figure 7- Modal functions for nonisotropic hub, frequency. 
one blade damper inoperative, SI = 255 rpm. rad/sec 12 



BLADE 
RESPONSE 




HUB 
RESPONSE 






TIME, sec 

Figure 8. Time history calculations for noniso- 
tropic hub, one blade damper inoperative, 
J) = 255 rpm. 




200 

ROTOR SPEED, RPM 



3 4 

ROTOR SPEED, Hz 



Figure 9. Modal damping and frequencies obtained 
for nonisotropic hub, one blade damper inopera- 
tive, using the smearing technique. 



m-N-s/rad 
22000 



19000 

16000 

BLADE 
LAG 13000 
DAMPING 

10000 
7000 



4000 



fHb-sec/rad 
15000 r 



_ 12000 



9000 



6000 



b 3000 



STABLE/^ UNSTABLE ^STABLE 



fr 



200 250 300 350 
ROTOR SPEED, RPM 



CT3 



4 5 
ROTOR SPEED. Hz 



Figure 10. Instability region as a function of 
blade lag damping for the nonisotropic hub and 
one blade damper inoperative. 



158 



THEORY AND COMPARISON WITH TESTS OF TWO FULL-SCALE PROPROTORS 



Wayne Johnson 

Research Scientist 

Laige-Scale Aerodynamics Branch 

Ames Research Center, NASA 

and 

U.S. Army Air Mobility R&D Laboratory 

Moffett Field, California 



Abstract 



A nine degrees-of-freedom theoretical model has been devel- 
oped for investigations of the dynamics of a proprotor operating 
in high inflow axial flight on a cantilever wing. The theory is 
described, and the results of the analysis are presented for two 
proprotor configurations: a gimballed, stiff-inplane rotor, and a 
hingeless, soft-inplane rotor. The influence of various elements of 
the theory is discussed, including the modeling used for the blade 
and wing aerodynamics and the influence of the rotor lag degree 
of freedom. The results from full-scale tests of these two prop- 
rotors are presented and compared with the theoretical results. 

Notation 
en blade lift-curve slope 

p wing torsion degree of freedom 

q t wing vertical bending degree of freedom 

q 2 wing chordwise bending degree of freedom 

R rotor radius 

V forward velocity 

blade flap degree of freedom 

(3 rotor coning degree of freedom 

0! c rotor tip path plane pitch degree of freedom 

(3, s rotor tip path plane yaw degree of freedom 

7 blade Lock number 

8 3 pitch/flap coupling 

f damping ratio of eigenvalue, -ReX/ 1 X | 

f blade lag degree of freedom 

f rotor collective lag degree of freedom 

f time derivative of f ; rotor speed perturbation degree of 
freedom for autorotation case 

?! c rotor vertical cyclic lag degree of freedom 

?! s rotor lateral cyclic lag degree of freedom 

X eigenvalue or root 

Vo rotating natural frequency of blade flap motion 

eg- rotating natural frequency of blade lag motion 

to frequency of eigenvalue, ImX 

J2 rotor rotational speed 

The tilting proprotor aircraft is a promising concept for 
short haul, V/STOL missions. This aircraft uses low disk loading 
rotors located on the wing tips to provide lift and control in 

Presented at the AHS/NASA-Ames Specialists' Meeting on Rotor- 
craft Dynamics, February 13-1!>, 1974. 



hover and low speed flight; it also uses the same rotors to provide 
propulsive force in high speed cruise, the lift then being supplied 
by a conventional wing. Such operation requires a ninety degree 
change in the rotor thrust direction, which is accomplished by 
mechanically tilting the rotor shaft axis. Thus the aircraft com- 
bines the efficient VTOL capability of the helicopter with the 
efficient, high speed cruise capability of a turboprop aircraft. 
With the flexible blades of low disk loading rotors, the blade 
motion is as important an aspect of tilt rotor dynamics as it is for 
helicopters. When operated in cruise mode (axial flight at high 
forward speed), the rotor is operating at a high inflow ratio (the 
ratio of axial velocity to the rotor tip speed); such operation 
introduces aerodynamic phenomena not encountered with the 
helicopter rotor, which is characterized by low inflow. The 
combination of flapping rotors operating at a high inflow ratio 
on the tips of flexible wings leads to dynamic and aerodynamic 
characteristics that are in many ways unique to this configura- 
tion. The combination of efficient VTOL and high speed cruise 
capabilities is very attractive; it is therefore important to estab- 
lish a clear understanding of the behavior of this aircraft and 
adequate methods to predict it, to enable a confident design of 
the aircraft. Experimental and theoretical investigations have 
been conducted over several years to provide this capability. 
However, much remains to be studied, both in the fundamental 
behavior and in the more sophisticated areas such as the design 
and development of automatic control systems for the vehicle. 
This paper presents the results of a theoretical model for a 
proprotor on a cantilever wing, including application to two 
proprotor designs." a gimballed, stiff-inplane rotor and a hingeless, 
soft-inplane rotor. Using these two cases, the influence on the 
system dynamics of several elements of the analysis was exam- 
ined, including the effects of the rotor blade lag motion and the 
rotor and wing aerodynamic models. The predicted stability 
characteristics are then compared with the results of full-scale 
tests of these two proprotor designs. The development of this 
theory is presented in detail in Reference 1, together with some 
additional applications to the analysis of proprotor aeroelastic 
behavior. 

Analytical Model 

Figure 1 shows the proprotor configuration considered for 
the theory and for the full-scale tests. The rotor is operating in 
high inflow axial flight on a cantilever wing. For the tests, the 
rotor was unpowered, hence operating in autorotation (really in 
the windmill brake state). This configuration incorporates the 
features of greatest importance to the aircraft: the high inflow 
aerodynamics of a flapping rotor in axial flow and the coupled 
dynamics of the rotor/pylon/wing aeroelastic system. Many 
features of the aircraft-coupled wing and rotor motion may be 
studied with such a model, theoretically and experimentally, 
with the understanding that the model must eventually incorpo- 
rate the entire aircraft. 



159 



The theoretical model of the proprotor developed in Refer- 
ence 1 consists of nine degrees of freedom: the first mode flap 
(out of disk plane) and lag (inplane) motion for each of three 
blades; and vertical bending, chordwise bending, and torsion for 
the cantilever wing. The degrees of freedom of the individual 
rotor blades are combined into degrees of freedom representing 
the motion of the rotor as a whole in the nonrotating frame. 
Thus the rotor flap motion is represented by tip path plane pitch 
and yaw (fii c and /3 X s ) and coning (/3 ) degrees of freedom. The 
rotor lag motion is represented by cyclic lag ^ c and fi s (lateral 
and vertical shift of the rotor net center of gravity) and collective 
lag f Q . Wing vertical and chordwise bending of the elastic axis (qj 
and q 2 ) and torsion about the elastic axis (p) complete the set of 
nine degrees of freedom. 

The rotor blade motion is represented by first mode flap 
and lag motion, which are assumed to be respectively pure 
out-of-plane and pure inplane deflection of the blade spar. For 
the gimballed and hingeless rotor blades considered here (except 
for the flap mode of the gimballed rotor), there is, in fact, some 
elastic coupling of the flap and lag modes, so that there is 
actually participation of both out-of-plane and inplane motion in 
each mode. In the coefficients giving the aerodynamic forces on 
the rotor, it is further assumed that the mode shapes are propor- 
tional to the radial distance from the hub, i.e., equivalent to rigid 
body rotation about a central hinge. The model based on these 
two assumptions, which considerably simplify the aerodynamic 
and structural terms of the rotor equations, proves to be an 
adequate representation of the proprotor dynamics. 

The theoretical results presented here will usually be for the 
rotor operating unpowered, i.e., windmilling or autorotation 
operation. An important element of autorotation dynamic 
behavior is the rotor speed perturbation. With no restraint of the 
rotor shaft rotation, this degree of freedom has considerable 
influence on the aeroelastic behavior of the proprotor and wing. 
The rotor speed perturbation is modeled by using the collective 
lag mode f Q . By setting the rotating natural frequency of this 
mode to zero, i.e., zero spring restraint, f becomes equivalent to 
the rotor speed perturbation. (The natural frequency for the 
cyclic lag modes, f a c and f i s , is not set to zero in the representa- 
tion.) The other extreme case is that of the hub operating at 
constant angular velocity (fi) with no speed perturbation, with 
f then the elastic inplane deflection of the blade with respect to 
the hub. This case will be considered to represent powered 
operation of the rotor, although it is really the limit of operation 
with a perfect governor on rotor speed. 

The proprotor operating in high inflow has simpler aero- 
dynamics than the low inflow rotor in forward flight. As for the 
case of low inflow (i.e., the hovering helicopter rotor), the 
symmetry of axial flow results in a corresponding symmetry in 
the equations of motion; it also means that the differential 
equations of motion have constant coefficients. In high inflow 
there is the additional fact that both out-of-plane and inplane 
motions of the blade produce significant angle-of-attack changes 
at the sections, and the resulting lift increment has significant 
components both normal to and in the disk plane. Hence the 
rotor aerodynamic forces are primarily due to the lift changes 
produced by angle-of-attack changes, i.e., the co^ terms in the 
aerodynamic coefficients. This is in contrast to low inflow, 
where, for example, the inplane blade motion produces signi- 
ficant contributions to the forces by the lift and drag increments 



due to the dynamic pressure changes, i.e., the eg and c^ terms in 
the aerodynamic coefficients. As a result, high inflow rotor 
aerodynamics are well represented by considering only the cg„ 
forces. If, in addition, the lift curve slope is assumed constant, 
then the aerodynamic coefficients depend on only two param- 
eters, the Lock number y and the inflow ratio V/JZR. Refer- 
ence 1 presents the aerodynamic coefficients also for a more 
complete theoretical model of the rotor aerodynamics, namely a 
perturbation about the local trim state, including the cg^, eg, c,j, 

and Cj terms (and also derivatives with respect to the Mach 

"a 
number). Such a model is in fact little more difficult to derive 

than with the cg a terms alone. The influence of its use in the 

theory is examined below. 

This nine degrees-of-freedom model will have nine roots or 
eigenvalues (really nine pairs of complex roots) and correspond- 
ingly nine eigenvectors or modes. Of course, each mode involves 
motion of all nine degrees of freedom. The modes are identifiable 
by their frequencies (which will be near the uncoupled natural 
frequencies, nonrotating for the rotor modes), and also by the 
participation of the degrees of freedom in the eigenvector. The 
nine modes will be denoted as follows (the approximate 
uncoupled, nonrotating natural frequency of the mode is given in 
parentheses): 

p wing torsion (w p ) 

q x wing vertical bending (w q ) 

q 2 wing chordwise bending C« q ) 

coning (i^) 

P + 1 high frequency flap (va + fi) 

P - 1 low-frequency flap (vo - fi) 

f collective lag (fg-) 

i +. 1 high-frequency lag(i>>. + S2) 

f - 1 low-frequency lag (v> - Q.) 

The basic theoretical model will consist of all nine degrees 
of freedom, autorotation operation, and just the cg a rotor aero- 
dynamic forces. The wing aerodynamic forces are also included 
(based on a strip theory calculation). The dynamic stability of 
the system, specifically the frequency and damping ratio of the 
modes, will be examined for variations of the forward velocity V 
and the rotor rotational speed S2. Both V and fi sweeps change 
the inflow ratio V/J2R, and hence the rotor and wing aero- 
dynamic forces. A variation of the rotor speed £1 also changes the 
values of the wing and rotor blade nondimensional (per rev) 
natural frequencies. The rotor frequencies may also vary with 
Y/CIR due to the change in the rotor collective pitch angle. 
Several elements of the theoretical model will be examined to 
determine their influence on the predicted proprotor dynamics: 
the influence of the blade lag motion (by dropping the f , c and 
Ji s degrees of freedom), the wing aerodynamics (by dropping the 
wing aerodynamic coefficients), the rotor speed perturbation 
(i.e., the autorotation and powered cases), and the more com- 
plete model of the rotor aerodynamics (compared to just the eg 
terms). 

Two Full-Scale Proprotors 

The theory described above will be applied to two full-scale 
proprotors. The first is a 25-ft diameter gimballed, stiff-inplane 
proprotor, designed and constructed by the Bell Helicopter 
Company, and tested in the Ames 40- by 80-ft wind tunnel in 
July 1970. The second is a 26-ft diameter hingeless, soft-inplane 



160 



proprotor, designed and constructed by the Boeing Vertol Com- 
pany, and tested in the Ames 40- by 80-ft wind tunnel in August 
1972. The configuration for the dynamics tests consisted of the 
vrindmilling rotor operating in high inflow axial flow on the tip 
of a cantilever wing, as shown in figure 1 . As far as their dynamic 
characteristics are concerned, the two rotors differ primarily in 
the placement of the rotating natural frequencies of the blade 
flap and lag motions. The Bell rotor has a gimballed hub and 
stiff-inplane cantilever blade attachment to the hub, hence 
Vo= 1/rev (nearly, for it does have a weak hub spring) and 
vy> 1/rev; it also incorporates positive pitch/flap coupling 
(5 3 < 0) to increase the blade flap/lag stability. The Boeing rotor 
has a cantilever or hingeless hub with soft-inplane blade attach- 
ment, hence va > 1/rev and i>*. < 1/rev. The different placement 
of the blade frequencies, at the opposing extremes of the possible 
choices, results in quite different dynamic characteristics for the 
two aircraft. 

The rotors are described in References 2 to 5. Table I gives 
the major parameters of the rotor, and of the cantilever wing 
used in the full-scale tests (a more complete description of the 
parameters required by the theory is given in Reference 1). The 
wing frequencies in the theory were match to the experimentally 
measured values by adjusting the spring constants. The typical 
wing frequencies given in table I are for the coupled motion of 
the system (including the rotor) at 100 knots and design £2. The 
blade rotating natural frequencies are shown in figures 2 and 3 
for the Bell and Boeing rotors, respectively. The variation of the 
Bell lag frequency (fig. 2b) with V/fiR is due to the collective 
pitch change. The Boeing rotor blade frequencies vary little with 
collective pitch (V/J2R) since the blade has nearly isotropic 
stiffness at the root. 

The damping ratio of the wing modes was measured in the 
full-scale tests by the following technique. The wing motion was 
excited by oscillating an aerodynamic shaker vane on the wing 
tip (visible in fig. 1) at the wing natural frequency. After a 
sufficient amplitude was achieved, the vane was stopped. Then 
the frequency and damping ratio were determined from the 
decay of the subsequent transient motion of the wing. 

Results and Discussion 

Gimballed, Stiff-inplane Rotor 

The effects of several elements of the theoretical model will 
be examined for a gimballed, stiff-inplane rotor (the Bell rotor). 
The theoretical results will then be compared with the results of 
full-scale tests. The test results and results from the Bell theories 
are from Reference 2. The predicted variation of the system 
stability with forward speed V at the normal airplane mode rotor 
speed (£2 = 458 rpm) is shown in figure 4, in terms of the fre- 
quency and damping ratio of the eigenvalues. The wing vertical 
bending mode (q x ) becomes unstable at 495 knots. The damping 
of that mode first increases with speed; the peak is due to 
coupling between the wing vertical bending (qi) and low- 
frequency rotor lag (f - 1) modes (it occurs at the resonance of 
the frequencies of these two modes). Figure 5 shows the influ- 
ence of the rotor lag motion, comparing the damping of the wing 
modes with and without the fj c and Si s degrees of freedom. The 
rotor low-frequency lag mode has an important influence on the 
motion, particularly on the wing vertical bending mode; the q[ 
damping is increased when its frequency is below that of the 



?- 1 mode (low V), and decreased when its frequency is above 
that value. The high speed instability is relatively unaffected, 
however, indicating that the mechanism of that instability 
involves primarily the rotor flap motion. Therefore, the net 
effect of the reduced damping at high speed due to the lag 
motion is a reduction of the rate at which the damping decreases, 
which is beneficial since the instability is then less severe. 

Figure 6 shows the influence of powered operation (stabiliz- 
ing) and of omitting the wing aerodynamics (destabilizing). The 
powered state effect is the influence of dropping the rotor speed 
perturbation degree of freedom. The wing aerodynamics effect is 
mainly the loss of the aerodynamic damping of the wing modes 
due to the angle-of-attack changes during the motion. Figure 7 
shows the influence of the more complete theoretical model for 
the blade aerodynamics, compared with the results using only the 
eg terms. The basic behavior remains the same, but the better 
aerodynamic model reduces slightly the level of the predicted 
damping ratio. The predicted speed at the stability boundary is 
significantly reduced, however, because of the gradual variation 
of the damping with speed. It is therefore concluded that for the 
prediction of the characteristics of an actual aircraft, the best 
model available for the rotor aerodynamics should be used. 

Figure 8 shows the variation of the dynamic stability with 
velocity at the normal rotor speed (J2 = 458 rpm), in terms of 
the frequency and damping ratio of the wing modes; the full- 
scale test results for the Bell rotor are compared with the pre- 
dicted stability. Also shown are predictions from the Bell linear 
and nonlinear theories, from Reference 2. Figure 8 shows reason- 
able correlation between the predicted and full-scale test stability 
results. Additional comparisons with the full-scale test data are 
given in Reference 1 . 

Figure 9 shows the influence of the rotor lag motion. The 
predicted and measured stability is shown for the Bell rotor on 
the full-stiffness wing, and on a quarter-stiffness wing. (By oper- 
ating on a quarter-stiffness wing at one-half design rotor speed, 
n = 229 rpm; the wing frequencies and inflow ratio are modeled 
for an equivalent speed twice the actual tunnel speed.) Also 
shown is the predicted stability for the rotor on the full-stiffness 
wing, but without the f j c and f i s degrees of freedom. Figure 9a 
shows the variation of the wing vertical bending mode damping. 
The full-scale experimental data show a definite trend to higher 
damping levels with the full-stiffness wing, and this trend corre- 
lates well with the results of the present theory. Figures 9b and 
9c show the predicted stability of all the wing modes. The 
difference in damping at the same inflow ratio is due to the rotor 
lag motion. Figure 9d shows the frequencies of the f - 1 , q, , and 
p modes for the full-stiffness and quarter-stiffness wings. The 
full-stiffness wing has a resonance of the f - 1 and q, modes 
which produces the peak in the damping. Slowing the rotor on 
the quarter-stiffness wing greatly increases the lag frequency (per 
rev), removing it from resonance with the qj mode. Another way 
to remove the influence of the rotor lag motion, in the theory, is 
simply to drop the f , c and fj s degrees of freedom from the full- 
stiffness wing case. When these degrees of freedom are dropped, 
the predicted wing vertical bending damping is almost identical 
to that for the quarter-stiffness wing (figs. 9a and 9b). Figure 10 
examines further the influence of the rotor lag motion on the 
wing vertical bending mode damping. Predicted stability with and 
without the ?i c and J k s degrees of freedom is compared with 
experimental results from tests of a 0.1333-scale model of a 



161 



gimballed, stiff-inplane proprotor. The test results are from Ref- 
erence 6; this rotor is a model of the Bell M266, similar in design 
to the full-scale rotor considered here. The experimental data 
correlates well with the predictions, including the influence of 
the rotor lag motion. 

Hingeless, Soft-Inplane Rotor 

The effects of several elements of the theoretical model will 
be examined for a hingeless, soft-inplane rotor (the Boeing 
rotor). Then the theoretical results will be compared with the 
results of full-scale tests, and with results from the Boeing theory 
(the latter are from Reference 3). The predicted variation of the 
system stability with forward velocity at normal rotor speed 
(Q, = 386 rpm) is shown in figure 11. The low-frequency flap 
(/3-1) mode becomes unstable at 480 knots. By the time this 
instability occurs, the mode has assumed the character of a wing 
vertical bending mode (i.e., the q x motion, and the associated p, 
fie. fis. and j motions); hence this instability has the same 
mechanism as does the Bell rotor. With the soft-inplane rotor, 
v* < 1/rev, the proximity of the f - 1 and qi mode frequencies 
significantly reduces the wing mode damping at low speeds; this 
effect is the air resonance phenomenon. A similar influence 
occurs with the resonance of the J - 1 and q 2 modes, leading to 
an instability of the, wing chord mode (this instability can occur 
because the wing chord mode aerodynamic damping remains low 
even at high speed). At higher £2, this q 2 mode instability is, in 
fact, the critical instability. The influence of the rotor lag motion 
is shown in figure 12. The substantial decrease in the damping of 
the wing vertical and chordwise bending modes due to the rotor 
lag motion is the air resonance effect. Figure 13 shows the 
influence of powered operation and of omitting the wing aero- 
dynamics, and figure 14 shows the influence of the better theo- 
retical model for the rotor aerodynamics on the predicted 
stability. The effects, and hence the conclusions from figures 13 
and 14 are similar to those for the Bell rotor. 

Figure 15 shows the variation of the predicted stability of 
the Boeing rotor with rotor speed at 50 knots. At this low speed, 
the resonance of the £ - 1 and qj mode frequencies actually 
results in an instability of the wing vertical bending mode. 
Figure 16 shows the stability variation with rotor speed at 
192 knots. The reduction in wing vertical bending mode damping 
due to air resonance is still present, but the increase in the rotor 
lag aerodynamic damping and wing vertical bending aerodynamic 
damping with flight speed has been sufficient to stabilize the 
motion even at resonance. Figure 17 summarizes the air 
resonance behavior of the Boeing rojor. 

Figure 18 compares the predicted and full-scale results for 
the stability of the wing modes for a velocity sweep of the 
Boeing rotor at £2 = 386 rpm. Figure 19 shows the variation of 
the wing vertical bending mode damping with rotor speed at 
V = 50 to 192 knots. These runs were conducted to investigate 
the air resonance behavior of this configuration, i.e., the influ- 
ence of the rotor lag motion. Reasonable correlation is shown 
between the predicted and measured stability, except at the 
higher speeds where tunnel turbulence made extraction of the 
damping ratio from the experimental transient wing motion 
difficult. Also shown are predictions from the Boeing theory, 
from Reference 3. Additional comparisons with the full-scale test 
data are given in Reference 1. 



Concluding Remarks 

This paper has presented theoretical results for the stability 
of a proprotor operating in high inflow on a cantilever wing : 
Some experimental results from full-scale tests have been pre- 
sented, showing reasonable correlation with the predicted stabil- 
ity. The nine degrees-of-freedom theoretical model has been 
established as a useful and accurate representation of the basic 
dynamic characteristics of the proprotor and cantilever wing 
system. The significant influence of the rotor speed perturbation 
degree of freedom (i.e., windmilling or powered operation), the 
wing aerodynamics, and the rotor aerodynamic model on the 
predicted stability have been shown, indicating the importance of 
including these elements accurately in the theoretical model. 
From a comparison of the behavior of the gimballed, stiff-inplane 
rotor and the hingeless, soft-inplane rotor, it is concluded that 
the placement of the natural frequencies of the rotor blade first 
mode bending — i.e., the flap frequency va and the lag frequency 
vy — has a great influence on the dynamics of the proprotor and 
wing. Moreover, the theoretical and experimental results have 
demonstrated that the rotor lag degree of freedom has a very 
important role in the proprotor dynamics, for both the soft- 
inplane (y* < 1/rev) and stiff-inplane (v* > 1/rev) configurations. 

References 



1. NASA TN-D (in preparation), THE DYNAMICS OF TILT- 
ING PROPROTOR AIRCRAFT IN CRUISE FLIGHT, 
Johnson, Wayne, 1974. 

2. NASA CR 114363, ADVANCEMENT OF PROPROTOR 
TECHNOLOGY TASK II — WIND TUNNEL TEST 
RESULTS, Bell Helicopter Company, September 1971. 

3. Boeing Vertol Company Report No. D222-1 0059-1, WIND 
TUNNEL TESTS OF A FULL SCALE HINGELESS 
PROP-ROTOR DESIGNED FOR THE BOEING 
MODEL 222 TILT ROTOR AIRCRAFT, Magee, John P., 
and Alexander, H. R., July 1973. 

4. NASA CR 114442, V/STOL TILT-ROTOR STUDY 
TASK II - RESEARCH AIRCRAFT DESIGN, Bell Heli- 
copter Company, March 1972. 

5. NASA CR 114438, V/STOL TILT-ROTOR AIRCRAFT 
STUDY VOLUME II - PRELIMINARY DESIGN OF 
RESEARCH AIRCRAFT, Boeing Vertol Company, March 
1972. 

6. Kvaternik, Raymond G., STUDIES IN TILT-ROTOR 
VTOL AIRCRAFT AEROELASTICITY, Ph.D. Thesis, Case 
Western Reserve University, June 1973. 



162 



TABLE I - DESCRIPTION OF THE FULL-SCALE 
PROPROTORS, AS TESTED IN THE AMES 40- BY 80-FT 
WIND TUNNEL. 





Bell 


Boeing 


Rotor 






Type 


gimballed, stiff- 


hingeless, soft- 




inplane 


inplane 


Number of blades 


3 


3 


'Radius, R 


3.81m (12.5 ft) 


3.96 m (13 ft) 


Lock number, y 


3.83 


4.04 


Solidity ratio 


0.089 


0.115 


Pitch/flap coupling, 
5, 


-15deg 





Rotor rotation direc- 


clockwise 


counterclockwise 


tion, on right wing 






Tip speed, SIR 


183 m/sec 


160 m/sec 


(cruise mode) 


(600 ft/sec) 


(525 ft/sec) 


Rotation speed, 


458 rpm 


386 rpm 


(cruise mode) 






Wing 






Semispan, y w /R 


1.333 


1.281 


Mast height, h/R 


0.342 


0.354 


Typical frequencies 






Vertical bending 


3.2 Hz 0.42/rev 


2.3 Hz 0.36/rev 


Chordwise bending 


5.35 0.70 


4.0 0.62 


Torsion 


J 9.95 1.30 


9.2 1.48 




<D g - 





\ 
\ 


\ 




~ 




\ 
\ 
^ 










a, rpm 


- 


\ 


^"^^, 


_^-229 




\ 


^' — .^ 




(b) 


I 


I 1 I I 


.-350 
J^458 
">550, 



1 

v/a rpm 

Figure 2. Blade rotating natural frequencies for the Bell rotor, 
(a) flap frequency vq (normal SI = 458 rpm), (b) lag 
frequency v> (V/S2R = 1 at 355 knots and normal SI). 




fl, rpm 
,l 93 



•300 ,386 



(a) 



^550 




V/flR 



Figure 1. Configuration of analytical model, and for full-scale 
tests: proprotor operating in high inflow axial flight on a 
cantilever wing. 



Figure 3. Blade rotating natural frequencies for the Boeing rotor 
(V/S2R =1 at 311 knots and normal St, 386 rpm). (a) flap 
frequency Pa, (b) lag frequency v* . 



163 





200 400 

V, knots 




600 



V, knots 



Figure 5. Effect of deleting the rotor lag degrees of freedom (f i c 
and f is ), Bell rotor velocity sweep at fi = 458rpm. 

(a) damping of wing vertical bending mode (q t ), 

(b) damping of chordwise bending (q 2 ) and torsion (p) 
modes. 




VELOCITY SWEEP 

»H 1 — I 1 — I — h— 

25 200 400 600 

V, knots 




2J0 



1.0 



Rex 



AUTOROTATION, WITH WING AERODYNAMICS 

POWERED 

MO WING AERODYNAMICS 



-.05 FOR q,, q 2 , p 




600 



Figure 4. Predicted stability of Bell rotor, velocity sweep at Figure 6. Influence of powered operation, and wing aerodynamic 
12 = 458 rpm. (a) frequency of the modes, (b) damping forces, Bell rotor velocity sweep at Q. = 458 rpm. 

ratio of the modes, (c) root locus. (a) damping of wing vertical bending mode (qj), 

(b) damping of chordwise bending (q 2 ) and torsion (p) 

modes. 



164 



.05 



MORE CQHPLETE MODEL FOR ROTOR AEROOTIIAitlCS 

OKLY C La TERMS Id ROTOR AERODtHAHIC C0EFF1GIEHTS 

H H- HELICAL TIP MACH NUMBER LIMITS 

CRITICAL SONIC 



POKERED 




.15 


- 


s~/ \ 










^^^AIITOROTATION \ 






POWERED // / KT~ V \\ 


.10 








yv / \*s\\- 






/ / / ' \\ 






A // \ 






/' / / T\ 






/' /"' w 


.05 




~j^^^=r^-_l'0*ERED \\ 

AUT0R0TATI0in\^ N^-\A "^ 




(b, , 


i i i \ \i \ r^-^ i 



200 400 

V, knots 



600 



Figure 7. Influence of a more complete model for the rotor 
aerodynamics, Bell rotor velocity sweep at SI = 458 rpm. 

(a) damping of wing vertical bending mode (q,), 

(b) damping of chordwise bending (qj) and torsion (p) 
modes. 



a 



(a! 



.05 



(b) 



j05 



(c) 



.05 - 



td) 



PRESENT THEORY 

NONLINEAR 



LINEAR 

o EXPERIMENT 



BELL THEORY 






" ^ HJ Ltj ' t! ■wWaaBua ^Basr- sf i " "rr i n* 



_ g_ 





100 



200 



V, knots 



Figure 8. Comparison with full-scale experimental data, Bell 
rotor velocity sweep at S2 = 458 rpm. (a) frequency of the 
modes, (b) damping of the wing vertical bending mode (qj ), 

(c) damping of wing chordwise bending mode (q 2 ), 

(d) damping of wing torsion mode (p). 



.05 




(a) 



i i i i ii i i i i 



.5 
V/flR 



1.0 




Figure 9. Influence of the rotor lag motion. Bell rotor velocity sweeps on the full-stiffness wing, on a quarter-stiffness wing, and on 
the full-stiffness wing without the $Vc and ?is degrees of freedom (theory only), (a) damping of the wing vertical bending 
mode (q, ), comparison with full-scale experimental data, (b) damping of wing vertical bending mode (q, ), (c) damping of wing 
chordwise bending (q 2 ) and torsion (p) modes, (d) frequency of the modes. 



165 



.15 



.10 



.05 



EXPERIMENT THEORY 

o FULL STIFFNESS WING 

• QUARTER STIFFNESS WINS 

FULL STIFFNESS WING, WITHOUT £ |C , £, s 




a 





-^ 




/V 




-^-_. 






- 








A 




^- 


(d) 


i 


A- 


i 

i 



1.0 



V/flR 



Figure 9. Concluded. 





o EXPERIMENT (0.1333 SCALE MODEL) 


.08 


WITHOUT £ |C , £ |S 


/\ 






/ \ 






/ \ 




°^^-i 


/ 




y/ 


> \ \ 




77 






a/ 


o\ \ 


.04 


// 


o\ 1 

\ ° 1 




yS 


\ ° 1 




^^^ 








V 




(a) 238 rpm 







1 1 


1 A 




200 x N 

V, knots 



400 



3 






+ 1 




? 






£ + 1 






P 









1 


_ 




A 


T~~ 






Qi\ 


C-K-— 


-— ^-1 




(ah- 


1 


1 1 1 


,A , 




-Rex 



-.05 FOR q |r q 2 



Figure 10. Comparison with experimental data from tests of a 
0.1333-scale rotor and cantilever wing model of Bell M266 

aircraft (experimental points from Reference 6), velocity Figure 11. Predicted stability of Boeing rotor, velocity sweep at 
sweeps at (a) 12 = 238 rpm, (b) £2 = 298 rpm, $2 = 386 rpm. (a) frequency of the modes, (b) damping 

(c) £2 = 358 rpm (equivalent full-scale V and J2). ratio of the modes, (c) root locus. 

166 



WITHOUT J |C , C (S 



WITH £.., { 




.10 



.06 



AUTOROTAT10H , 

WITH WINS AERODYNAKICS 

POWERED 

HO WING AERODYNAMICS 



.15 



.10 



.05 - 




600 



200 400 

V, knots 



600 



Figure 1 2. Effect of deleting the rotor lag degree of freedom (f t c 
and ?! s ), Boeing rotor velocity sweep at to = 386 rpm 
(a) damping of wing vertical bending (q t ) and flap ((3-1) 
modes (the (3 - 1 mode is shifted by 250-300 knots to 
higher speed by the removal of the lag influence, beyond 
the scale shown), (b) damping of wing chordwise bending 
(q 2 ) and torsion (p) modes. 



Figure 13. Influence of powered operation, and wing 
aerodynamic forces, Boeing rotor velocity sweep at 
to = 386 rpm. (a) damping of wing vertical bending (q a ) 
and rotor flap 03-1) modes, (b) damping of wing 
chordwise bending (q 2 ) and torsion (p) modes. 





600 



v, knots 



Figure 14. Influence of a more complete model for the rotor aerodynamics. Boeing rotor velocity sweep at to = 386 rpm (a) damp- 
ing of wing vertical bending (q, ) and rotor flap 03—1) modes, (b) damping of wing chordwise bending (q t ) and torsion (p) 
modes. - 

167 





.15 r 




600 




600 



Figure 15. Predicted stability of Boeing rotor, rpm sweep at Figure 16. Predicted stability of Boeing rotor, rpm sweep at 
50 knots, (a) frequency of the modes, (b) damping ratio of 192 knots, (a) frequency of the modes, (b) damping ratio of 

the modes. the modes. 




400 

a, rpm 



600 



Figure 1 7. Air resonance behavior of soft-inplane hingeless rotor. Boeing rotor at 50 to 192 knots, variation of damping of 

wing vertical bending mode (q, ) with rotor speed. 



168 



•PRESENT THEORY 



2 r 



JUL 
a 



BOEING THEORY 

O EXPERIMENT 



-g— o- 



"2 



(0) 4 I 




400 



Figure 18. Comparison with full-scale experimental data, Boeing rotor velocity sweep at SI = 386 rpm. (a) frequency of the modes, 
(b) damping of the wing vertical bending (qi ), chord bending (q 2 ), and torsion (p) modes; the experimental data is for q! only. 



• 60 knots 



EXPERftiEHT 



o 50 knots 

.03, 80 J* fto t S ) PRESEKT THEORY p 

50 knots I 

BOEING THEORV, 60 knots 








On o 


/T 


o \»P 

(b) IOO knots ^X 
1 1 ^ii- 


/ / 
// 

f i 



PRESENT THEORY 

BOEING THEORY 

o EXPERIMENT 



J03 



.02 - \> 



" s -^. ,o © 


1 


\>oo 
0H50 


°° // 


\P[ 


f 1 




► 1 


(c! 140 knots o 


" \ 1 



300 



400 500 

a, rpm 



\ G oo o 

-\v °^° / 

(d) 192 knots 



600 300 



400 500 

a, rpm 



600 



Figure 1 9. Comparison with full-scale experimental data, Boeing rotor rpm sweeps, damping of wing vertical bending mode at 

(a) 50-60 knots, (b) 1 00 knots, (c) 140 knots, (d) 192 knots. 



169 



EXPERIMENTAL AND ANALYTICAL STUDIES IN TILT- ROTOR AEROELASTICITY ' 

Raymond G. Kvaternik 

Aerospace Technologist 

NASA Langley Research Center 

Hampton, Virginia 



Abstract 

An overview of an experimental and analytical 
research program underway within the Aeroelasticity 
Branch of the NASA Langley Research Center for 
Studying the aeroelastic and dynamic characteris- 
tics of tilt-rotor VTOL aircraft is presented. 
Selected results from several joint NASA/contractor 
investigations of scaled models in the Langley 
transonic dynamics tunnel as well as some results 
from a test of a flight-worthy proprotor in the 
NASA Ames full-scale wind tunnel are shown and dis- 
cussed with a view toward delineating various 
aspects of dynamic behavior peculiar to proprotor 
aircraft. Included are such items as proprotor/ 
pylon stability, whirl flutter, gust response, and 
blade flapping. Theoretical predictions, based on 
analyses developed at Langley, are shown to be in 
agreement with the measured stability and response 
behavior. 

Notation 

e Blade flapping hinge offset 

H Rotor normal shear force 

cfflt/oNx Rotor normal shear force component in 
phase with pitch angle 

5H/5q, Rotor normal shear force component in 
phase with pitch rate 



R 


Blade radius 


R 


0.75 blade radius 


AT 


Rotor perturbation thrust 


V 


Airspeed 


v F /nR 


Flutter advance ratio 


w 

g 


Vertical component of gust velocity 


a 

m 


Mast angle of attack 



a Oscillation amplitude of airstream 
oscillator 

B Blade flapping angle 1 

3(3 /da Blade flapping derivative 

8, Pitch-flap coupling angle 



Presented at the AHS/NASA Ames Specialists' Meeting 
on Rotorcraft Dynamics, February 13-15, 197^- 



e Gust- induced angle of attack 
g 

£_ Hub damping ratio 

+ Aircraft yaw rate 

SI Rotor rotational speed 

Q Frequency 

oo Blade flapping natural frequency 

u> Pylon pitch frequency 

co. Pylon yaw frequency 

The feasibility of the tilt-proprotor com- 
posite aircraft concept was established in the mid 
1950' s on the basis of the successful flight 
demonstrations of the Bell XV-3 and Transcendental 
Model 1-G and Model 2 convert iplanes. Flight 
research conducted with the XV-3 identified 
several dynamic deficiencies in the airplane mode 
as technical problems requiring further atten- 
tion.-'- A more serious proprotor dynamic problem 
was identified in a 1962 wind-tunnel test of the 
XV-3- In that test, conducted in the Ames full- 
scale tunnel, a proprotor /pylon instability simi- 
lar in nature to propeller whirl flutter was 
encountered. Clearly, to maintain the viability 
of the tilt-proprotor concept it remained to 
demonstrate that neither the whirl flutter anomaly 
nor the major flight deficiencies were endemic to 
the design principle. An analytical and experi- 
mental research program having this objective was 
undertaken by Bell in 1962. Results of this 
research, which defined the instability mechanism 
and established several basic design solutions, 
were reported by Hall. 2 Edenborough? presented 
results of subsequent full-scale tests at Ames in 
1966 which verified the analytical prediction tech- 
niques, the proposed design solutions, and demon- 
strated stability of the XV-3 through the maximum 
wind-tunnel speed of 100 m/s (195 kts). 

In 1965 the U.S. Army inaugurated the Com- 
posite Aircraft Program which had the goal of 
producing a rotary- wing research aircraft combin- 
ing the hovering capability of the helicopter with 
the high-speed cruise efficiency and range of a 
fixed-wing aircraft. Bell Helicopter Company, 
with a tilt-proprotor design proposal, was awarded 
one of two exploratory definition contracts in 
1967 . The Model 266 was the design resulting from 
their work (Fig. 1). The research aircraft pro- 
gram which was to have been initiated subsequent 
to the exploratory definition phase was never 
begun, however, primarily due to lack of funding. 



171 




Figure 1. Artist's conception of Bell Model 266 
tilt-proprotor design evolved during the Army 
Composite Aircraft Program. 

Concurrent with the developments described 
above, various VTOL concepts based on the use of 
propellers having independently hinged blades were 
proposed with several reaching flight-test status. 
These included the Grumman proposal in the Tri- 
Service VTOL Transport competition, the Vertol VZ-2 
built for the Army, and the Kaman K-l6 amphibian 
built for the Navy. A vigorous investigation of the 
whirl flutter phenomenon peculiar to conventional 
propellers had been initiated in i960 as a result 
of the loss of two Lockheed Electra aircraft in 
fatal accidents. The possibility that hinged 
blades could adversely affect the whirl flutter 
behavior of a propeller undoubtedly contributed 
considerable impetus to examine the whirl flutter 
characteristics of these flapping propellers. Wort 
related to these efforts was reviewed by BeedA 

The foregoing constitutes a resume of 
proprotor-ralated experience through 1967. This 
paper will present an overview of a research pro- 
gram initiated within the Aeroelasticity Branch of 
the NASA Langley Research Center.- Included in this 
program are joint NASA/contractor wind-tunnel 
investigations of scaled models in the transonic 
dynamics tunnel and the in-house development of 
supporting analyses. For completeness, motivating 
factors leading to the work and the scope of the 
investigation are outlined below. 

A 0.133- scale semispan dynamic and aeroelastic 
model of the Model 266 tilt rotor built by Bell in 
support of their work pertaining to the Composite 
Aircraft Program was given to Langley by the Army 
in 1968. The availability of this model and the 
interest of both government and industry in the 
tilt- rotor VTOL aircraft concept suggested the use- 
fulness of continuing the experimental work ini- 
tiated by Bell with the model to further define the 
aeroelastic characteristics of proprotor-type air- 
craft. Because both the XV-3 experience and studies 
conducted during the Composite Aircraft Program 
identified certain high-risk areas associated with 
operation in the airplane mode of flight, 



specifically proprotor /pylon stability (whirl 
flutter), blade flapping, and flight mode stability, 
it was judged that the research effort would be 
primarily directed to these areas. 



The experimental portion of the research pro- 
gram was initiated in September 1968 in a joint 
NASA/Bell study of proprotor stability, dynamics, 
and loads employing the 0. 133-scale semispan model 
of the Model 266, Several other cooperative 
experimental studies followed this investigation. 
The models employed in these studies are positioned 
in chronological order in the composite photo giverj 
in Figure 2. Briefly, these other studies include^: 
(l) A study of a folding proprotor version of the/ 
tilt-rotor model used in the first study, (2) a 
parametric investigation of proprotor whirl flutter, 
(3) a stability and control investigation employing 
an aerodynamic model, and (k) a "free- flight" 
investigation of a complete tilt-rotor model. 

TILT-ROTOR AEROELASTIC RESEARCH 

LANGLEY TRANSONIC DYNAMICS TUNNEL 




Figure 2. Tilt-rotor models tested in the Langley 
transonic dynamics tunnel. 

The results pertaining to the above-mentioned 
studies are quite extensive. The particular results 
to be presented herein have been selected with a 
view toward highlighting some of the dynamic 
aspects of proprotor behavior, delineating the 
effects of various design parameters on proprotor/ 
pylon stability and response, and providing valida- 
tion of analyses developed at Langley. The results 
pertaining to investigations conducted in the 
Langley transonic dynamics tunnel are presented 
first. These are arranged in chronological order 
according to Figure 2. To provide additional data 
for correlation, some experimental results obtained 
by Bell in tests of a semispan model and a full- 
scale flight-worthy proprotor are also included. 
In each case both experimental and analytical 
results are for the pylon fully converted forward 
into the airplane mode of operation and the rotors 
in a windmilling condition. Equivalent full-scale 
values are given unless noted otherwise. 



172 



Model Tests in Langley Transonic Dynamics Tunnel 

Bell Model 266 

(a) September I968 

Although the 0.133- scale semispan model of the 
Bell Model 266 was not designed to permit extensive 
parametric variations, in that it represented a 
'specific design, it did permit a fairly diversified 
|test program. The principal findings of this inves- 
tigation have been published and are available in 
the literature. 5, 6 seme results adapted from 
Reference 6 pertaining to stability and gust 
response are discussed below. 

Proprotor /Pylon Stability. To provide an - 
indication of the relative degree to which stabil- 
ity could be affected, and to provide a wide range 
of configurations for correlation with analysis, 
several system parameters were varied either indi- 
vidually or in combination with other parameters 
and the level of stability established. 

A baseline stability boundary, based on a 
reference configuration, was first established. 
The degree to which stability could be affected was 
then ascertained by varying selected system param- 
eters (or flight conditions). Stability data were 
obtained by holding rpm constant as tunnel speed 
was incrementally increased, transiently exciting 
the model by means of lightweight cables attached 
to the model, and analyzing the resulting time 
histories to determine the damping. The reference 
configuration consisted of the basic Model 266 
parameters with the pylon yaw degree of freedom 
locked out and the wing aerodynamic fairings 
removed. A 100$ fuel weight distribution was 
maintained by appropriately distributing lead 
weights along the wing spar. The hub flapping 
restraint was set to zero and the S3 angle to 
-0.393 radian (-22.5°). The reference stability 
boundary as well as changes in this boundary due to 
several parameter variations are shown in Figure 3- 

For the reference configuration instability 
occurred in the coupled pylon/wing mode in which 
the pylon pitching angular displacement is in phase 
with the wing vertical bending displacement. A 
characteristic feature of this coupled mode is the 
predominance of wing bending (relative to pylon 
pitch) and the frequency of oscillation, which is 
near the fundamental wing vertical bending natural 
frequency. For descriptive purposes this flutter 
mode is termed the "wing beam" mode herein. Negli- 
gible wing chordwise bending or rotor flapping 
(relative to space) was observed. The pylon/rotor 
combination also exhibited a forward whirl preces- 
sional motion, the hub tracing out an elliptical 
path in space. However, because of the large ratio 
of pylon yaw to pylon pitch stiffness the pylon 
angular displacement was primarily in the pitch 
direction. The flutter mode of the model in each 
of its perturbations from the reference configura- 
tion was essentially the same as for the reference 
configuration. 

The proprotor /pylon instability described 
above is similar in nature to classical propeller 



RPM 

4001- 



* 

I 

ft 




1 Flight condition 
for gust response 



Measured Calculated 

q _^_ — i — ,_ Reference boundary 

A *-' Altitude, 3668 m (12000 ft) 

O __-.__-— Pylon yaw unlocked (fy aw » 6 Hz) 

O -- — :.., Hub restraint 

□ — ^ — Wing aerodynamics 



J_ 



300 400 

Airspeed, knots 

I 



Figure 3- 



Airspeed, meters/sec 



Effect of several system parameters on 
proprotor/pylon stability. 



whirl flutter. However, because of the additional 
flapping degrees of freedom of the proprotor the 
manner in which the precession generated aerody- 
namic forces act on the pylon is significantly 
different." Specifically, while aerodynamic cross- 
stiffness moments are the cause of propeller whirl 
flutter, the basic destabilizing factors on 
proprotor/pylon motion are aerodynamic in plane 
shear forces which are phased with the pylon motion 
such that they tend to increase its pitching or 
yawing velocity and, hence, constitute negative 
damping on the pylon motions. 

(1) Altitude - Altitude has a highly benefi- 
cial effect on proprotor/pylon stability. This 
increased stability is a consequence of the fact 
that the destabilizing rotor normal shear forces 
decrease with altitude for pylon pitch frequencies 
near the fundamental wing elastic mode frequencies. 
This means that a given level of these destabiliz- 
ing shear forces is attained at progressively 
higher airspeeds as altitude increases. 

(2) Hub Flapping Restraint - A stabilizing 
effect due to moderate flapping restraint is also 
indicated in Figure 3- Increasing the flapping 
restraint increased the flapping natural frequency 
from its nominal value of about 0.8o/rev bringing 
it closer to the "optimum" flapping frequency in 
the sense of Young and Iytwyn.f They showed that 
this increased stability because the pylon support 



173 



stiffness requirements were reduced as the optimum 
flapping frequency was approached. 

(3) Wing Aerodynamics — Figure 3 indicates 
that wing aerodynamic forces have a slight stabi- 
lizing effect. Now the stiffness of a strength- 
designed wing for tilt-rotor application is 
generally sufficiently high to relegate the flutter 
speed of the pylon/wing combination (with blades 
replaced by lumped concentrated weights) to speeds 
well beyond the proprotor mode flight envelope. 
This suggests that wing aerodynamics will contrib- 
ute primarily to the damping of any coupled rotor/ 
pylon motions. This is substantiated in Figure It-, 
which shows the variation of the wing beam mode 
damping with airspeed through the flutter point for 
the reference configuration and the corresponding 
configuration with the wing airfoil segments 
installed. The damping of the mode is increased; 
however, the magnitude of the increase is small 
indicating that proprotor aerodynamic forces are 
predominant in the ultimate balance of forces at 
flutter. This provides some justification for 
neglecting, in this flutter mode at least, wing 
aerodynamics as a first approximation. 



a - 5 Hz (298 EPM) 
Measured Calculated 



o 

A 



Without wing aerodynamics 
With wing aerodynamics 



■Blade inplane flexibility included 



V 




50 



200 300 

Airspeed, knots 

_1 l_ 



100 .150 

Airspeed, meters/sec 



Figure h. Comparison of measured and calculated 
wing beam mode damping for reference 
configuration. 

The initial increase in the stability of the 
wing beam mode before instability occurs is asso- 
ciated with the fact that dH/dq, the component of 
the normal shear force associated with pylon pitch 
rate, initially becomes more stabilizing with 
increasing airspeed until 5H/ck%, the component of 
the normal shear force in phase with pylon pitch 
angle, becomes sufficiently large to lower the 
coupled pylon pitch frequency to a level where 
cffl/dq becomes increasingly destabilizing with 
increasing airspeed. " The increased damping 
response at about 103 m/s (200 kts) is due to 
coupling of the blade first inplane cyclic mode 
with the wing beam mode. Uote, however, that the 



predicted flutter speed is not sensitive to blade 
inplane flexibility for the Model 266. 

{k) Pylon Restraint — When the pylon yaw 
stiffness was reduced by unlocking the pylon yaw 
degree of freedom and soft-mounting the pylon in 
yaw relative to the wing tip the stability 
decreased slightly (Fig. 3). The particular yaw :' 
flexibility employed in this variation effectively f 
produced a more nearly isotropic arrangement of the f 
pylon support spring rates. Since the region of !f 
instability in a plot of critical pylon yaw stiff- 
ness against critical pitch stiffness is extended 
along the line representing a stiffness ratio of 
unity, the configuration approaching isotropy in 
the pylon supports is more prone to experience an 
instability than one in which one of the stiff- 
nesses is significantly less than the other. 

The general trend of decreasing stability with 
increasing rotor speed shown in Figure 3 was found f 
for all values of the adjustable parameters of the* 
model. In each case the, predicted flutter mode 
and frequency were in agreement with the correspond- 
ing measured mode and frequency. 

Gust Response . Analytical methods for deter- 
mining aircraft response to turbulence are usually 
based on power spectral analysis techniques which 
require the definition of the aircraft frequency 
response function, that is, the response to sinu- 
soidal gust excitation. A study to assess the 
feasibility of determining these frequency response 
functions for fixed-wing aircraft utilizing models 
in a semi-free-flight condition using a unique air- 
stream oscillator system in the transonic dynamics 
tunnel has been underway within the Aeroelasticity 
Branch for several years. ° This system (Fig. 5) 
consists of two sets of biplane vanes located on 
the sidewalls of the tunnel entrance section. The 




Figure 5. Langley transonic dynamics tunnel air- 
stream oscillator showing cutaway of driving 
mechanism. 

vanes can be oscillated in phase or l80 c out of 
phase to produce nominally sinusoidal vertical or 
rolling gusts, respectively, over the central por- 
tion of the tunnel. The gusts are generated by the 
cross- stream flow components induced by the trail- 
ing vortices from the tips of the vanes. With a 



174 



view toward the possible application of this tech- 
nique to rotary-wing aircraft the airstream oscil- 
lator was employed to excite the model for several 
"flight" c, Llity 

boundary. Alt i '■ee" the 

data so obtained did give an indication of the 
frequency response characteristics of the c&ntl- 
levered mc m of the 

effects of airspeed, rotor speed, and rotor and 
wing aerodynamics on the overall dynamic response. 

A measure of the gust- induced angle of attack 
(or stream angle) was provided by means of a small 
balsa vane flow direction transmitter (see Fig. 6) 
which gave readings proportional to the stream 
angle. The variation of the vertical component of 



been normalized by the maximum amplitude of the 
stream angle using the curve of Figure 7« 




Figure 6. 0.133- scale semispan tilt-rotor model in 
simulated conversion mode showing boom-mounted 
flow direction transmitter. 

the stream angle for in phase (symmetrical) oscil- 
lation of the biplane vanes is shown in Figure f. 
The curve shown is actually an average of data 
obtained from runs at several tunnel speeds and 
air densities. The amplitude of the stream angle 
has been normalized on the maximum amplitude of 
oscillation of the biplane vanes arid plotted 
against the frequency parameter os/v, where m is 
the frequency of oscillation of the biplane vanes 
in rad/sec and V is the tunnel speed in m/s 
(ft /sec). This parameter is proportional to the 
reciprocal of the wavelength (spacing) between 
vortices shed from the tips of the oscillating 
vanes. 

The frequency response of wing vertical bend- 
ing moment was taken as one measure of system 
response to vertical gust excitation. To ascertain 
the relative influence of rotor and wing aerody- 
namics, three model configurations were employed: 
wing only, with the rotor blade weight replaced by 
an equivalent lumped weight; rotor only, with the 
wing aerodynamic fairings removed; wing and rotor 
combined. For the "flight" condition indicated in 
Figure 3 the r fects of rotor and wing 
aerodynamics are & ■ -, . I ' rjures 8 and 9. In 
each of these figures the wing bending moment has 



/ \ 

Proprotor 




Flow direction ' 
transmitter- 



rad/ft 

_i L 



J_ 



.3 .6 .9 1.2 1.5 1.8 

rad/m 

Wavelength parameter, w/V 

Figure 7- Measured variation of vertical component 
of gust angle with frequency parameter for vanes 
oscillating in phase. 

Comparison of the rotor-on and rotor-off 
response curves for the wing panels on configura- 
tion is shown in Figure 8. Two proprotor-related 
effects are indicated: first, the significant 
contribution of the rotor inplane normal force 
(H-force) to wing bending response, as indicated by 
the relative magnitudes of the bending moments; and 
second, the rotor contribution to wing beam mode 
.damping,* as indicated by the relative sharpness of 
the resonance peaks. The peak amplitudes occur 
when the gust frequency is in resonance with the 
wing beam mode frequency. The peak for the blades- 
off condition is shifted to the higher frequency 
side of the rotor-on peak because the rotor H-force 
decreases the frequency of the wing beam mode. For 
the rotor-on case the bending moment is consider- 
ably larger than for the. rotor-off case throughout 
the range of gust frequencies investigated. The 
wing chord mode frequency (about 2.8 Hz) is within 
the gust frequency range but is absent from the 
.response curves because the gust excitation is 



At this particular airspeed, the rotor was 
still contributing positive damping to the wing 
beam mode. 



175 




Wing airfoil segments installed 
V . 200 knots (102.8 m/s) 

Measured Calculated 

O a - 238 RPM (4 Hz) 

A 



- Rotor off (rotor weight replaced by 
equivalent lumped weight) 



~u\5 O O 275 CS O - 

Simulated full-scale gust frequency, Hz 



Figure 8. Effect of proprotor aerodynamics on wing 
root tending moment amplitude response function. 



in-lb/deg 



N-m/rad 




a « 238 HPM (4 Hz) 

V - 200 knots (102.8 m/s) 

Measured Calculated 

O Wing airfoil segments removed 

A Wing airfoil segments installed 



_1_ 



J_ 



J_ 



_1_ 



_L_ 



_i_ 



0.5 1.0 1.5 2.0 2.5 3A~ 

Simulated full-scale gust frequency, Hz 



J 



Figure 9> Effect of wing aerodynamics on wing root 
bending moment amplitude response function. 



primarily vertical and there is very little 
coupling between the wing beam and chord modes. 

Figures 8 and 9 quite clearly illustrate that 
proprotor s operating at inflow ratios typical of 
tilt-rotor operation in the airplane mode of flight 
are quite sensitive to vertical gusts. This sensi- 
tivity is due to the fact that the proprotors, 
being lightly loaded in the airplane mode of flight, 
operate at low blade mean angles of attack (a) and 
any gust- induced angle of attack is a significant 
fraction of a. 

Hote that good correlation is achieved for 
frequencies up to about 2 Hz beyond which the cal- 
culated responses are much lower than the measured 
values. This discrepancy is thought to be a con- 
sequence of the deviation of the induced gust from 
its nominally one- dimensional nature to one which 
is highly two-dimensional (i.e., varies laterally 
across the tunnel) at the higher frequencies. The 
analytical results shown are based on the assump- 
tion of a one- dimensional gust. Unsteady aerody- 
namic effects may also be a contributing factor to 
the discrepancy. v 

A comparison of the wing panels-on and wing 
panels-off response curves for the rotor-on con- 
figuration is given in Figure 9- As might be 
expected, the wing response for the case in which 
the wing airfoil segments are installed is higher 
than for the rotor alone. The reduced magnitude of 
the response at resonance for the rotor-plus-wing 
combination relative to the rotor alone is due to 
the positive damping contributed by the wing aero- 
dynamics. This increased damping is evident by 
comparing the widths of the resonance peaks. 

Close examination of Figures 8 and 9 reveals 
a very heavily damped, low amplitude resonance 
"peak" at a gust frequency of about 0.8 Hz. This 
resonance is a manifestation of the low-frequency 
(i.e., £1 - fita) flapping mode. Analyses have indi- 
cated that the flapping modes are generally well 
damped for moderate or zero values of flapping 
restraint." These results constitute an experi- 
mental verification. 

These results indicate that "free- flight" 
tilt-rotor models could be used to measure the 
frequency response functions needed in gust 
response analyses. This would be a fruitful area 
for future analytical and experimental research. 

(b) January 1970 

A joint HASA/Bell/Air Force test program was 
conducted in the transonic dynamics tunnel in 
January 1970 for the purpose of investigating any 
potential problem areas associated with the folding 
proprotor variant of the tilt-rotor concept. The 
model used in this study was the same model 
employed in the first investigation but modified to 
permit rapid feathering and unf eathering of the 
proprotor and to include a blade fold-hinge. The 
main objectives were to investigate stability at 
low (including zero) rotor rotational speeds, 



176 



during rotor stopping and starting, and during 
blade folding. All objectives of the test program 
were met. No aeroelastic instabilities were 
encountered during the blade folding sequence of 
transition, the blade loads and/or the feathering 
axis loads inboard of the fold hinge being identi- 
fied as the critical considerations from a design 
point of view. The stop- start portion of the test 
indicated that additional flapping restraint would 
be required to minimize flapping during rotor 
stopping.* Stability investigations conducted over 
a wide range of rotor speed identified an apparently 
new form of proprotor instability involving the 
rotor at low and zero rotational speeds. The 
influence of several system parameters on this 
instability was established both experimentally and 
analytically. 6 

Proprotor /Pylon Stability . For the stability 
investigation a reference configuration was again 
established. This consisted of the basic Model 266 
configuration with the pylon locked to the wing tip 
in both pitch and yaw, a hub restraint of 
117,685 N-m/rad (86,800 ft-lb/rad), 03 = -0.595 rad 
(-22.5°), a simulated wing fuel weight distribution 
of 15$, and the wing aerodynamic fairings installed. 
The flutter boundary obtained for this configuration 
and that for 03 = -0.558 rad (-52°), are shown in 
Figure 10 as a function of rotor speed. Open sym- 
bols denote flutter points. Excessive vibration 
resulting from operation near resonances with the 
pylon/wing or blade modal frequencies often limited 
the maximum attainable airspeed. These points are 
indicated by the solid symbols. The annotation to 
the right of the flutter boundaries indicates that 
the model experienced several modes of flutter. 
The predicted flutter modes and frequencies were 
in agreement with the experimental results. The 
nature of these flutter modes is discussed below. 

For Q greater than about k Hz (2^0 rpm) 
instability occurred in the wing beam mode and had 
the characteristics described earlier for the 
September 1968 test. For fi between about 2 Hz 
(120 rpm) and k Hz (2^0 rpm) the motion at flutter 
was predominantly wing vertical bending and rotor 
flapping with the hub precessing in the forward 
whirl direction. Examination of the root loci 
indicated that this instability was associated with 
the low- frequency (i.e., 0. - cup) flapping mode root 
becoming unstable. The subcritical response through 
flutter for 83 = -0.558 rad (-52°) and a = 2.86 Hz 
(172 rpm) is shown in Figure 11 where, in addition 
to the measured wing beam mode damping and frequency, 
the calculated variation of both the wing beam and 
low-frequency flapping modes is shown. These 
results illustrate an interesting modal response 
behavior similar to that described by Hall. 2 The 
wing beam mode, being least stable at low airspeeds, 
is at first dominant. As airspeed increases, how- 
ever, its damping continually increases. The damp- 
ing of the fl - <»q flapping mode meanwhile is 
continually decreasing. Crossover occurs analyti- 
cally at Ikh m/s (280 lets) at a damping of 17$ of 

These aspects of this investigation are given 
detailed treatment in Reference 9- 




300 400 

Airspeed, knots 



100 200 

Airspeed, meters/sec 



Figure 10. Model 266 flutter boundaries showing 
variation in character of flutter mode as rpm is 
reduced to zero. 

critical. Beyond 280 knots, the fi - fflp flapping 
mode is the dominant mode and very abruptly becomes 
unstable as airspeed is increased. Hence, a tran- 
sition from a dominant wing beam mode to a dominant 
flapping mode with an accompanying change in fre- 
quency. Since the flapping mode frequency is only 
slightly less than the wing beam mode in the 
vicinity of flutter there is only a gradual, albeit 
distinct, transition in the frequency of the wing 
beam mode as the flapping mode begins to predominate 
over the wing mode. Examination of the Q - cup 
flapping mode eigenvector indicated that a larger 
amount of wing vertical motion was evident in this 
mode than in the wing beam mode eigenvector. This 
implies that the predominant motion in the flutter 
mode is not necessarily determined by the root 
which analytically goes unstable as airspeed is 
increased but the frequency at which a root goes 
unstable. 

Below about 2 Hz (120 rpm) instability is in 
the high-frequency (i.e., JJ + coo) flapping mode and 
is characterized by large amplitude flapping, the 
rotor tip-path-plane exhibiting a precessional 
motion in the forward whirl direction. The modes 
of instability at zero rotational speed were similar 
in character to those at low rotor speeds but with 
larger amplitudes of flapping. Although the rotor 
was not turning, the flapping behavior of the blades 



177 





^ 



200 300 

Airspeed, knots 



Wife" 



100 150 

Airspeed, meters/sec 



Figure 11. System response characteristics for 
flutter at Jl = 172 rpm and S3 = -3?.". 

was patterned such that the tip-path-plane appeared 
to be wobbling or whirling in the forward direction. 
Negligible wing motions accompanied the flapping 
motion. Figure 12 shows the variation of flap 
damping with airspeed. A hub damping of £g = 0.015 
was originally used in calculating the stability 



6 3 » -.393 rad (-22.5°) 



? R - .020 




O Measured 
Calculated 



300 



„sr 



50 



Airspeed, knots 

_l_ 



100 
Airspeed, meters/sec 



150 



boundaries, leading to very conservative values for 
the flutter speed at the low rotor speeds. Based on 
the results of Figure 12, which indicate that the 
rotor hub structural damping is closer to £ R = 0.025,'^ 
the stability boundaries were recalculated using 
£r = 0.025. The predicted boundaries in Figure 10 
reflect this change. 

The small region of increased stability in the 
region of 0.8 Hz (kQ rpm) is due to a favorable 
coupling of the flapping mode with wing vertical 
bending. 

The instabilities encountered at low and zero 
values of rotational speed were quite mild and had 
a relatively long time to double amplitude. The 
necessity of limiting the flapping amplitude during 
the feathering sequence of transition dictates that 
significantly increased values of hub restraint are 
needed as rotor rotational speed is reduced to zero. 
Since increased flapping restraint was found to 
stabilize this mode° this instability is probably 
only of academic interest, at least for the config- 
uration tested. However, since it was a new 
phenomenon and was not understood a,t the time of 
the test, attention was directed to assessing the 
effect of the variation of several system parameters 
on the flutter speed. Both experimental and analyt- 
ical trend studies were conducted for this purpose. ° 
Based on these studies it was concluded that rotor 
precone was the primary cause of the instability. 

Blade Flapping . In the feathering sequence of 
transition flapping sensitivity to a given mast 
angle of attack varies with rotor rotational speed. 
A typical variation of steady- state one-per-rev 
flapping response is given in Figure 1>. These 
data were taken to establish a steady-state flapping 
response baseline for evaluating the transient 



rad 

.ior 



a 

«; .04 



Measured 

o 



- .02618 rad (1.5°) 
-a_ - .01745 rad (1.0°) 








100 200 

Proprotor speed, RPM 

1 1 



Proprotor speed, Hz 



Figure 12. Variation of fl + (Da flapping mode Figure 13. 
damping with airspeed for zero rpm. 



Variation of blade flapping with rotor 
rpm. 



178 



flapping response during the feathering portion of 
the test. Since the proprotor mast was not affixed 
to a rigid backup structure the wind- on mast angle 
of attack was not known (it was nominally 1° ) . The 
important conclusion following from Figure 15 is 
that the measured trend is predicted correctly. 
The peak in the flapping response occurs when the 
rotor rotational speed is in resonance with the 
flapping natural frequency in the rotating system. 

Grumman Helicat (March 1971 ) 

A wide variety of technical considerations 
confront the structural dynamicist in the design 
of a proprotor VTOL aircraft. Perhaps the most 
celebrated consideration has been that of prop- 
rotor/pylon whirl flutter, having been the concern 
of many investigators in both government and 
industry. Several years ago Baird 1 - raised the 
question of whether proprotor whirl flutter, in 
particular forward whirl flutter, could be pre- 
dicted with confidence. His skepticism was 
prompted by the lack of agreement between the 
experimental results obtained with several small 
models of flapping-blade propellers and the corre- 
sponding theoretical predictions.^ To provide a 
large data base from which to assess the predict- 
ability of proprotor whirl flutter, a joint NASA/ 
Grumman investigation was conducted in the tran- 
sonic dynamics tunnel employing an off-design 
research configuration of a 1/4.5- scale semispan 
model of a Grumman tilt-rotor design designated 
"Helicat" (Fig. 14). This design is characterized 




Figure 14. Grumman "Helicat" tilt-rotor model in 
whirl flutter research configuration. 

by a rotor which incorporates offset flapping 
hinges in contrast to the Bell rotor in which the 
blades are rigidly attached to, the hub which is in 
turn mounted on the drive shaft by a gimbal or 
universal joint housed in the hub assembly. The 
Helicat model was specifically designed to permit 
rather extensive parametric changes in order to 
provide a wide range of configurations. These 
variations included pylon pitch and yaw stiffness 
and damping, hinge offset, and pitch- flap coupling. 
To obtain flutter at low tunnel speeds, a reduced- 
stiffness pylon-to-wingrtip restraint mechanism 



which permitted independent variations in pitch 
and yaw stiffness was employed. The resulting 
pylon-to-wing attachment was sufficiently soft to 
insure that the wing was effectively- a rigid backup 
structure. Details concerning this model as well 
as a summary of results are contained in 
Reference 11. 

Some whirl flutter results are given in Fig- 
ures 15 to 17, where flutter advance ratio Vp/QR 
is plotted versus pylon frequency nondimensionalized 
by the rotor speed. The effect of 85 on stability 



K 

> 



1 



1.4 


Symmetrical pylon frequencies 
e/R » .05 




1.2 


- 


6 3 > .349 rad (20°) 


1.0 




6- « .118 rad (6.75°) / 

i / 6 3 " - 524 rad ( 30 °) 


.8 




P / ! 

I of 4 




.6 
.4 




■■/ /*/ 

/ A^ 7 

1 f£{ Measured 

'/ / ° 

' / A 

/ □ 

1 1 1 


Calculated 


.2 


1 1 



1.0 



Pylon frequency, cycles/rev 



Figure 15. 



Effect of pitch-flap coupling on whirl 
flutter. 



is shown in Figure 15 for the case in which the 
pylon pitch and yaw frequencies are identical and 
e/R set to 0.05- Many of the configurations were 
not exactly symmetrical in the frequencies. These 
data were adjusted to reflect a symmetric frequency 
support condition using Figure 18 of Reference 11. 
The results show a strong increase in flutter 
advance ratio (and hence flutter speed for a fixed 
rpm) with increasing pylon support stiffness and 
decreasing 63. All flutter was in the forward 
whirl mode except for the two points denoted by the 
solid symbols, which were in the backward mode. 
The analytical results shown assumed a symmetric 
frequency configuration and, since the structural 
damping varied somewhat, an average value of damp- 
ing of i = 0.01 in pitch and I = 0.02 in yaw. 
The analytical results shown were obtained using the 
theory of Reference 6 which is based on the assump- 
tion of a gimbaled rotor. For analysis purposes the 
restoring centrifugal force moment from the offset 
flapping hinge was represented by introducing an 



179 



equivalent hub spring which preserved the blade 
in-vacuum flapping natural frequency in the manner 
indicated in Appendix B of Eef erence 6. 

The beneficial effect of increased hinge off- 
set is demonstrated in Figure 16. The results for 
the 13$ hinge offset are particularly noteworthy 



fa 

> 



1.4 


Symmetrical pylon frequencies 
6 3 • .349 rad (20°) 






1.2 




e/R 


. .05 




1.0 
.8 




e/R m .13 / 
/ / 

^ / 
/ / 

/ / 
/ °/ 






.6 










.4 


- 


A / / 
/ (J 

/ y 

/ Measured 

' O 
A 


Calculated 




.2 












1 1 ' 


1 


.1 







.2 .4 .6 


.8 


i.o 



Figure 16. 



Pylon frequency, cycles/rev 

Effect of hinge offset on whirl 
flutter. 



in that both forward and backward whirl motions 
were found to occur simultaneously; in effect, the 
flutter was bimodal. Theory also predicted this 
bimodal behavior, the forward and backward whirl 
modes being within a few knots of each other 
analytically. 

The effect of asymmetry in the pylon support 
stiffness is shown in Figure 17 . Again the sym- 
metric frequency data reflect adjustments to true 
symmetry for configurations which were nearly, but 
not exactly, symmetric. The nonsymmetric results 
reflect actual measured values, the lower of 
either the pitch or yaw frequencies being plotted. 
It was analytically shown^ that for sufficient 
asymmetry in the pylon support stiffness increas- 
ing the asymmetry more does not increase the flut- 
ter speed. The data for the nonsymmetric fre- 
quencies are an experimental demonstration of this 
fact. Flutter in all the asymmetric conditions 
was in the backward whirl mode. 



1.4 



1.2 



l.o- 



> 



I 



6 3 - .349 rad (20°) 




e/R « .05 


/ 
/ 









1 . 




1 1 




p I 


_ 


1 




1 




1 / 


o< 


I 


Bl 


o 


1 




1 




1 


OT 


$1 




mi 




£<V 




/ y 




uy 


Measured Calculated 


/<§7 




*-* W B' ty 


1/ 

1 
1 

1 


A W 


1 1 1 



1.00 

1.50 

.75 



.4 . .6 

Pylon frequency, cycles/rev 



1.0 



Figure 17. 



Effect of pylon support stiffness on 
whirl flutter. 



Bell Model 300 
(a) August 1971 

A joint NASA/Bell investigation employing a 
1/5- scale aerodynamic model of a Bell tilt-rotor 
design designated the Model 300 was conducted in 
the transonic dynamics tunnel in August 1971 for 
the purpose of providing the longitudinal and 
lateral static stability and control characteris- 
tics and establishing the effect of proprotors on 
the basic airframe characteristics in both air and 
freon. Use of freon permitted testing at full- 
scale Mach numbers and near full-scale Reynolds 
numbers.. Flapping was measured in both air and 
freon for several values of tunnel speed over a 
range of sting pitch angles. The resultant flap- 
ping derivatives, obtained by evaluating the slopes 
of the flapping amplitude versus pitch angle curves, 
are shown in Figure 18. Since the range of inflow 
ratios over which the derivatives were measured was 
the same in air and freon and the test medium 
densities at the simulated conditions were about 
the same, an indication of the effects of Mach 
number on the flapping derivatives can be obtained 
by comparing the air and freon results. The speed 
of sound in freon is approximately half that in 
air so that for a given tunnel speed (or inflow 
ratio) the Mach number in freon is about twice that 
in air. The calculated results reflect the 



180 



o 

to 






Measured Calculated 



Air 
Freon 




shafting is also employed in wind-tunnel models. 
The availability of thrust damping to provide a 
stabilizing force for yawing motion is dependent 
on the structural integrity of this cross- shafting 
and has implications which are pertinent to both 
full-scale flight and model testing. Consider the 
case of a windmilling "free-flight" model. A fully 
effective interconnect maintains synchronization of 
the rotor speeds during any motions. A yawing 
motion of the model to the left, say, as might 
occur during a disturbance, generates blade angle- 
of -attack changes which decrease the lift of blade 
elements on the right rotor and increase the lift 
of blade elements on the left rotor. This produces 
resultant perturbation thrust changes which tend to 
damp the yawing motion, as depicted in the sketch 
in the right-hand portion of Figure 19. If the 



Rotor interconnect shaft 
Engaged (AT > 0) 
Disengaged (AT * 0) 



Measured Calculated (Bell) 

O 



1 


Inflow ratio at 0.75 Wade 


radius, 


v/m 

1 




1 


.10 .20 

Mach number in air 

.1 i 




.30 
1 1 



■a so 
| 



.40 .60 

Mach number in freon 



.80 



Figure 18. Effect of Mach number on proprotor 
flapping. 

variation of Sx with blade pitch. Drag was 
neglected in the calculated results shown for air 
but was accounted for, in an approximate manner, 
in the results shown for freon." The drag rise 
associated with operation at high Mach numbers is 
seen to reduce flapping as Mach number is increased 
and suggests that calculations based on the neglect 
of blade drag will predict conservative values of 
flapping at Mach numbers where drag is important. 

These data are believed to be the first which 
provide an indication of the effects of Mach number 
on blade flapping. 

(b) March 1972 

The most recent investigation conducted in 
the transonic dynamics tunnel utilized a l/5- scale 
dynamic and aeroelastic "free-flight" model of the 
Bell Model 300 tilt rotor for the purpose of 
demonstrating the required flutter margin of safety 
and to confirm that the aircraft rigid-body flight 
modes are adequately damped.- 1 - 2 During this test 
the importance of rotor thrust damping on stability 
of the Dutch roll mode was investigated. This 
damping is associated with rotor perturbation 
thrust changes which can be generated during axial 
oscillations of the rotor shaft and constitutes a 
positive damping force on aircraft yawing motions. 

The rotors of tilt-rotor aircraft are gener- 
ally designed to have an interconnecting shaft 
between the two rotor/engine systems to provide 
synchronization of the rotor speeds and to insure 
that in the event of an engine failure either 
engine may drive both rotors. Interconnect 




180 220 260 300 340 

Tunnel speed, ft/sec 

50 ii loo 

Tunnel speed, meters/sec 

Figure 19. Thrust damping effects on tilt-rotor 
Dutch roll mode stability. 

interconnect is absent, the rotors are able to main- 
tain their inflow angle and, hence, angle of attack 
by increasing or decreasing rotor speed. The per- 
turbation thrust changes thus go to zero and the 
stabilizing contribution of this damping to the air- 
craft yawing motion is lost. The effects of thrust 
damping on the stability of the Dutch roll mode was 
investigated by measuring the Dutch roll mode damp- 
ing as a function of tunnel speed for the cases in 
which the model interconnect was engaged and dis- 
engaged. Some typical results are shown at the 
left of Figure 19 along with the damping levels 
predicted by Bell. The substantial contribution of 
thrust damping to total damping is quite apparent. 
It is of interest to point out that for the rotors 
contrarotating in the direction indicated in the 
sketch at the right of Figure 19 (inboard up) the 
perturbation thrust changes accompanying an aircraft 
rolling angular velocity are destabilizing on Dutch 
roll motion. For contrarotating rotors turning in 
the opposite direction (inboard down) the &T due 
to both yawing and rolling motion are stabilizing 
on Dutch roll motion. 

Rotor rpm governors of the type which maintain 
rpm by blade collective pitch changes while main- 
taining constant torque are being considered for 
use on full-scale tilt-rotor aircraft. With the 
interconnect engaged, full thrust damping is avail- 
able (assuming a perfect governor). However, in 
the event of an interconnect failure, the governors 



181 



would respond to any rpm changes by varying blade 
collective pitch in a manner which tends to main- 
tain the original blade angle- of- attack distribu- 
tion and hence torque. This is aerodynamically 
equivalent to the windmilling case with no inter- 
connect. It is axiomatic that tilt-rotor aircraft 
must be designed to have stable Dutch roll charac- 
teristics should an interconnect failure occur 
anywhere within the flight envelope. 

Some Additional Results Applicable to the 
Bell Model 300 Tilt Rotor 



- nnrHS-nOFQ o <~> 








Measured 




D Design stiffness test stand 




O Ifi design stiffness test stand 




- A 1/5, scale aeroelastic model 




Calculated (design stiffness test stand) 








Blades rigid inplane 





A dynamic test of a flight-worthy proprotor 
for the Bell Model 300 tilt- rotor aircraft was 
conducted in the NASA Ames full-scale wind tunnel 
in July 1970 (Fig. 20). Two different test stands 




Figure 20. Bell 25-foot flight-worthy proprotor 
in NASA Ames full-scale tunnel for dynamic 
testing. 

were used. One duplicated the actual stiffness 
characteristics of the Model 300 wing; the other 
was one- fourth as stiff. By using the reduced 
stiffness spar and operating the proprotor at one- 
half its design rotational speed it was possible 
to preserve the per-rev natural frequencies of the 
wing and simulate, at any given tunnel speed, the 
inflow of flight at twice that speed. This expe- 
dient did not, however, maintain the blade per-rev 
elastic mode frequencies or simulate compressibil- 
ity effects on rotor aerodynamics. 

Some results from the full-scale test are 
compared with data obtained from a test of a 1/5- 
scale model and theory in Figure 21. Note that 




Airspeed, knots 



100 200 

Airspeed, meters/sec 



Figure 21. Model/full-scale comparisons of wing 
beam mode damping and frequency variation with 
airspeed for Bell Model 300. 

the calculated results are based on the use of the 
design stiffness test stand characteristics. To 
provide for an indication of the effect of blade 
inplane flexibility on stability, the predicted 
results for the case in which the blades are 
assumed to be rigid inplane are also shown. The 
predicted increase in damping at about 103 m/sec 
(200 kts) for the case in which blade inplane flex- 
ibility is included is associated with coupling of 
the blade first inplane cyclic mode with wing verti- 
cal bending. For the range of tunnel speed over 
which full stiffness test stand data are available, 
the results are in good agreement with theory 
assuming flexible blades. Note that a significant 
stabilizing effect is predicted for the Model 300 
as a consequence of blade inplane flexibility. 
This trend is in contrast to that predicted for the 
Model 266. The data for the quarter- stiffness test 
stand are in agreement with theory assuming rigid 
blades because operation at half the design rpm has 
effectively stiffened the blades by a factor of h. 
The 1/5- scale model data are also seen to be in 
better agreement with analysis based on the assump- 
tion of rigid blades. This is because the model 
hub employed at the time the data were obtained was 
too stiff. If this increased stiffness is taken 
into account the predicted damping is in agreement 
with theory (Fig. 22). The model/full-scale com- 
parisons shown in Figure 21 indicate that assessment 
of full-scale stability can be made on the basis of 
results of small-scale model tests. 



182 



10 



,„- 4 



Q- 

s 



MEASURED CALCULATED 



O 
A 



WING BEAM MODE 
• WING CHORD MODE 



O O 




(P'zoo 



n^ 



'100 



300 
AIRSPEED, knots 



400 



500 



150 200 

AIRSPEED, meters/sec 



250 



Figure 22. Variation of wing beam and chord mode 
damping with airspeed for 1/5- scale aeroelastic 
model of Bell Model 300. 

Conclusions 

An overview of an experimental and analytical 
proprotor research program being conducted within 
the Aeroelasticity Branch of the NASA Langley 
Research Center has been presented. On the basis 
of the particular results shown herein the follow- 
ing basic conclusions can be drawn: 

(1) A proprotor /pylon/wing system can exhibit 
a wide variety of flutter modes depending on the 
degree of fixity of the pylon to the wing, rotor 
characteristics, and rotor rotational speed. In 
particular, for pylons which are rigidly affixed 
to the wing tip, the instability can occur in 
coupled pylon/wing, pylon/wing/rotor, or rotor 
modes j for pylons which are soft-mounted to the 
wing, a true whirl instability akin to classical 
propeller whirl flutter can occur. 

(2) Lightly loaded proprotors operating at 
inflow ratios typical of tilt-rotor operation in 
the airplane mode of flight exhibit a marked 
sensitivity to gust excitation. 

(3) Blade inplane flexibility can have a 
significant effect on stability. 

(k) A significant contribution to aircraft 
lateral-directional (Dutch roll) stability arises 
from rotor thrust damping. Since the availability 
of this thrust damping is dependent on the integ- 
rity of the rotor interconnect shaft, tilt-rotor 
aircraft must be designed to have acceptable 
lateral-directional response characteristics should 
an interconnect failure occur anywhere within the 
operating envelope. 



(5) Proprotor whirl flutter, both backward 
and forward, can be predicted with simple linear- 
ized perturbation analyses using quasi- steady rotor 
aerodynamics. 

(6) For strength designed wings, wing aerody- 
namics have only a slight stabilizing effect on 
proprotor flutter speeds. 

(7) The drag rise associated with proprotor 
operation at high Mach numbers reduces blade flap- 
ping and suggests that calculations based on the 
neglect of blade drag will predict conservative 
values of flapping at Mach numbers where drag is 
important. 

The analytical portion of this research pro- 
gram is continuing. Attention is presently being 
directed toward refining the existing stability and 
response analyses and extending them by including 
additional degrees of freedom. 

Acknowledgment s 

The author acknowledges the assistance pro- 
vided by Bell and Grumman in preparing and testing 
the models employed in the investigations conducted 
in the transonic dynamics tunnel. Particular thanks 
are extended to Troy Gaffey of Bell for his general 
advice and assistance since the initiation of this 
research program and to Jerry Kohn of Grumman for 
performing the correlations with the data obtained 
during the whirl flutter investigation using the 
Helicat model. 

References 

1. Deckert, W. H. , and Ferry, R. G. , LIMITED 
FLIGHT EVALUATION OF THE XV-3 AIRCRAFT, Air 
Force Flight Test Center, Report TR-60-lt-, 
May i960. 

2. Hall, W. E. , PROP-ROTOR STABILITY AT HIGH 
ADVANCE RATIOS, Journal of the American Heli- 
copter Society, June 1966. 

3. Edenborough, H. K., INVESTIGATION OF TILT- 
ROTOR VTOL AIRCRAFT ROTOR-PYLON STABILITY, 
Journal of Aircraft, Vol. 5, March-April 1968. 

k. Reed, W. H., III., REVIEW OF PROPELLER- ROTOR 
WHIRL FLUTTER, NASA TR R-261+, July I967. 

5. Gaffey, T. M. , Yen, J. G. , and Kvaternik, R. G. , 
ANALYSIS AND MODEL TESTS OF THE PROPROTOR 
DYNAMICS OF A TILT-PROPROTOR VTOL AIRCRAFT, 
presented at the Air Force V/STOL Technology 
and Planning Conference, Las Vegas, Nevada, 
September 1969. 

6. Kvaternik, R. G. , STUDIES IN TILT-ROTOR VTOL 
AIRCRAFT AEROELASTICITY, Ph. D. Dissertation, 
Case Western Reserve University, June 1973. 

7. Young, M. I., and Lytwyn, R. T., THE INFLUENCE 
OF BLADE FLAPPING RESTRAINT ON THE DYNAMIC STA- 
BILITY OF LOW DISK LOADING PROPELLER- ROTORS, 
Journal of the American Helicopter Society , 
October 1967. 



183 



8. Gilman, J. , Jr. , and Bennett, R. M. , A WIND- 
TUNNEL TECHNIQUE FOR MEASURING FREQUENCY- 
RESPONSE FUNCTIONS FOR GUST LOAD ANALYSIS, 
Journal of Aircraft, Vol. 3, November- 
December 1966. 

9. Yen, J. G. , Weber, G. E. , and Gaf f ey, T. M. , 
A STUDY OF FOLDING PROPROTOR VTOL AIRCRAFT 
DYNAMICS, AFFDL-TR-71-7 (Vol. I), September 
1971. 

10. Baird, E. F. , CAN PROP-ROTOR STABILITY BE 

PREDICTED?, presented at the Aerospace Flutter 
and Dynamics Council Meeting, San Francisco, 
California, November 12- l^, 1969. 



11. Baird, E. F. , Bauer, E. M., and Kohn, J. S. , , 
MODEL TESTS AND ANALYSIS OF PROP-ROTOR DYNAMICS 
FOR TILT-ROTOR AIRCRAFT, presented at the Mid- 
east Region Symposium of the American Helicop- 
ter Society, Philadelphia, Pennsylvania, 
October 1972. 

12. Marr, R. L. , and Neal, G. T. , ASSESSMENT OF 
MODEL TESTING OF A TILT-PROPROTOR VTOL 
AIRCRAFT, presented at the Mideast Region 
Symposium of the American Helicopter Society, 
Philadelphia, Pennsylvania, October 1972. 



184 



COMPARISON OF FLIGHT DATA AND 

ANALYSIS FOR HINGELESS ROTOR 

REGRESSIVE INPLANE MODE STABILITY 

by 

W. D. Anderson 

and 

J. F. Johnston 

Lockheed California Co. 

Burbank, California 



Abstract 

During the development of the AH-56A, a considerable 
amount of analytical and experimental data was obtained on the 
stability of the regressive inplane mode, including coupling with 
other modes such as body roll and rotor plunge. The data were 
obtained on two distinctly different control systems; both gyro 
controlled, but one with feathering moment feedback and the 
other with direct flapping feedback. The paper presents a review 
of the analytical procedures employed in investigating the 
stability of this mode, a comparison of analytical and experi- 
mental data, a review of the effect of certain parameters, 
including blade droop, sweep, 63, aj , vehicle roll inertia, inplane 
frequency, rpm and forward speed. It is shown that the stability 
of this mode is treatable by analysis and that adequate stability is 
achievable without recourse to auxiliary inplane damping devices. 



M. 



Notation 

B subscript referring to blade feathering 

Cj j2 measure of damping, cycles to half amplitude 

F subscript referring to fuselage 

g structural damping ratio 

I imaginary part of root, rad/sec 

Kg collective feathering stiffness, ft-lb/rad/blade 

Ko root flapping moment per unit of blade flapping, ft-lb/rad 

L rotor lift, pounds 



M 



moment, ft-lb 



Presented at the AHS/NASA-Ames Specialists' Meeting on 
Rotorcraft Dynamics, February 13-15, 1974. 



xy 



N R 



Xc 



Po 

63 

a 

€ x 



blade product of inertia about feathering axis, slug-ft 2 

normal rotor speed 

real part of root, per second, subscript referring to rotor, 
or rotor radius, ft 

airspeed, knots 

airframe longitudinal motion, ft 

airframe lateral motion, ft 

airframe vertical motion, ft 

pitch lag coupling - positive nose up feather due to lag aft 
of blade 

rot6r blade collective flapping or coning, radians 

pitch flap coupling angle - tan"' (-8/p) 

to indicate partial differentiation 

rotor blade cyclic inplane motion sine component, positive 
forward, radians 

rotor blade cyclic inplane motion cosine component, 
positive to the right, radians 



t, 


fraction of critical damping 


e 


pitch motion, radians 


% 


blade collective feathering, radians 


\ 


blade effective sweep angle, radians 


is 


servo time constant, sec 


♦ 


roll motion, radians 


10 


frequency, rad/sec 


nip 


inplane natural frequency, rad/sec 


Q 


rotor rotational speed, rad/sec 



185 



In hingeless rotors two fundamental types of coupled rotor 
body inplane mode stability problems exist. One is associated 
with a soft inplane system having the inplane frequency less than 
rotational speed, and the other with a stiff inplane system where 
the inplane frequency is above rotational speed. The soft inplane 
system when coupled With a basic body mode is unstable in the 
absence of aerodynamics, and therefore its stability must be 
provided by aerodynamic or auxiliary damping. This type of 
system is discussed in References 1, 2, and 3. In contrast, the 
stiff inplane system does not exhibit this inherent mechanical 
instability, and so its stability is less dependent upon aero- 
dynamic or auxiliary damping. 

Both types of modes, however, are subject to aeroelastic 
phenomena which can be stabilizing or destabilizing. Also, both 
types can exhibit response characteristics caused by pilot and/ or 
gust inputs which are undesirable. The critical inplane mode in 
the soft inplane system is advancing in the stationary system, 
whereas for the stiff inplane system, the mode is regressive. The 
frequency of the mode in each case is the magnitude of (w . 



n)or(n-w mp ). 



nip 



This paper deals specifically with the stiff inplane system. 
The various types of coupled rotor body regressive inplane 
stability /response problems associated with this type of system 
are discussed. The paper deals with both a feathering moment 
feedback gyro-controlled system and a direct flapping moment 
feedback gyro-controlled system. 

These two types are described in more detail in Reference 4. 
The inplane mode characteristics of the direct flapping moment 
feedback type system would be more characteristic of any direct 
control hingeless rotor system employing a stiff inplane rotor. 

The stiff inplane hingeless rotor system is worthy of serious 
consideration because of its inherent characteristics of being free 
from ground/ air resonance mechanical instability type 
phenomena and its ability to provide a stable, highly maneuver- 
able, rotary wing vehicle. 

The absolute level of the stability of the inplane mode is not 
the only consideration in establishing design criteria. An equal or 
even more important criterion is that of response of the mode as 
a result of pilot or gust disturbances. The stability of the mode 
can appear adequate, but if it is easily excited by either pilot or 
gust inputs, the mode can be unsatisfactory. Conversely, the 
mode may exhibit very low damping characteristics, but not be 
easily excited by either pilot and or gust excitations, and be quite 
satisfactory because no high loads or undesirable body motions 
occur. 

Besides the basic stability considerations of the regressive 
inplane mode, certain other basic types of coupled rotor body 
regressive inplane mode stability /response problems may be 
encountered. Some of these may be either low or high airspeed 
phenomena or virtually independent of airspeed. 

It is not intended here to go into a complete theoretical 
treatise describing each of these types of phenomena, but the 
effects of some parameters and flight conditions on stability/ 
response characteristics for particular rotor vehicle configurations 
are presented. Because stability and response characteristics 
depend on considerations of the detail design, generalized 
conclusions cannot always be drawn as to the effect of each 
parameter discussed. 



The fundamental types of regressive inplane mode stability/ 
response problems discussed include those associated with: 

• The basic regressive inplane mode. 

• The coupled regressive inplane body roll mode, 

• A coupled regressive inplane-roll-rotor plunge mode. 



The first type can exhibit itself to the pilot as an apparent 
rotor weaving or rotor disc fuzziness with very little body 
response. The second type appears to the pilot as rotor tip path 
plane response and body roll or just body roll oscillation. 
Depending on the frequency of this mode, and the gain of the 
feed forward loop of the control system, this mode may be 
subject to pilot-induced or pilot-coupled oscillations. The last 
type basically exhibits itself to the pilot as a rotor umbrella 
mode and a vertical plunge of the vehicle. This mode can exist 
in the absence of the inplane mode but can be seriously 
affected by its presence. The mode has been characterized as 
a "Hop" mode because of the plunge response of the airframe. 



Analytical Method 

The analytical method employed in the study consists of a 
fundamental 13-degree-of-freedom representation of the coupled 
rotor body control system. The body is characterized by 5 
degrees of freedom — yaw being ignored. Likewise, the rotor is 
represented by eight multiblade coordinates including rotor disc 
plunge, pitch and roll; lateral and longitudinal inplane; and pitch, 
roll, and collective elastic feathering/ torsion. The model is shown 
schematically in Figure 1. The equations are solved as linear 
constant coefficient equations, modified as required to represent 
the control system being considered. 

For the solutions shown, certain simplifying assumptions 
were made. These include neglecting the effects of retreating 
blade stall, reverse flow, and advancing tip Mach number. 




COLLECTIVE FLAPPING 
AND FEATHERING 



FUSELAGE MOTIONS 



Figure 1 . Description of Analytical Model. 



186 



The model includes the effects of elastic coupling 
phenomena of inplane moments times flapping deflections 
causing feathering moments. A simplified inflow model is used 
which characterizes the induced velocity from low transition 
speed to high speed as a trapezoidal distribution withupwash at 
the front of the rotor and downwash at the back of the rotor, for 
positive rotor lift 

To gain a fairly comprehensive understanding of such modes 
as the regressive inplane mode as well as the other coupled rotor 
body fundamental aeroelastic modes, it is felt this type of model 
is a necessity. 

Effect of Parameters 

Following is a discussion of each type of mode and the 
relevant parameters which affect the stability /response of the 
mode together with some parametric effects. 

Basic Regressive Inplane Mode 

The basic regressive inplane bending mode can be lightly 
damped without any problem, provided it is not easily excited 
by the pilot or by gusts. The mode if its frequency is well 
separated from any other rotor body control mode frequencies, 
behaves very much as a single blade would behave. 

Figure 2 shows a complex plane plot of a typical mode of 
response of the regressive inplane mode. It is noted that motions 
of all other degrees of freedom are small compared with the 
response of the blade inplane € x and 6y. This would be for a 
case where the inplane mode frequency is well separated from 
other rotor-body control mode frequencies, and the inter- 
coupling with these modes is not large. In this case the inplane 
mode is reasonably above the body roll mode, with a frequency 
ratio of 1.32 in nonrotating coordinates. 






*F 




| UP 



^T 



FWD 



-^ 



VIEW LOOKING IN AT 
TIP OF BLADE 



For this case, the effect of several parameters is examined. 
The stability/response of the mode is largely controlled by such 
parameters as discussed in Reference 4 and 5. That is, parameters 
such as blade kinematic and elastic pitch-flap-lag couplings are 
extremely important. Also, such items as precone, feather 
bearing location, hub/blade stiffness distributions and control 
system flexibility play important roles in the stability of the 
mode. A discussion of each of several parameters affecting the 
stability of the mode follows. 



Inplane Damping. Figure 3 shows a locus of roots as a 
function of equivalent structural damping in the inplane mode. 
Identified on this figure are lines of constant damping in terms of 
one over cycles to half amplitude, 1/C| m. This figure shows the 
expected results. That is, computing the approximate change in 
fraction of critical damping from the root locus plot by taking an 
increment of change in the real part of the root due to the 
change in modal structural damping, g§j, and dividing it by the 
sum of the imaginary part of the root and rotational speed 
results in a value of damping roughly half the change in 
equivalent structural damping. This is consistent with the well 
known relationship of g =«2C where g « 1. 



Since some centrifugal stiffening effect is existant in the 
inplane mode, the actual change in damping, 24 , is less than the 
change in equivalent structural damping in the mode. 




U) 



-3 -2 -1 

REAL PART OF ROOT 



Figure 2. Complex Plane Plot of Typical Regressive 
Inplane Mode, V = 20 KN, w € /u Roll = 1.32. 



Figure 3. Locus of Roots 
Damping, V = 



-Effect of Structural 
= 20 KN. 



187 



Kinematic Pitch Lag Coupling. Figure 4 shows the change in 
damping due to a variation in pitch lag coupling. The indicated 
sense of this coupling for improved stability is nose down 
feathering due to lag aft of the blade for the stiff inplane system 
whereas Reference 6 showed that the opposite coupling is stabi- 
lizing for articulated or soft inplane system. The effect of this 
parameter on the stability of the regressive cyclic inplane mode is 
similar to the effect on the stability of the reactionless inplane 
mode as indicated in Reference 5. The fundamental mechanism 
of the a j coupling is to cause blade flapping to couple through 
coriolis forces to damp the inplane mode. This can be deduced 
by examining Figure 2. 



of up flapping velocity at the time the blade is moving aft, 
causing a Coriolis force forward to reduce the inplane motion. 
The effect of blade droop on the regressive inplane mode is 
shown in Figure 5. 




CO 



-3 -2 -1 

REAL PART OF ROOT 



Figure 4. Locus of Roots - Effect of Pitch Lag 
( a j) coupling, V = 20 KN. 



A schematic of the response of a single blade for this mode 
looking in at the blade tip is shown in the lower left corner of 
Figure 2. The response shown is a stable response. With the 
inplane frequency above the basic flapping frequency, nose 
up blade feathering when the blade is forward, positive ct|, will 
cause the blade to flap up as the blade is going aft. The up 
flapping velocity of the blade generates a Coriolis force which 
reduces the inplane motion. 

Blade Droop. Blade droop is the built in vertical angular 
offset of the blade below the feathering axis (see Reference 5) 
and causes an elastic pitch lag coupling which is similar in effect 
to stabilizing aj coupling. The droop effect though is somewhat 
more effective in stabilizing this particular mode since some 
additional phase lag results in the response of the elastic blade 
feathering which improves the amount of flap induced Coriolis 
damping in the mode. This is accomplished through an increment 




CO 



-3 -2 

REAL PART OF ROOT 



Figure 5. Locus of Roots 
Angle, V = 



- Effect of Blade Droop 
20 KN. 



An additional insight into the effect of droop on the charac- 
teristics of the system is shown in Figure 6. Shown is a predicted 
frequency response of inplane response and of vehicle roll rate 
response due to lateral stick excitation as a function of excitation 
frequency. It is noted that the inplane becomes quite responsive 
at low values of blade droop. It is also noted that even with the 
fairly large separation of the roll mode and inplane mode 
frequency, an influence of blade droop is seen on the roll mode. 
This influence is seen to make an increase in the peak response of 
roll rate at its peak response frequency with increasing blade 
droop. 

Other Parameters. Other parameters such as built in blade 
sweep forward or aft of the feathering axis, 6 3 coupling, control 
system flexibility, stiffness distribution of the blade and hub and 
location of the feather bearings influence the stability of this 
mode. Again, it is pointed out that the influence of each 
parameter depends to a large part on the detail design. However, 
in general, for a stiff inplane hingeless rotor, couplings which 
result in nose down feathering due to lag aft of -the blade add 
damping to the regressive inplane mode. Also, with the inplane 
mode frequency above the flapping mode frequency, couplings 
which act as a negative spring increment to the flapping mode or 
a positive spring increment to the inplane mode are stabilizing to 
the inplane mode. These couplings may, however, influence the 
stability /response characteristics of other modes, in particular the 
roll mode. 



188 



RESPONSE PER INCH OF 
LATERAL STICK 



16 



<o 

tr uj 

-jSS 

O ui 
CC D 



12 





,0 


^ 








r 




" 


s 1° \ 
\ J 


t 








"^3° -^ 





1- 

2 

Ui 

o z 

UJ j 

2 ■ 

< 5 
-j 

Q. in 
2 2 







1 

\ DROOP 
J ANGLE 






1 10 






1 


r 






J&^? 







2 3 

FREQUENCY - CPS 



Figure 6. Effect of Blade Droop on Inplane and Vehicle 
Roll Frequency Response Characteristics, V = 20 KN. 



Coupled Regressive Inplane Body Roll Mode 

Next is considered the coupled regressive inplane bending- 
body roll mode where the frequency of the inplane mode and of 
the roll mode are nearly coalescent. Figure 7 is a complex plane 
plot of a typical coupled regressive inplane-roll mode for a direct 
flapping moment gyro control type system where the inplane to 
roll mode frequency ratio is 1.1. Comparing this figure with 
Figure 2, it is noted that the roll response of the airframe relative 
to the inplane is significantly larger in this mode. In this case, the 
phase relationships between inplane, the rotor pitch and roll, and 
the cyclic blade angle are still in a damping phase for the inplane 
but the rotor roll-airframe roll phasing is such as to provide a 
slight driving to the airframe roll motion. For this particular case, 
the net damping of the regressive inplane mode would be some- 
what reduced. Again, a discussion on the effect of significant 
parameters which influence the characteristics of this mode 
follows. 



Inplane Frequency . Figure 8 shows the influence of inplane 
frequency on coupled regressive inplane bending-roll mode 
damping. Data are shown for a low-speed, 20-knot case and a 
high-speed, 235-knot condition (compound helicopter flight 
mode). It is interesting to note that at low speed, the roll mode 
loses damping due to frequency coalescence whereas at the high- 
speed condition, it is the inplane mode that tends to lose 
stability. 




Figure 7. Complex Plane Plot of Typical Coupled Regressive 
Inplane Roll Mode, V = 1 60 KEAS, w e / w RoU = 1.1. 



235 KNOTS 
20 KNOTS 




OJ 



-3 -2 -1 

REAL PART OF ROOT 



Figure 8. Locus of Roots — Effect of Inplane Frequency. 



189 



Blade Droop. The influence of blade droop, where the 
inplane mode and roll mode frequencies are close, is shown in 
Figure 9. These data show a significant effect of droop on the 
tradeoff of damping between the two modes. It is noted that 
increasing droop has a significant effect in increasing the damping 
of the inplane mode, but an equally significant effect in reducing 
the damping of the roll mode. 

Vehicle Roll Inertia . Figure 1 shows the influence of 
vehicle roll inertia. In the case shown, a small reduction in damp- 
ing of the regressive inplane mode and a significant improvement 
in damping of the roll mode result from increasing the roll 
inertia. A reduction in roll frequency is also seen. Ordinarily, a 
roll mode in the 0.6 to 1 .3 Hz frequency region, with the damp- 
ing sufficiently low, can be subject to pilot-coupled oscillations. 
As can be seen from Figure 10, this problem is avoided by the 
corresponding large increase in damping of the roll mode as the 
frequency decreases into this range with the increasing roll 
inertia. 

Pitch Flap (63) Coupling. The influence of pitch-flap 
coupling on coupled regressive inplane-roll mode stability is 
shown in Figure 1 1 . This figure shows the inplane mode to be 
little affected by decoupling with flap-up, pitch-down coupling 
being slightly stabilizing. The effect on the roll mode is to 
increase its frequency with positive coupling and also to improve 
its damping. The inplane mode frequency is decreased as was 
expected, but the damping increase was not expected. For a case 
(not shown) Where the inplane frequency was considerably above 
the roll mode frequency, the influence of the more positive pitch 
flap coupling was to destabilize the inplane mode slightly with a 
more significant effect of improving the stability of the roll 
mode. 




REGRESSIVE 
INPLANE MODE 

7,000 

i 

13,000 

& 16,000 



16 



14 



■ 12 



10 



REAL PART OF ROOT 



Figure 1 0. Locus of Roots - 
Roll Inertia, V = 



- Effect of Airframe 
20 KN. 



OJ 




(j0 
33 
> 



REAL PART OF ROOT, 1/SEC 

Figure 9. Locus of Roots - Effect of Blade 
Droop Angle, V = 20 KN. 




CJ 



-3 -2 

REAL PART OF ROOT 



Figure 1 1 . Locus of Roots — Effect of Pitch Flap 
(63) Coupling, V = 20 KN. 



190 



Feedback Ratio . Feedback ratio, \ , is the ratio of the 
moment applied to the control gyro by rotor cyclic flapping 
moment or shaft moment to the corresponding rotor shaft 
moment. This parameter is described in detail in Reference 4. 
It is used both to prevent excessive rotor shaft moments while 
the vehicle is in contact with the ground, and to aid in tailoring 
the vehicle handling qualities. This ratio is defined by the 
following equation: 



M, 



gyro 



M, 



shaft 



The influence of this parameter on coupled roll regressive 
inplane mode stability is fundamentally on the roll mode. As 
shown in Figure 1 2, increasing the magnitude of this parameter 
increases the frequency of the roll mode and reduces its damping 
while increasing the damping of the inplane mode. 

Servo Time Constant. For the configuration being discussed, 
the blade cyclic feathering is obtained through irreversible servo 
actuators which are slaved to the control gyro. The lag in the 
servos then causes a lag in the response of the cyclic blade feath- 
ering as commanded by the control gyro. The influence of the 
cyclic servo time constant is shown in Figure 13. It is noted that 
the effect of increasing servo time constant is to reduce the fre- 
quency and damping of the roll mode and the damping of the 
inplane mode. The effect of the cyclic servo time constant 
becomes increasingly important with increasing speed in deter- 
mining the damping of the third type of mode, discussed below. 





-4 -3 -2 -1 

REAL PART OF ROOT 

Figure 1 3. Locus of Roots - Effect of Cyclic Servo 
Time Constant, V = 20 KN. 

Coupled Regressive Inplane-Roll-Rotor Plunge Mode 

This mode is most critical in high-speed flight. It has been 
characterized as a Hop mode because plunging of the rotor disc 
results in a vertical bounce of the airframe. The parameters 
strongly influencing the stability of this mode in a feathering 
moment feedback system are inplane frequency, collective 
control stiffness, pitch-flap coupling, blade product of inertia 
relative to the feathering axis and blade sweep. A typical mode 
shape for this type of mode is shown in Figure 14. It is noted 
that a considerable amount of rotor inplane pitching, rolling and 
plunging, and airframe vertical and rolling motion occurs. The 
mode may become critical with increasing speed if the rotor 
plunge mode and coupled roll inplane mode are allowed to 
approach coalescence. 




1 

>R 




\4-r fx 


R 


) ^s 


Z F 


k 





REAL PART OF ROOT 

Figure 12. Locus of Roots — Effect of Feedback 
Ratio, V= 20 KN. 



Figure 1 4. Typical Mode Shape of Coupled Regressive 

Inplane Bending - Roll — Rotor Plunge 

Mode, V= 180 KEAS. 



191 



This coalescence can be caused by the influence of several 
factors. First, any couplings that cause the rotor plunged mode 
to decrease in frequency with increasing forward speed may 
cause coalescence. This can be due principally to collective con- 
trol system flexibility, blade sweep, adverse pitch-flap coupling 
and blade product of inertia effects; all affecting the collective 
pitch response to vehicle normal acceleration or rotor coning. On 
a first-order basis, a negative or positive aerodynamic spring on 
the collective plunge or coning mode of the rotor can be 
expressed by the following equation: 



9L 
80,, 



*4( K *o 



5, - JTZMxy 



- K /?)' 



3L 



where 
knots. 



3 L approximately doubles between hover and 1 20 



Another source of coalescence or near coalescence can be 
due to the coupled roll-inplane mode increasing in frequency 
with increasing speed. As the lift due to collective blade angle 
increases with speed, so do the aerodynamic derivatives associ- 
ated with the cyclic motions of the rotor disc, and the aero- 
dynamic coupling terms between cyclic and collective rotor disc 
motions. Any kinematic or aeroelastic couplings that phase these 
aerodynamics to act to stiffen the coupled roll inplane mode 
with increasing speed will cause an increase in the frequency of 
this mode with speed. 

Absolute coalescence of these two modes is not necessary 
for instability to occur. Both coupling between the modes and 
frequency proximity are key to stability. Couplings which cause 
cyclic aerodynamic forces or moments due to lift or plunge res- 
ponse of the rotor, which in turn result in cyclic response of the 
rotor disc which cause rotor disc lift or plunge driving forces, can 
be destabilizing. When these couplings are sufficiently strong and 
properly phased, the system will be unstable. 

The significant aerodynamic coupling terms between these 
two modes, which are strongly affected by forward speed, are a 
rolling moment on the rotor due to change in collective blade 
angle, a pitching moment on the rotor due to change in coning of 
the rotor, and lift or plunge aerodynamic loadings due to roll 
velocity of the rotor or to change in longitudinal cyclic blade 
angle. These are direct aerodynamic couplings between these two 
modes. 

Indirect aerodynamic couplings exist through the inplane 
response of the rotor system. This is particularly true in a feath- 
ering moment feedback system, because relatively high inplane 
exciting forces are generated as a result of changes in rotor lift. 
The resulting inplane responses can couple through blade static 
and elastic coning relative to the feather axis and cause perturba- 
tional cyclic feathering responses. These cyclic featherings result 
in aerodynamic forces which can be either stabilizing or 
destabilizing. 

Again Figure 14 shows a typical mode shape or eigenvector 
for this type of coupled roll-regressive inplane bending rotor 
plunge mode. In this case, which happens to be stable but lightly 
damped, the collective feathering is at an amplitude and phase 
with respect to O , collective coning of the rotor, to act as a 



negative aerodynamic spring on the coning mode. Likewise, Q , 
collective blade angle, acts in conjunction with longitudinal 
cyclic blade angle in causing the rotor to pitch up. As can be 
seen from this figure, the rotor pitch response is lagging the 
collective blade angle response by approximately 45 degrees, 
whereas the coning response is actually leading the collective 
blade feathering response by a small phase angle. 

Further examination of Figure 14 shows that the coning 
response, in addition to the rotor pitch response, is also being 
driven by longitudinal cyclic blade angle. It is interesting to note 
that the lateral inplane response is leading the rotor coning res- 
ponse by approximately 90°. Positive lift on the rotor combined 
with lateral cyclic blade angle causes a lateral inplane excitation. 
Positive lift results in an increase in lateral inplane bending to the 
left, which is aft bending on the aft blade and forward bending 
on the forward blade. The fact that the lateral inplane response is 
lagging its excitation by approximately 90° and the response is 
virtually pure regressive indicates that the inplane mode is very 
close to being in resonance. The inplane response, in coupling 
through the feathering axis, is a prime source of the longitudinal 
cyclic blade angle. 

Even though the mode shown in Figure 1 4 is stable, one can 
see the potential for the mode to lose damping, which it does for 
the case shown, with increasing air speed. Lift and rotor disc 
rolling moment due to 6 and longitudinal cyclic both increase 
with air speed, as well as rotor disc pitching moment, due to con- 
ing of the rotor. These aerodynamic terms in conjunction with 
the inplane aerodynamics due to rotor coning are the principal 
coupling terms between rotor disc plunge and coupled roll regres- 
sive inplane response. It is through these terms and the choice of 
rotor/control system parameters that the coupled rotor vehicle 
system can be rriade to have adequate damping at high speed. 

Figure 1 5 shows the effect of pitch flap coupling, blade 
product of inertia, control system collective stiffness, and blade 
sweep which, as indicated earlier, are key parameters in influen- 
cing the stability of this mode. 

Studies are also presented for the stability characteristics of 
this type mode, for the direct flapping moment feedback type 
control system. In this system, one other parameter was intro- 



+1r 



O 
O 

E£ 
Li. 
O 
H 
CC 
< 



12 3 

SWEEP ANGLE-DEGREES 




10 20 30 



K„ 10 3 FT-LB/RAD 



+1f 



-1 
-20 




-10 



10 



v MXY - SLUG-FT* 



-.6 -.4 -.2 

S3 COUPLING, -9/p 



Figure 15. Effect of Parameters on Stability of Coupled 

Regressive Inplane Bending — Roll - Rotor Plunge 

Mode, V= 180 KEAS. 



192 



duced which has a significant effect on this mode. The parameter 
is the time constant or frequency response characteristic of the 
main power cyclic actuators. It is through these actuators that 
the control gyro commands cyclic blade feathering. As indicated 
earlier, a lag in the servo response results in a lag in the cyclic 
blade, feathering, which can have an adverse effect on the stabil- 
ity of the hop mode. Inasmuch as the hingeless rotor depends on 
corrective control such as by stabilizing gyro to prevent pitchup 
at high speed, it is recognized that lag in the corrective control 
may lead to dynamic instability. This influence or effect is shown 
in Figure 16. 

Figure 1 6 shows also the effect of 63 coupling as well as 
collective control system stiffness and inplane frequency. 



REGRESSIVE INPLANE MODE 

ROLL MODE 




1 
2 

S 3 

o 



o 

z 



SERVO TIME CONSTANT 
.01 .02 .03 .04 .OS 



S3 COUPLING, -0/(3 
-0.1 0.1 0.2 0.3 





- — 


— 








1 

2 
3 


k, 























































< 
a 



K(J -% NOMINAL 
20 40 60 80 100 



\\ 



INPLANE FREQUENCY-P 
1.2 1.3 1.4 1.5 1.6 



■» mm •■" 



Figure 1 6. Effect of Parameters on Rotor Vehicle High Speed 

Dynamic Stability — Direct Flapping Moment Feedback 

Control System, V = 280 KEAS. 

Experimental and Analytical Comparison 

The experimental and analytical comparison is based upon 
aata obtained during the development of the AH-56A. Early in 
the development of the AH-56A, a vehicle equipped with an 
experimental rotor system in which the blades had been modified 
by adding torsional doublers encountered a dynamic Hop 
phenomenon. The principal effect of the torsional doublers on 
this mode was to lower the inplane frequency and cause it to 
become more critically coupled with the rotor plunge/body roll 
mode. An analytical study was undertaken to define this 
phenomenon which extended the coupled rotor body linear 
analysis method available at that time and led to the 
development of the linear math model discussed earlier. 

Figure 17 shows a comparison of the experimental and 
analytical data obtained for the vehicle configuration which 
initially encountered the Hop or coupled roll-regressive inplane 
bending-rotor plunge mode phenomenon. Additionally, the 
following table summarizes the normalized roll rate, chord 
moment and collective control load comparison obtained for this 
condition. In both the experimental and analytical data, the 
responses are due to a roll doublet excitation and are normalized 
on vehicle e.g. vertical acceleration. 







95% N R 
Test Analysis 


100% N R 
Test Analysis 


Roll Rate, 
deg/sec/g 




20 19.1 


15.5 ■ 9.2 


Collective Control 
Load, Ib/g 




2700 2820 


2370 2610 


Inplane Moment, 
in. Ib/g 




424K 420K 


770K 670K 


Frequency Ratio, co/n 


0.52 0.51 


0.54 0.54 


Speed, KEAS 




190 180 


178 180 






- ANALYSIS 






O 


97% N R "I 






Q 


95% N R V TEST DATA 




A 


94% N R J 





c 

3 



C 



o 





i100% N 


% 












^95% N R 













120 160 200 240 

FORWARD SPEED - KNOTS 



280 



Figure 1 7. Comparison of Theory and Test Damping vs Speed 

For Initial Encounter With Hop on Early AH-56A 

Development Configuration. 



The loss in damping was caused by a coalescing of the rotor 
body roll mode, the inplane mode (with both modes exciting 
blade cyclic feathering), and the rotor plunge mode. The terms 
discussed in the equation for 9L/gg previously given were such as 
to cause the rotor coning or plunge mode to decrease in fre- 
quency with increasing speed. In hover, the frequency of this 
mode was close to IP. With increasing speed, the frequency drop- 
ped into the 0.5 to 0.6P frequency range in the 200-knot speed 
regime and coalesced with the lower-frequency body roll, regres- 
sive inplane modes. This resulted in the observed reduction in 
damping of the Hop mode with increasing forward speed. 

A modification was made to the system which included 
approximately doubling the collective control system stiffness, 
reducing the pitch-flap coupling from a value of 0.27 to a value 
of 0.05 at a collective blade angle of 5 degrees, increasing the 
blade sweep from 2.5 to 4 degrees sweep forward, and reducing 
the inplane frequency from approximately 1.55P to 1.4P. The 



193 



reduction in pitch-flap coupling and the increase in collective 
control system stiffness were done specifcally to eliminate the 
Hop phenomenon within the flight envelope. The increase in 
sweep and reduction in inplane frequency were done to improve 
certain handling quality characteristics. These changes resulted in 
the frequency of the collective coning mode remaining virtually 
constant with increasing forward speed. The changes also resulted 
in the coupled roll regressive inplane mode remaining at nearly a 
constant frequency with speed. The resultant effect was to 
increase significantly the speed at which the predicted coupling 
between these modes became critical. 

Figure 1 8 shows a comparison of the predicted damping of 
the coupled roll-regressive inplane bending mode with test results 
for this modified configuration as a function of speed. This figure 
indicates fairly good agreement between the measured and pre- 
dicted values. Figure 1 9 shows a comparison between the 
predicted and measured chord-bending response due to a lateral 
stick doublet at 170 knots. As can be seen, good agreement 
between the two responses was obtained. 



The rotor system was then modified to increase the blade 
droop from 2°20' to 3°10' (Reference 5). this configuration 
change had little effect on the high-speed coupled roll-regressive 
inplane mode stability characteristics, and the vehicle was sub- 
sequently flown to 240 knots' true airspeed with no indication of 
a high-speed dynamic stability problem. 

This latter configuration change however, did, lower the' 
damping of the coupled roll regressive inplane mode in hover and 
low-speed flight because of the increase in blade droop. The 
mode was characterized by roll oscillation and inplane response 
due to pilot lateral stick inputs. The frequency of the mode was 
approximately 1 Hz. This, coupled with the roll oscillation of the 
airframe, made the mode susceptible to pilot coupled oscillation. 

Figure 20 shows a comparison of the experimentally 
determined and predicted roots of the coupled roll-regressive 
inplane mode for the two different blade-droop configurations. 
Again, fairly good agreement is seen between experimental and 
analytical results. 



ANALYSIS 



• EXPERIMENT 



< 



P 1 



O 

1- 



o 
> 



— m ( 



120 160 200 240 

FORWARD SPEED-KNOTS 



280 



Figure 18. Comparison of Theory and Test Damping vs Speed 
For Modified Rotor — Control Configuration. 

RT 

PO 
top 
-JO • 

< w 5 

< Q ANALYSIS 

---EXPERIMENT 





A major revision was then made to the control system 
which replaced the feathering-moment feedback system with a 
direct flapping-moment feedback system. This change necessi- 
tated placing the main cyclic power actuators between the 
control gyro and blade feathering instead of between the pilot 
and the gyro. 



2° 20' DROOP 

3° 10' DROOP 

OPEN SYMBOLS - ANALYSIS 

CLOSED SYMBOLS - TEST 




10 u> 



ELAPSED TIME - SEC 

Figure 1 9. Comparison of Experimental and Analytical 

Transient Chord Bending Response Due to Lateral 

Stick Doublet, V = 175 KN. 



REAL PART OF ROOT 

Figure 20. Effect of Droop Angle on Low Speed Roll Mode 

Stability - Comparison of Experiment and Analysis, 

V = 20 KN. 



194 



The rotor and control system parameters were selected to 
provide a system that was completely free of either the undesir- 
able Hop or roll mode characteristics discussed earlier. The 
various parameters, which were established to be critical by 
extensive parametric studies, using the linear analysis method 
adapted to computer graphics, were established and controlled 
very carefully. These parameters included both cyclic and collec- 
tive pitch flap coupling, inplane frequency, main cyclic power 
actuator time constant, gyro to blade-feathering gear ratio and 
phasing, blade sweep and droop, and shaft-moment to gyro- 
moment feedback ratio. 

The initial configuration, when tested on the whirl tower, 
was determined to have met all criteria except that the inplane 
frequency was below the criteria value by about 0.05P, or 0.21 
Hz. Some limited flight testing was performed with this configur- 
ation to validate the criterion, after which the final configura- 
tion, conforming to the original criteria, was reached by 
removing 6.8 pounds tip weight from each blade. The results 
with both configurations, are discussed in the following. 



a 

z 

i m 

o -I 



-100K l \^y ^ 



100K 
100K 



rf 1- 
x 5 -i° 0K 

a! 100K 

9S 






95% NR 



100% NR 



105% NR 



Figure 21 shows the effect of predicted rotor vehicle 
responses as a function of rotor speed for a lg, 1 60-knot flight 
condition with the degraded inplane frequency. The excitation in 
each case is a lateral stick doublet at 1.5 Hz which is the tech- 
nique used in flight test for exciting coupled rotor-body dynamic 
modes to determine their stability characteristics. 



It is noted that with increasing rotor speed, the damping of 
all responses decreases, and the magnitude of the pitch response 
of the rotor disc in the mode increases. This was noted by the 
pilot as a characteristic of the mode in that, with similar excita- 
tions, the rotor disc tip path oscillations would be imperceptible 
at lower rotor speeds but would become increasingly responsive 
at higher rotor speeds. 



Figure 22 shows a comparison between the calculated roots 
and the experimentally determined damping and frequency for 
this configuration at 1 60 knot airspeed. The actual vehicle res- 
ponses contain a varying mix of the two roots, increasing in 
inplane content with increasing rpm. 



-H ANALYSIS 

• EXPERIMENT - FROM RESPONSE OF ROLL RATE 




2 3 

ELAPSED TIME - SEC 



3 -2 -1 

REAL PART OF ROOT 



Figure 21 . Rotor Speed Effect on Transient Response Due Figure 22. 

to Lateral Stick Doublet, V= 160KEAS. Analysis - 



Locus of Roots — Comparison of Test and 
Reduced Inplane Frequency, V = 160 KN. 



195 



Figure 23 shows the predicted effect of increasing the 
inplane frequency by removal of tip weight on the frequency and 
damping of the coupled roll-regressive inplane modes. Figure 24 
shows the corresponding predicted transient response at 105 
percent of normal rotor speed with the tip weight removed. The 
comparison in roots shown on Figure 23 and the comparison of 
the 1 05 percent rpm transient response in Figure 24 with the 95 
and 105 percent rpm cases in Figure 22 show a significant 
improvement in the damping and transient response character- 
istics at rotor overspeed for the configuration with the tip weight 
removed to give the desired inplane frequency placement. 

For the final configuration, Figure 25 shows a comparison 
of the measured and predicted damping as a function of forward 
speed. The data show good agreement in measured and predicted 
damping levels from hover through transition and in higher-speed 
flight. Experimental data on damping of the regressive inplane 
mode consist of only one point because even though the mode 
was not excessively damped, it was extremely difficult to excite 
by the pilot with lateral stick doublet type excitations to amp- 
litudes sufficiently large to obtain a reliable determination of its 
stability. This final configuration was tested over a very large 
flight envelope covering speeds to 220 knots true airspeed and 
maneuvering load factors from -0.2g to 2.6g in the 1 80 to 
200-knot true airspeed flight regime. The pilot reported "excel- 
lent" to "deadbeat" damping and minimal responses to air 
turbulence in high-speed flight. 

X TIP WEIGHT REMOVED 
O TIP WEIGHT IN 





280 




REAL PART OF ROOT 

Figure 23. Locus of Roots - Effect of Tip Weight, 
V=160KEAS. 



12 3 

ELAPSED TIME - SEC 

Figure 24. Transient Response Due to Lateral Stick Doublet 
Tip Weight Removed, V = 160 KEAS. 

ANALYSIS 

O EXPERIMENT 
FORWARD SPEED - KNOTS 
80 120 160 200 240 



n 1 



a 

z 
a. 
S 
< 
Q 



Figure 25. Damping vs Forward Speed — Comparison of Test 
and Analysis For Final AH-56A (AMCS) Configuration. 



Conclusions 

Several types of modes can exist in a stiff inplane hingeless 
rotor which involve coupling with the regressive inplane mode. 
These include phenomena where the inplane mode is not well 
coupled with the rest of the system, phenomena where the 
inplane mode and body roll mode are the primary participants, 
and even phenomena where the rotor plunge mode is heavily 
involved in the total system dynamic behavior. These phenom- 
ena, particularly the first two types, are not limited to any 
particular flight regime but can be critical in either stability or 
response in either low- or high-speed flight or can exhibit char- 
acteristics which are virtually independent of speed. Each of the 
modes is treatable by analysis, and certain parameters such as 
blade droop, control system stiffness, pitch-flap and pitch-lag 
couplings, blade sweep, blade product of inertia, inplane fre- 
quency and control system parameters are influential in control- 
ling the stability and response characteristics of these modes. 
Additionally, a totally satisfactory system can be achieved 
without recourse to auxiliary damping devices. 



196 



References 

1. Burkam, J.E.; Miao, Wen-Liu; EXPLORATION OF 
AEROELASTIC STABILITY BOUNDARIES WITH A 
SOFT-IN-PLANE HINGELESS-ROTOR MODEL, Journal 
of the American Helicopter Society , Volume 1 7, Number 4, 
October 1972 

2. Huber, H.; EFFECT OF TORSION-FLAP-LAG COUPLING 
ON HINGELESS ROTOR STABILITY, Preprint No. 731, 
Presented at the 29th Annual National Forum, Washington, 
D.C., May 1973 

3. Donham, R.E.; Cardinal, S.V.; Sachs, LB.; GROUND AND 
AIR RESONANCE CHARACTERISTICS OF A SOFT 
IN-INPLANE RIGID-ROTOR SYSTEM, Journal of the 
American Helicopter Society , Volume 14, Number 4, 
October 1969 

4. Potthast, A.J.; Blaha, J.T.; HANDLING QUALITIES 
COMPARISON OF TWO HINGELESS ROTOR CONTROL 
SYSTEM DESIGNS, Preprint No. 741, Presented at the 
AHS 29th Annual National Forum, Washington, D.C., May 
1973 

5. Anderson, W.D.; INVESTIGATION OF REACTIONLESS 
MODE STABILITY CHARACTERISTICS OF A STIFF 
INPLANE HINGELESS ROTOR SYSTEM, Preprint No. 
734, Presented at the AHS 29th Annual National Forum, 
Washington, D.C., May 1973 

6. Chou, P.C.; PITCH-LAG INSTABILITY OF HELICOPTER 
ROTORS, Journal of the American Helicopter Society , 
Volume 3, Number 1, July 1958. 



197 



HUB MOMENT SPRINGS ON TWO-BLADED TEETERING ROTORS 

Walter Sonneborn 
Grp. Eng. Mechanical Systems Analysis 

Jing Yen 

Grp. Eng. VTOL Technology 

Bell Helicopter Company 

Fort Worth, Texas 



MAGNITUDE OF HUB MOMENT 



Two-bladed teetering rotors with 
elastic flapping hinge restraint are 
shown to be suitable for zero-g flight. 
The alternating moment component intro- 
duced into the fuselage by the hinge 
spring can be balanced about the aircraft 
center of gravity by alternating hub 
shears. Such shears can be produced in 
proper magnitude, frequency, and phase 
by additional underslinging of the hub 
and by judicious choice of the location 
of the first inplane cantilevered natural 
frequency. Trends of theoretical results 
agree with test results from a small scale 
model and a modified OH-58A helicopter. 



Centrally hinged rotors have traditionally 
relied upon thrust vector tilt for gener- 
ating control moments about the helicopter 
eg. All present production two-bladed 
rotors have central teetering (flapping) 
hinges. Such rotors, without hub restraint, 
have no control power in zero-g flight. 
Recent military specifications for 
transport and attack helicopters call for 
the ability to sustain zero-g flight for 
several seconds. Helicopter control under 
this condition of no rotor thrust requires 
hub moments which in two-bladed rotors can 
be generated by springs restraining the 
flapping hinge. The resulting flapping- 
dependent hub moment, when observed in the 
fixed system, has a mean value in the 
direction of and proportional to the 
maximum flapping relative to the shaft. 
A 2/rev oscillatory moment with an ampli- 
tude equal to this mean value results in 
both the fore and aft and lateral direc- 
tions. This paper discusses methods for 
producing 2/rev hub shears for balancing 
the oscillatory component of the spring 
moment about the helicopter center of 
gravity. Practical magnitudes of hub 
moments are defined by minimum control 
power requirements for zero-g flight, and 
maximum values are limited by a variety of 
factors. Test results from a 1/12- scale 
Froude model and flight test results from 
an OH-58A helicopter with variable hub 
restraint are presented. 



The basic benefits of hub moment are 
better aircraft rate damping and positive 
control power in zero-g flight. Minimum 
hub moment requirements have been investi- 
gated by analysis and testing of an OH-58A 
helicopter. Zero-g flight was demonstra- 
ted with this helicopter using only stiff 
elastomeric bearings in the see- saw hinge 
for hub restraint increasing the 1-g 
control power by 10%. Figure 1 shows a 
record of the maneuver. Only small 



oo 
§ 

1-1 H 

Eh ci 

t-H 

00 B^ 

O 

Ph - 

►J § 

q fa 
ei 

o 

o 



d 
o 

H 
O 

3 



100 



80 



60 



i) Q Q 



Q 

1-3 



o 







00 



< 



2 

( 

i 
o 

20' 



LONGITUDINAL 
9 <t> 



[| □ tl 



LATERAL 

a A a 



(I 



o 












~(> J 2.9 SEC < .25 g 1 © 
°t> « t\ <a /» ° 















I* 



-20 



ii_©ja. 



ROLL. 



-B-f- 



O 

PITCH 



12 3 4 5 

•TIME - SECONDS 

FIGURE 1. Model OH- 58 A pushover with 
elastomeric flapping bearings. 



199 



lateral inputs were made during the maneu- 
ver, and the roll angle did not exceed +5 
degrees. The lateral SCAS was engaged and 
contributed significantly to roll stabili- 
zation. This test and analysis indicated 
that some 25% of the 1-g control power is 
adequate for zero-g flight. The OH- 58 A 
helicopter was subsequently fitted with 
ground- adjustable hub torsion springs 
which added 23% or 37% to the 1-g control 
power (see Figure 2). The pilot's reac- 
tions were favorable with regard to the 
lower spring value, but the stiff er spring 
made roll control power excessive and also 
increased the gust sensitivity noticeably. 



LENGTH FOR 1 
ll32 FT-LB/DEGI 



LENGTH FOR 
1210 FT-LB/DEG 



SLIP RING 




EFFECT OF UNDERSLINGING 
AND CHORDWISE FREQUENCY 

The effects of underslinging on 2/rev 
hub shears and of hub restraint on 2/rev 
hub moments are shown by an analysis of 
the simple rigid body model shown in 
Figure 3. The kinematics of an underslung, 




<t MAST 




VIEW 
FROM 
TOP . 



2/REV CIRCULAR PATH 
OF ROTOR eg 



FIGURE 3. Simple rigid body math model 
of two-bladed rotor. 

flapping, two-bladed rotor with a teeter- 
ing hinge cause the center of mass of this 
rotor to travel in a circular path at 
2/rev if the mast does not oscillate about 
point A (this assumption will be partly 
justified below). The resulting centrif- 
ugal forces introduce hub shears S for 
fore and aft (F/A) flapping: 

2 
Sp/ A =2ai<u u m cos2<ut . (la) 



S, . 1 ,=2a, <u u m sin2&)t (lb) 



FIGURE 2. Experimental Model 640 rotor 
with ground-adjustable torsion tube 
flapping restraint. 

Other considerations limiting the hub 
restraint of two-bladed rotors are: 

- Fuselage vibration caused by 
oscillatory hub moment at 
2/rev. 

- Increase in beamwise bending 
of flexure. 

- A weight penalty of about 90 
pounds per 1000 ft-lb/deg of 
effective control moment. 

- Instability of the coupled 
pylon/rotor system at extreme 
spring stiffnesses. 1 

These tend to discourage the designer from 
introducing significantly more hub moment 
than that equivalent to 25% of the 1-g 
control power from thrust vector tilt. 



where 

a-, = F/A flapping angle 

m = rotor mass 
cot = rotor azimuth 
The moments M from the hub spring are: 
M F / A =- O.SajKjjQ + cos2«t) (lc) 



M LAT =-0.5a,Kr, sin2eut 



(Id) 



Now taking moments about point A below 
the rotor (M. = M + Sh) , it is evident 
that all oscillatory components can be 
cancelled in both the fore and aft and 
lateral direction if 



— a l K H = (2a,&> u m) h 



(2a) 



200 



and du _^ 



dK, 



H 



4w hm 



(2b) 



When this condition is met a rigid 
mast will not oscillate, but merely 
experience a steady tilt, the amount of 
which is determined by the mean value of 
: the hub spring moment and the stiffness 
Kp of the pylon spring. 

The dynamic analysis was extended to 
include the effects of the first inplane 
mode and the rotor coning mode. Also 
included were aerodynamic calculations at 
the 3/4 blade radius and a modal repre- 
sentation of the pylon support system. 
The set of five differential equations 
\was solved on a hybrid computer (Bell 
J program ARHB2). The solution" showed that 
,the location of the first inplane canti- 



\levered blade natural frequency ok. has 
k prounounced effect on hub shears, 
figure 4 shows how the requirement for 
uhderslinging u of the eg changes with 



Since the loads induced by spring 
restraint are not in phase with the loads 
of the unrestrained rotor, small amounts 
of hub restraint can reduce chordloads 
(see Figure 5). In general, the spring 
induced loads are small when compared 
with the + 7000- inch- pounds loads occur- 
ring in tEe unrestrained rotor at V v in 
level flight. 



"H 



K H = 132 FT-LB/DEG 

VECTORS INDICATE 
MAGNITUDE AND 
PHASE OF MAX. 
FWD. BLADE BENDING 



w = 0.94/REV 
K H = 



= 210 FT-LB 




FWD FLAPPING 



Rigid Blade Theory 




»-<y; 



*c 



FIGURE 4. Underslinging requirement 
versus blade cantilevered first inplane 
natural frequency. 

The blades act like a dynamic absorber 
whenever alternating hub spring moments 
begin to induce pylon motion. They retain 
this absorber function over a large range 
of blade frequencies because of the 
relatively large absorber (blade) mass. 
The chordwise bending moments induced at 
the blade root in this manner have been 
computed for the experimental Model 640 
rotor, (see Figure 5. The pylon param- 
eters used are representative of an OH- 58 A 
helicopter). This rotor has a cantilever- 
ed blade natural frequency of 0.94/rev. 
(This frequency is raised to 1.4/rev in 
the coupled rotor/pylon system. ) 



FIGURE 5. Blade root chord loads as a 
function of hub moment spring rate. 

The ideal moment balance about point 
A, as suggested by equation (2a) and the 
above discussion, is actually not fully 
achieved. When the underslinging on a 
flapping hub- restrained rotor is varied, 
the complete calculations show a residual 
pylon oscillation remaining and the phase 
of the pylon response changing in a contin- 
uous manner (see Figure 6). 'The reason 



FWD FLAPPING = 6 DEG 



=132 FT-LB/DEG 



20.- 



§ 



1 
m 

o w 
< to 






g +9<£ 

6-" 

1-1 W 

04 w o 

Bs .go 01 

>* 

04 






132 FT-LB/DEG\ 
0.94 




132 FT-LB/DEG 



.10 



.15 



.20 



UNDERSLINGING, u - feet 



FIGURE 6. Pylon response versus under- 
slinging. 



201 



for this is that the airloads of the free- 
flapping rotor are slightly modified when 
the airload moment due to hub spring is 
considered. Figure 7 shows the lift and 
drag increments on a lifting rotor 
resulting from this, and it is evident 
that an inplane shear 90 degrees out of 
phase from the desired shear results. 
(If the rotor were not producing any 



AL 



%0K R / .75R 


/^~ ^^^\ 


4 AL 


/ \ 


1 + 1 


L Ca 




Wad adW 


VIEW 
SIDE, 


FROM 


VIEW FROM 
TOP 



FIGURE 7. Change AL in blade lift tp 
balance hub moment, and resulting drag 
components. 

net lift then both blades would experi- 
ence a drag increase, leaving no net hub 
shear) . Figure 8 shows records of 
computed pylon responses with and without 
consideration of the inplane shears due 
to airloads. 

AIRLOADS INCLUDED AIRLOADS OMITTED 



IN HUB SHEARS 



FROM HUB SHEARS 



^ = 

u = BASELINE 



"H 



210 FT- LB 
DEG 

u = OPTIMUM 







f 












:::; 






.'.". 








;•■: 


















■■: 




— i 




- 










[ 




I!:: 


















:<£ 


-r. 










< 










::■: 


<£ 


s ; 










« 






•':'.': 


'.i.: 




! >L 














K 


::'.': 


j::i 




















'■•':. 




cr 











: . -/ 


r 












































.: . 


■■■■■if 


i 






















■ \ 


U- ' 






















:-:^ 


i ■' ■ 














— -u 







W^. = 1.4/REV 



FIGURE 8. Comparison of hub acceleration 
responses. 



EXPERIMENTAL RESULTS 

Model Tests 

A test was conducted on a 1/12- scale/ 
Froude model. Figure 9 shows the appara-/ 
tus also shown schematically in Figure 3./ 
The model was operated at a fixed cyclic 

and collective pitch and at 900 rpm. Th€ 
mahogany blades were heavier than typical 
for helicopter design practice, hence 
little coning took place, and even at the 
smallest possible amount of under slingin$ 
(determined by the bearing diameter of 
the see- saw hinge) some hub spring was 
required for smooth running of the mode] 
Accelerometers above and below the plane's 
of the pylon gimbal support detected thijs 
smooth condition. The amplitude of the/ 
2/rev acceleration was a function of thg 
hub spring rate and the amount of under?- 
slinging, as shown in Figure 10. The data 
scatter near the equilibrium position is 
indicative of the residual ( oscillation. 
The phase change of the pylon response 
occurred in the. gradual manner found in 
the analysis. However, it was noted that 
additional underslinging was only about 
half as effective as anticipated in 
balancing hub moments. (The cantileVered 
blade frequency is 1.5/rev). A partial 
explanation is in the different mast bend- 
ing due to a shear and a moment (see 
Figure 11a). 

■■■■■■■HMHHPHP 

■■aDJUSTA! i . j\l 

if 






feT""*" 




eromete: 



FIGURE 9. (Model (1/12 Froude Scale) with 
variable underslinging and hub restraint. 



202 



w . 

g 

P-i 



UNDERSLINGING -IN. 
(FULL SCALE EQUIVALENT) 




KjjXlO J FT-LB/DEG 
(FULL SCALE VALUES) 



. FIGURE 10. Model pylon accelerations 
•versus hub spring. 

The mast deflection of the tip of the 
mast under an applied moment is greater 
than that resulting from the balancing 
shear force by an amount equal to 20% of 
the radius of the circle described by the 
eg of the underslung rotor. In addition, 
the excursions of the rotor's center of 
mass are reduced when part of the total 
disk flapping occurs in blade flexures. 
The magnitude of this effect is dependent 
on rotor coning (see Figure lib). Both 
of the above effects were omitted from 
the analysis. 




SPRING 
MOMENT 



HUB 
SHEAR 




(a) Moment Diagrams 



M J ' 

"blade » 




(b) Shift of Rotor C.G. Due 
to Flexure Bending 



FIGURE 11. Factors reducing the effect of 
unde r s 1 ing ing . 



Flight Test Results 

The OH-58A helicopter shown in Figures 
2 and 12 was flown at 3250 pounds gross 
weight with hub restraints of 0, 132, and 




FIGURE 12. Modified 0H-58A helicopter 
with restrained flapping hinge. 

210 foot pounds per degree. The eg was 
varied from station 106.1 to 111.8 
Figure 13 shows the 2/rev vibration 
measurements at the pilot's seat. 



.3 



60 
+1 

« 

> 

> 

w 

PS 



2- 



,1- 



G.W. 



O 

a 

A 





= 3250 
e.g. 


LB 
K H -FT-LB./DEG 


106.1 
111.8 
106.1 
111.8 






132 

132 





a 
ft 



a 







60 80 100 
AIRSPEED - KIAS 



120 



FIGURE 13. Vertical vibration of pilot's 
seat. 

The influence of hub restraint (132 ft- lb/ 
deg) is negligible compared with the 
increase in vibrations with forward speed. 
There are several reasons for this. The 
helicopter has a focused pylon isolation 
system^ which is effective for isolating 
inplane hub shears and hub moments. 
(There is no vertical isolation) . In 
addition to the moment and horizontal 
shear isolation, the predicted dynamic 
absorber effect appears to take place. 
The first cantilevered inplane natural 
frequency of the test blade is located 
at 0.94/rev and it was shown in Figure 4 
that for this frequency placement nearly 
no additional underslinging is required. 
The dynamic absorber effect of the blade 
is reflected in the oscillatory chordwise 



203 



loads. The measured change in these loads 
in hover as a function of hub restraint 
and flapping is similar to the computed 
values : 



K H 

FT- LB 

DEG 


AFT 


FLAPPING 


OSG. CHORD MOMENT @ 

STA 7. 8- IN. LB 
MEASURED COMPUTED 







4.1° 


2900 


2175 


132 




3.4° 


900 


2000 


210 




2.6° 


3000 


3000 



No computations were made for the forward 
flight case. It is helpful, however, to 
compare the oscillatory moments introduced 
by the hub spring with the moments about 
the aircraft eg due to the 1/rev hub 
shears from airloads. These shears are 
estimated from the modal shear coeffic- 
ient of the first inplane cyclic mode. 
The + 7000- inch-pound chordwise moment 
at V„ corresponds to a hub shear of + 80 
pounas per blade, which is equivalent to 
a 2/rev moment about the helicopter eg of 
+ 450 ft- lb. The maximum oscillatory 
spring moment was only 50% of this value 
when flapping reached 3.3 degrees in 
hover at the forward eg and with the 132 
ft-lb/deg hub spring. (This amount of 
flapping is usually not exceeded in normal 
maneuvers). This comparison shows that 
the vibratory excitation introduced by 
the hub spring is relatively small to 
begin with. 

The pylon isolation system and the 
placement of the blade first cantilevered 
inplant frequency near 1/rev made 
additional underslinging unnecessary in 
this aircraft for vibration isolation. 
(The baseline underslinging for the 
experimental Model 640 rotor was 2.375 
inches). 

CONCLUSION 

(1) Two bladed rotors with hub restraint 
are suitable for zero-g flight. 

(2) Hub restraint which added some 27% 
to the one-g control power of an 
OH-58A helicopter with a Bell Model 
640 rotor caused a negligible 
increase in 2/rev vibrations during 
hover and level flight. 

(3) The 2/rev oscillatory moment compon- 
ent due to hub restraint in a two- 
bladed rotor can be balanced about 

a point below the rotor hub by 
additional rotor underslinging. The 
amount of this underslinging depends 
on the location of the natural 
frequency of the first cantilevered 
inplane blade mode. 



REFERENCES 

Gladwell. G.M.L. and Stammers, C. W./, 
On the Stability of an Unsvmmetricaj 
Rigid Rotor Supported in Unsymmetrical 
Bearings , Journal of Sound and 
Vibrations, 3,(3), (1966), pp. 221-' 
232. 

Balke, R. W. , Development of the 
Kinematic Focal Pylon Isolation 
System for Helicopter Rotors . The ', 
Shock and Vibration Bulletin, 38,(3), 
November (1968), p. 263. 



204 



OPEN AND CLOSED LOOP STABILITY OF HINGELESS ROTOR 
HELICOPTER AIR AND GROUND RESONANCE 



Maurice I. Young* 



David J. Bailey**, f and Murray s- 
The University of Delaware 
Newark, Delaware 



Hirschbein** 



Abstract 

The air and ground resonance insta- 
bilities of hingeless rotor helicopters 
are examined on a relatively broad para- 
metric basis including the effects of 
blade tuning, virtual hinge locations, and 
blade hysteresis damping, as well as size 
and scale effects in the gross weight 
range from 5,000 to 48,000 pounds. A spe- 
cial case of a 72,000 pound helicopter air 
resonance instability is also included. 
An evolutionary approach to closed loop 
stabilization of both the air and ground 
resonance instabilities is considered by 
utilizing a conventional helicopter swash- 
plate-blade cyclic pitch control system in 
conjunction with roll, roll rate, pitch 
and pitch rate sensing and control action. 
The study shows that nominal to moderate 
and readily achieved levels of blade in- 
ternal hysteresis damping in conjunction 
with a variety of tuning and/or feedback 
conditions are highly effective in dealing 
with these instabilities. Tip weights and 
reductions in pre-coning angles are also 
shown to be effective means for improving 
the air resonance instability. 

Notation 

C = landing gear equivalent viscous 

damping coefficient, lb/ft/sec 
C = pneumatic shock strut viscous 

damping coefficient, lb/ft/sec 
C. = tire viscous damping coeffi- 
cient, lb/ft/sec 
CG = helicopter center of gravity 
I = moment of inertia about x axis, 

x slug-ft2 
K = landing gear equivalent spring 

e rate, lb/ft 
K = non- linear, pneumatic shock 

s strut spring rate, lb/ft 
K = tire spring rate, lb/ft 

M = mass of helicopter 

M.. = control moment acting in later- 
1 al swashplate equation of mo- 

tion, ft- lb 

Presented at the AHS/NASA-Ames 
Specialists' Meeting on Rotorcraft Dyna- 
mics, February 13-15, 1.974. 

Acknowledgement is made of the support of 

the U.S. Army Research Office, 

Durham, N. C. 

under Grant DA-ARO-D-1247G112. 

•Professor of Mechanical and Aerospace 
Engineering. **Graduate student and re- 
search assistant. tCurrently U.S. Army 
Transportation Engineering Agency, Fort 
Eustis, Virginia. 



M„ 



N 

T/W 

XYZ 

db 

e, 



t 
x,y,z 



V a 2 
«1' «2 



8'k 



n j 



a2 



JM 



JU 



control moment acting in longi- 
tudinal swashplate equation of 
motion, ft- lb 
number of blades 
thrust to weight ratio 
inertial coordinate system 
decibels 

offset of virtual flapping 
hinge , ft 

offset of virtual lead-lag 
hinge , ft 

distance between center of mass 
of helicopter and coordinate 
system axis, ft 

lateral and roll coupling para- 
meter 

helicopter longitudinal, lateral 
and vertical displacements, ft 
helicopter pitch and roll angu- 
lar displacements, rad 
helicopter pitch and roll rate, 
rad/sec 

flapping angular displacement 
of kth blade, rad 
lead-lag angular displacement 
of k th blade, rad 
logarithmic decrement 
non-dimensionalized (by rotor 
radius) displacement of virtual 
flapping hinge from rotor cen- 
ter of rotation, ft/ft 
generalized fuselage and rotor 
system degrees of freedom 
constrained swashplate-blade 
pitch degrees of freedom 
percent of uncoupled critical 
roll damping 

percent of uncoupled blade lead- 
lag damping 

azimuthal coordinate of the kth 
blade , rad 

out-of-plane or flapping fre- 
quency ratio, cycles/revolution 
in-plane or lead-lag frequency 
ratio, cycles/revolution 

Introduction 



In recent years an intensive re- 
search and development effort within gov- 
ernment and industry has focused on hinge- 
less rotor helicopters with a view towards 
mechanical simplification, improved flying 
qualities and greater aerodynamic clean- 
ness. The approach being employed capi- 
talizes on modern structural materials and 
technology which, in principle, permit the 
hingeless rotor blades to flap and lead- 
lag by flexing elastically, rather than by 
the use of mechanical hinges. In order to 
keep cyclic bending fatigue stress and 



205 



blade weight within bounds, the in-plane 
or lead-lag hingeless blade fundamental 
natural frequency ratio, as a practical 
matter, inevitably falls within the range 
.6-. 9 cycles per revolution, although fre- 
quency ratios as small as .5 or as great 
as 1.2 are possible. As a consequence of 
this .6-. 9 range of frequency ratios, both 
ground and air resonance instabilities can 
still occur which stem from this frequency 
ratio being less than unity. 

There arises the added concern that 
slight amounts of internal blade structural 
damping to be expected in hingeless rotor 
blades can cause such instabilities to be 
much more severe and difficult to control 
than in an articulated rotor case, where 
mechanical lead-lag dampers would be a 
standard design feature. On the other 
hand, the elastic flapping of hingeless ro- 
tor blades and the presence of large blade 
structural moments which are aeroelasti- 
cally coupled to the fuselage oscillations 
both in hovering and on its landing gear, 
and the aforementioned relatively high fre- 
quency ratio of hingeless blade lead-lag 
oscillations compared to those of conven- 
tional articulated rotors (.2-. 4 cycles 
per revolution) , present the favorable 
possibility of significant alterations in 
the ground and air resonance stability 
characteristics. This is in contrast to 
centrally hinged, articulated rotors, 
where flapping motion would be expected to 
have negligible effect on such instabili- 
ties. Several recent investigations! #2,3 
have contributed to increased understand- 
ing of hingeless rotor helicopter ground 
and air resonance characteristics, but in 
each case were directed principally at de- 
sign and development of a particular ma- 
chine with its unique size, structural and 
operational characteristics, rather than 
at broad development of parametric trends 
and general principles, as well as the 
possibilities for enhancing system stabil- 
ity by application of modern control engi- 
neering/techniques in conjunction with ex- 
isting/ conventional blade pitch control 
systems . 

In this study, the effects of the 
various design and operating parameters 
which traditionally influence the ground 
and air resonance instabilities of articu- 
lated rotor helicopters have been con- 
sidered, but with the addition of the uni- 
que hingeless rotor helicopter parameters 
such as blade internal damping and virtual 
hinge locations. The effect of scale on 
stability is investigated by considering 
aerodynamical ly similar designs which 
range in gross weight from 5,120. pounds to 
48,000 pounds by keeping tip speed and 
mean rotor lift coefficient constant. Sev- 
eral other cases of general interest are 
also considered, such as off-loading, rpm 



reduction, increasing blade number, etc. / 
In view of the enormous control power availf 
able with a hingeless rotor due to its / 
structural characteristics and the possible 
need for or desirability of full artificial 
stabilization or stability augmentation of 
certain design configurations or operating 
conditions, a closed loop stabilization 
approach is also investigated. It is 
viewed as an evolutionary approach which 
would employ a conventional helicopter 
swashplate type of control system of blade 
collective and cyclic pitch. A variety of 
output variables and their derivatives are 
examined as possible sources of closed 
loop feedback information for control ac- 
tuation. The roll and the roll rate vari- 
ables are seen to be highly effective. 
The dynamics of cyclic and collective 
pitch change are also examined4 as part of 
such a closed loop stabilization system 
for ground and air resonance where the 
control process is seen to be that of a 
multiple input-multiple output, interact- 
ing control system5. 

Detailed parametric studies of the 
ground and air resonance stability bound- 
aries are carried out using a standard 
eigenvalue routine. The parameter combin- 
ations which can result in the instabili- 
ties are examined with a view towards com- 
paring designs with inherent stability 
with those that are a result of artificial 
stabilization. Finally those combinations 
of design, operating and stability augmen- 
tation parameters yielding hingeless rotor 
type aircraft free of the ground and air 
resonance instabilities are obtained. 

Analysis 

The analysis is carried out with the 
objective of developing a broad understand- 
ing of the influence of the principal de- 
sign and operating parameters on both the 
system air and ground resonance instabili- 
ties. Consequently the degrees of freedom 
chosen for the analytical model are those 
which can be expected to be common to all 
hingeless rotor helicopter designs in hov- 
ering and on the ground, irrespective of 
size and gross weight, operational require- 
ments or specific structural design ap- 
proaches . 

The fuselage body degrees of free- 
dom are taken as those which would repre- 
sent both the hovering and ground oscilla- 
tions of a single rotor helicopter either 
in the air or on a three point, conven- 
tional oleo-shock strut type of landing 
gear. These then follow as the lateral, 
longitudinal and vertical translational 
degrees of freedom and the angular roll 
and pitch degrees of freedom. A yawing 
degree of freedom is not included, since 
it is deemed an unnecessary and unproduc- 



206 



tive complication of marginal significance. 
This follows from the large yawing inertia 
of the body, the close proximity of the 
aircraft center of gravity to the two main 
landing gear and the rotor thrust line, 
the net effect of which is to virtually 
decouple the yawing freedom from the 
others, and thereby effectively eliminates 
its influence on the air and ground reson- 
ance instabilities. 

The landing gear type and arrange- 
ment used in the analysis of ground reson- 
ance are viewed as typical, but by no 
means universal. However, the effective 
spring and viscous damping restraints 
which are arrived at in the landing gear 
analysis are sufficiently broad in charac- 
ter to be representative of the many diff- 
erent landing gear systems currently in 
use. The two most prevalent systems are 
the skid type, and pneumatic shock strut 
and tire type configurations. Since the 
skid-type landing gear represents a spec- 
cial case of the more general shock strut 
and tire formulation, an analytic model of 
the latter has been employed. This formu- 
lation has the added advantage of permit- 
ting various effects, such as the shock 
strut damping, the non-linear pneumatic 
spring rate and the combined spring rate 
of the tire and landing surface to be more 
easily studied and is developed in detail 
in Reference 6. 

The hingeless rotor blades are flex- 
ible, cantilever structures which flap 
elastically in oscillations normal to the 
plane of rotation and lead-lag elastically 
in the plane of rotation. A generalized 
coordinate, normal mode type of analysis 
could be employed effectively for the 
structural dynamic aspects. However this 
does not lend itself well to a simple de- 
termination of the aerodynamic forces and 
moments which play a central role in the 
stabilization process because of the blade 
bending curvature during the oscillations. 
Consequently the concept of virtual springs 
and hinges 7 ' 8 for the flapping and lead- 
lag oscillations of the blade is used, 
where quasi-rigid body blade motions are 
introduced to replace the continuous, elas- 
tic bending deformations of the real 
blades. These degrees of freedom are il- 
lustrated in Figure 1. An isometric view 
of the body degrees of freedom is also 
shown . 

The blade pitch changes are treated 
as constrained degrees of freedom in the 
stability analysis. That is the blade 
pitch can be changed collectively or cy- 
clically by displacement or tilting of a 
swashplate mechanism. In the open loop 
case this is done by the pilot displacing 
the collective or cyclic pitch control 
sticks. This results in a transient re- 
sponse of the aircraft either about its 
initial hovering state or on its landing 
gear by altering the aerodynamic forces 
and moments produced by the hingeless rotor. 



Since it takes the form of a reference in- 
put or external disturbance, it has no ef- 
fect on the system stability as long as 

these disturbances are reasonably small. 
In the general closed loop case the air- 
craft roll position, roll rate, pitch posi- 
tion and pitch rate are sensed and used to 
drive a system of swashplate actuators with 
a view towards employing the enormous con- 
trol power inherent in the cantilever blade 
design of hingeless rotor systems. This 
can yield full stabilization, if required, 
or it can augment the inherent stability of 
the system when design and operating condi- 
tions permit. The swashplate-blade pitch 
change arrangement and the system block 
■ diagram are shown schematically in Figure 2. 
More sophisticated closed loop control 
system arrangements offer the possibility 
of enhanced performance and optimization 
,of the system at the expense of complexity 
or possible reduction in reliability. For 
example an inner control loop on rotor 
blade bending deflections by strain gage 
techniques, as well as sensing of body 
translational displacements and velocities 
offer interesting possibilities which are 
considered in Reference 9. Needless to 
say, departure of blade pitch from the set- 
tings called for by the control system com- 
plicates and may degrade the stability and 
controllability of the system. For example 
blade torsion which is not included in this 
study is an important factor considered in 
Reference 16. 

The combination of the fuselage, 
landing gear (when applicable) and the ro- 
tor blade systems yields 5+3N freedoms in 
the closed loop case and 5+2N freedoms in 
the open loop case where the blade number 
N is at least four. The minimum number of 
four blades follows from the possibilities 
of a dynamic instability unique to two- 
blade systems^ and resonant amplification 
of three blade aerodynamic loadings in the 
case of three bladesll which must be avoid- 
ed by using a minimum of four blades in a 
hingeless rotor system. 

The number of blade freedoms is re- 
duced by introduction of quasi-normal ccr, 
ordinates to describe the rotor motions iz ' J - J . 
This approach reduces the complexity of the 
analysis by eliminating all blade motions 
which do not couple with the body in a co- 
herent manner during open and closed loop 
oscillations. These coordinates describe 
the various significant patterns of blade 
motion by five degrees of freedom in the 
open loop case. These are the rotor cone 
vertex angle, the lateral and longitudinal 
tilt of the rotor cone, and the lateral 
and longitudinal displacements of the blade 
system center of gravity with respect to 
the geometric center of the rotor (due to 
lead-lag motion in the rotating frame of 
reference) . In the closed loop case three 
freedoms are added through the displace- 
ments of the swashplate for blade collec- 
tive pitch changes and by the angular tilt- 
ing of the swashplate for blade lateral 
and longitudinal cyclic pitch changes. 



207 



The analysis proceeds assuming 
that the rotor system has four blades. 
The single exception to this is the con- 
sideration of a very heavy helicopter 
(72,000 lbs.) air resonance behaviour. 
In this case a six blade design obtained 
by adding two blades to a four blade, 
48,000 lb. design is examined. This 
leads to a final quasi-normal coordinate 
model which has ten degrees of freedom 
for the open loop case and thirteen for 
the closed loop case. These equations of 
motion are then reduced to a canonical 
form suitable for application of a stan- 
dard digital computer routine for deter- 
mining the complex eigenvalues and eigen- 
vectors of the system. In effect twice 
the number of first order, linear differ- 
ential equations with constant coeffi- 
cients result. This is a twenty-sixth 
order system in the closed loop case, if 
ideal actuators are assumed. As more 
realistic models of the control hardware 
are employed (due to leakage across hy- 
draulic seals, imperfect relays, ampli- 
fier frequency response characteristics, 
etc.) the order of the system would in- 
crease further. This is deemed to be a 
specialized design problem which needs 
attention on an ad hoc basis. 

Discussion of Numerical Results 

The discussion of the numerical re- 
sults begins with the open loop stability 
or stability boundary characteristics of 
the hingeless rotor helicopter ground re- 
sonance problem and is then followed by an 
examination of the potential influence of 
closed loop, feedback control in system 
stability. This approach is then repeated 
for the air resonance problem. The dis- 
cussion closes with an overview of the po- 
tential of closed loop control for both of 
these hingeless' rotor helicopter instabil- 
ities. 

Ground Resonance 

In order to develop insight into 
the nature of the ground resonance insta- 
bility as it might occur for a typical 
helicopter employing a hingeless rotor, a 
12,000 pound reference case based on the 
S-58 helicopter^ i s considered first. 
The rotor is modelled as one with four 
hingeless blades with a flapping frequen- 
cy ratio of 1.15 cycles per revolution, 
and a lead-lag frequency ratio of .70 cy- 
cles per revolution at a rotor tip speed 
of 650 ft/sec. The wheels are first as- 
sumed to be locked, preventing the air- 
craft from rolling freely in a longitu- 
dinal direction. The uncoupled lateral 
and longitudinal translation modes of the 
aircraft are assumed to have five percent 
and three percent of critical damping, re- 
spectively, as a result of tire hysteresis 



losses. As the thrust to weight ratio is 
varied from zero to unity the vertical 
loading on the landing gear decreases. 
The stability of the small, coupled oscill- 
ations about a series of initial steady 
states determined by the thrust to weight 
ratio (T/W) is then studied as a function 
of oleo-shock strut damping for several 
small, but typical values of blade hys- 
teresis lead-lag damping. Both damping 
parameters are expressed in terms of per- 
cent of equivalent viscous critical damp- 
ing. 

The unstable mode of oscillation is 
found in all cases to be dominantly a 
fuselage rolling mode with a small amount 
of lateral translation coupling, and still 
lesser amounts of pitching and longitu- 
dinal motion. Release of the brakes, per- 
mitting the aircraft to move freely longi- 
tudinally, has a slightly stabilizing ef- 
fect, but of minor importance compared to 
the influence of oleo-shock strut damping 
and blade internal damping. The numerical 
results of the study with brakes on are 
presented in Figure 3. Equivalent viscous 
damping of the uncoupled rolling mode ex- 
pressed in percent of critical damping is 
taken as the abscissa, while thrust to 
weight ratio is the ordinate. The hori- 
zontal dash line at T/W = .9 is a visual 
reminder that this is an unrealistic condi- 
tion and that the stability data beyond 
this value is probably unreliable, since 
the analytical modelling of the landing 
gear depends on the questionable assumption 
of an initial steady state for thrust to 
weight ratios greater than nine tenths. 
The aircraft is, of course, in the trans- 
ient condition of landing or take-off. 

It is seen that if blade hysteresis 
damping should be equivalent to one percent 
of critical lead-lag damping, then slight 
amounts of oleo-damping of the rolling 
mode produce stable oscillations . If the 
blade internal damping is as little as one 
quarter of a percent of critical, stability 
can still be achieved for all thrust to 
weight ratios, if roll damping is equiva- 
lent to fourteen percent of critical damp- 
ing. Internal blade damping of one percent 
or greater is found to eliminate the in- 
stability entirely, if only slight amounts 
of landing gear damping are available, for 
example from tire hysteresis . Thus the 
ground resonance instability for the refer- 
ence case is found to be quite mild and 
easily eliminated with the moderate amounts 
of blade and landing gear damping normally 
present. 

In order to understand the influence 
of the tuning of a hingeless rotor on this 
desirable result, the lead-lag frequency 
ratio is varied about reference frequency 
ratio of .7 cycles per revolution as the 



208 



flapping frequency ratio is held constant 
at 1.15 cycles per revolution. Blade damp- 
ing is taken at one-half percent of criti- 
cal while roll damping is held fixed at 
eight percent of critical. Figure 4 shows 
the effect of this tuning on the unstable 
mode by plotting the log decrement of this 
mode versus thrust to weight ratio. It is 
seen that increasing the lead-lag frequen- 
cy ratio above .7 makes the system stable, 
while decreasing it below this reference 
value makes it progressively more unstable. 
Figure 5 considers the effect of the off- 
set of the virtual flapping hinge and tun- 
ing of the flapping frequency ratio on the 
instability with respect to the reference 
case. It is seen that a flapping frequen- 
cy ratio of 1.0 corresponding to a conven- 
tional, articulated rotor is considerably 
more unstable than the reference case. It 
is seen that increasing the offset and 
frequency ratio to progressively higher 
values is beneficial and stabilizing al- 
though tending to reach a point of dimin- 
ishing returns at a flapping frequency 
ratio of 1.20 cycles per revolution. 

Size and scale effects are investi- 
gated by considering the coupling of the 
lateral and rolling motion as the distance 
between the rotor hub and the center of 
gravity of the aircraft is varied with re- 
spect to the reference case, where it was 
assumed to be at a distance of seven feet. 
As this distance is decreased to five feet, 
the instability is observed to change in 
relationship to the thrust to weight ratio, 
but not in general character. On the 
other hand as the coupling increases by in- 
creasing the distance to nine feet, there 
is a stabilizing effect. This is illus- 
trated in Figure 6- This result can be 
understood in terms of the coupled rolling 
natural frequency, which tends to decrease 
as this distance increases. Thus if the 
lead-lag natural frequency ratio is held 
fixed at .7, stability can be improved by 
detuning the fuselage coupled rolling mode 
to a lower frequency. This result is typ- 
ical of all helicopter ground resonance 
instabilities . 

The influence of large size and 
scale changes is considered by studying 
the stability of two additional hingeless 
rotor helicopters of 5,120 and 48,000 
pounds, respectively, which are obtained 
from the reference case by aerodynamic 
scaling. That is the rotor diameter and 
overall proportions of the aircraft were 
altered to accomodate the gross weight 
changes at the same mean rotor lift co- 
efficient and tip speed. It is seen in 
Figure 7 that aircraft smaller than the 
reference case of 12,000 pounds tend to- 
ward inherent stability with the blade 
tuning and nominal amounts of damping as- 
sumed. On the other hand the relatively 



heavy machines tend to a more severe insta- 
bility at slightly higher thrust to weight 
ratios than the reference case, but still 
well within the range of achieving inher- 
ent stability with moderate amounts of 
blade hysteresis damping and oleo-shock 
strut damping of the unstable, coupled 
rolling mode. 

Ground Resonance with Feedback 

As an alternative or as a supple- 
ment to parameter selection which results 
in stable oscillations, closed loop feed- 
back control is considered. Since pro- 
portional control action (at least quali- 
tatively) alters the frequency of oscilla- 
tion of simple systems by adding or sub- 
tracting a virtual spring effect, depend- 
ing on whether feedback is negative or pos- 
itive, the reference case was used as a 
basis for investigating this possibility. 
Figure 8 shows the effect of proportional 
roll feedback and control action (in this 
case positive feedback is actually employ- 
ed) in detuning an unstable coupling by 
depressing the critical fuselage roll mode 
frequency. It is seen that this is very 
effective in stabilizing the system. It 
should be noted that in the case of other 
design reference parameters, proportional 
feedback and control action of opposite 
sign might be beneficial, if the detuning 
of the critical fuselage roll frequency 
required increasing, rather than decreas- 
ing. The application of this control ac- 
tion is deemed beneficial, but is best de- 
cided on an ad hoc. basis. 

A more conventional use of feedback 
control is considered in Figure 9 which 
shows the effect of negative feedback with 
derivative or rate control action. This 
tends to augment the damping of the criti- 
cal fuselage rolling mode. This is seen 
to be highly effective also, and, at least 
to a first approximation, is interchange- 
able with oleo-shock strut damping of the 
unstable roll mode. 

A logical extension of the fore~ 
going application of feedback control to 
the stability of ground resonance is the 
blending of both proportional and deriva- 
tive control action. In this case the 
critical roll mode can be both detuned and 
damped to approach an optimum. This is 
shown to be the case in Figure 10. Here 
the system is made progressively more sta- 
ble over the entire range of thrust to 
weight ratios. It is not the intention 
here to optimize the stability boundary, 
but to show that this is possible even 
with small values of blade internal hys- 
teresis damping and the normal amounts of 
landing gear damping of the reference case, 
In view of the relatively unimportant in- 
fluence of the pitching, and longitudinal 



209 



degrees of freedom for the reference case, 
pitch rate feedback and control action was 
not deemed effective. However, this re- 
mains a potentially useful and important 
tool in the event that special design or 
operational requirements modify the open 
loop system. 

Air Resonance 

The basic reference helicopter of 
12,000 pounds gross weight is examined for 
its air resonance stability as a function 
of lead-lag frequency ratio for several 
values of flapping frequency ratio. It is 
seen in Figure 11 that lead-lag frequency 
ratios of .70 or less result in instabil- 
ity over the structurally feasible range 
of flapping frequency ratios between 1.10 
and 1.20. It is also to be noted that in 
the neighborhood of neutral stability (for 
the assumed blade equivalent viscous in- 
ternal damping of 1/2%) , increasing flap- 
ping frequency ratio is stabilizing. This 
interaction effect between these two key 
blade natural frequency ratios is further 
illuminated in Figure 15. It can also be 
seen that the lighter blades (i.e. an Over- 
all mass fraction of 4%% rather than 6%%) 
require higher frequencies for neutral sta- 
bility. It is shown in Reference 12 that 
in the stable range of lead-lag frequency 
ratios, an increasing helicopter blade 
mass fraction improves stability further. 
On the other hand, it is also shown that 
for an unstable configuration, increasing 
blade mass fractions can further degrade 
stability. 

The critical effect of internal 
damping of the blade lead- lag motion is 
presented in Figure 13 for the reference 
case with a flapping frequency ratio of 
1.10 (comparable results were obtained at 
frequency ratios of 1.15 and 1.20). It is 
seen that increased internal damping en- 
hances air resonance stability and inter- 
nal damping levels of 1% of critical vir- 
tually eliminate the air resonance insta- 
bility for a lead-lag frequency ratio of 
.75 or greater (since hingeless rotor 
flapping frequency rates greater than 1.10 
improve stability further) . 

Although the frequency ratios for 
lead-lag and flapping motions of hingeless 
rotor blades have a fundamental effect on 
the air resonance stability boundaries, de- 
sign differences in structure, materials, 
and proportioning of such blades can re- 
sult in differences in the virtual hinge 
locations and stiffness with important 
modifications in the stability boundaries. 
These effects are presented in Fiaure 14, 
which show that more outboard location 
of the virtual hinges for lead-lag mo- 
tion tends to stabilize, although not 
by a substantial degree. This effect 



is believed to stem from a decrease in the 
relative energy level of the blade in-plane 
motion, just as in classical ground reson- 
ance. 

Size effects as distinguished from 
gross weight are presented in Figure 15. 
It is seen that the reference helicopter 
air resonance stability is virtually un- 
affected by large changes in the body 
pitch and roll moments of inertia, pro- 
vided that the lead- lag frequency ratio is 
sufficiently large for stability {<»)££=. 75) . 
However relatively large machines are seen 
to be less unstable, if an air resonance 
instability exists. The influence of gross 
weight changes through aerodynamic scaling 
is presented in Figure 16 for 5120, 12,000 
and 48,000 pound machines which operate at 
the same mean rotor lift coefficient and 
tip speed. It is seen that very large in- 
creases in gross weight tend to be stabil- 
izing with respect to the minimum lead- lag 
frequency ratio required for neutral sta- 
bility, although gross weight effects for 
machines in the 5,000 to 12,000 pound class 
are not clear-cut because of the greater 
sensitivity to all the other system para- 
meters. In fact, it may be difficult to 
obtain a rational trend when blade mass 
fraction is held constant, when in reality 
the very small machines will tend toward 
larger blade mass fractions. In contrast 
to this, if the gross weight of the refer- 
ence machine is decreased by off-loading 
(cargo, for example) , there is a clear-cut 
improvement in the air resonance stability. 
This is shown in Figure 17 and stems from 
the reduction in blade initial coning. 
The effect of coning is discussed further 
below. 

Built-in pre-coning angles are nor- 
mal in the design of hingeless rotors", to 
minimize steady bending stress is a rou- 
tine consideration. Figure 18 shows the 
stability boundary for the reference case 
and the effect of deviating from the nom- 
inally ideal case of built-in pre-coning 
matching the coning that would result from 
a 1-g load of a centrally hinged, articu- 
lated rotor. It is seen immediately that 
"over-coning" destabilizes and "under-con- 
ing" stabilizes for the entire range of 
lead-lag frequency ratios. This suggests 
that a direct, profitable trade-off between 
steady bending stress and air resonance 
stability exists. That is reduce coning 
by structural action and enhance stability. 
Figure 19 continues this theme by showing 
the influence of a concentrated tip weight 
on the air resonance instability. In this 
case it is seen that tip weight is benefi- 
cial and stabilizing, providing the lead- 
lag frequency ratio is of the order of 
three-fourths or greater (^£^.75) . Fig- 
ure 20 shows the design effect of an RPM 
reduction at fixed gross weight. This 



210 



would increase coning and the data shows a 
consistent loss of stability for the var- 
ious lead-lag frequency ratios . 

Aerodynamic scaling for very large 
helicopters appears to be barred by the 
adverse trend of coning at constant mean 
rotor lift coefficient and tip speed, un- 
less blade number is increased beyond four 
blades. For example increasing gross 
weight from 48,000 pounds to 72,000 pounds 
was considered by increasing disk loading 
and solidity by fifty percent - that is 
adding two blades to the original four- 
blade design. This yields the beneficial 
effect of no increase in coning angle and 
only a minor modification of the stability 
boundary. This is illustrated in Figure 
21. The dash or ghost line on this figure 
represents a second mode of marginal sta- 
bility at a high frequency. This is dis- 
cussed at length in reference 15. The im- 
plication is that a high frequency air re- 
sonance instability might become a factor 
in very large hingeless rotor helicopters. 
However, the effect of including the addi- 
tional rotor degrees of freedom suppressed 
by the "quasi-normal" or "multi-blade" co- 
ordinate transformations requires addi- 
tional, careful study since the current 
analysis limits rotor flapping type mo- 
tions to those which result in either col- 
lective or cyclic flapping of the indivi- 
dual blades . 

Air Resonance with Feedback 

Proportional feedback and control 
action proves to be a very effective means 
of stabilizing air resonance. Figure 22 
shows the influence of proportional roll 
control action for the reference helicop- 
ter; roll corrections alone are found to 
be highly effective over the entire range 
of lead-lag frequency ratios, whereas air- 
craft pitching motion is found to be a 
relatively small component of the air re- 
sonance instability mode and not a produc- 
tive avenue for closed loop stabiliza- 
tion. 15 Figure 23 examines the efficacy 
of proportional roll control for a case of 
maximum air resonance instability when 
u)££=.60. It is seen to be very effective 
and virtually a linear influence on sta- 
bility over the range of practical inter- 
est. 

Sensing aircraft roll rate is also 
found to be highly effective in closed 
loop control, but less so for pitch rate 
because of the relatively small participa- 
tion of pitch in the air resonance insta- 
bility. However the complex phase rela- 
tionships which exist in the mode of air 
resonance instabilityl 5 make it very de- 
sirable that aircraft roll and pitch con- 
trol actions be mixed (i.e. the interact- 
ing control actions referred to above 5 ) . 



This is illustrated in Figure 24 which 
shows the influence of pitch control ac- 
tion for several levels of roll control ac- 
tion (where both are based on roll rate 
feedback information) . The linearity of 
this stabilization method is made evident 
by cross-plotting the influence of pitch 
control action for a magnitude of roll 
control action which results in (almost) 
neutral stability. 

Closed Loop Stabilization 

The foregoing data illustrates that 
an appropriate mix of aircraft roll and 
roll rate information, in conjunction with 
aircraft roll and pitch control action, 
permits straightforward artificial stabili- 
zation of both the air and ground reson- 
ance instabilities of hingeless rotor heli- 
copters under very adverse design condi- 
tions. More importantly, perhaps, the da- 
ta indicates that the marginally stable 
configurations resulting from the lead-lag 
frequency ratio being tuned to .70-. 80 
and/or internal damping levels for this os- 
cillation being of the order of %% of cri- 
tical or less can be easily stabilized by 
utilizing existing, conventional control 
systems . 

A significant difference between 
the ground resonance and air resonance 
modes of instability is the phase rela- 
tionship between rotor cone tilting and 
fuselage rolling motion. Also the fact 
that a slight positive or regenerative roll 
feedback and control action can be benefi- 
cial in stabilizing ground resonance. The 
reverse is true for air resonance. The 
common, beneficial element for both insta- 
bilities is in sensing aircraft roll rate 
and utilizing this information for nega- 
tive feedback to implement roll control 
action. This in effect is stability aug- 
mentation of the aircraft roll damping both 
on the ground and in the air. The addi- 
tional control action for aircraft pitch 
has been found to be beneficial for sta- 
bilizing air resonance 3 - 5 and not detrimen- 
tal for stabilizing ground resonance. 6 
Thus the interacting, closed loop control 
system driven by roll rate information 
emerges as a simple, evolutionary approach 
to complete artificial stabilization, or 
stability augmentation of the hingeless 
rotor helicopter air and ground resonance 
instabilities . 

Conclusions 

1. The ground and air resonance in- 
stabilities of hingeless rotor helicopters 
are marginal ones, but they will persist 
as design considerations because of the 
natural tendency of the lead-lag frequency 
ratios to be less than unity (and conceiv- 
ably as small as .60), while internal damp- 



211 



ing levels will be slight, unless special 
materials and design measures which in- 
crease internal damping can be found and 
which are acceptable with respect to other 
design and operating constraints. 

2. The air resonance instability 
is very sensitive to blade coning, while 
ground resonance is not. Reductions in 
coning by a variety of means are benefi- 
cial, but the possibility of accepting a 
modest level of steady bending stress in 
lieu of other approaches (such as tip 
weights) is worthy of more consideration 

(since this would also reduce Coriolis- 
type fatigue loads in steady forward 
flight) . 

3 . Completely artificial stabiliz- 
ation of both the air and ground resonance 
instabilities is possible, by utilizing 
the concept of interacting controls. This 
is not suggested as a serious approach to 
design, but as an indication that a modest 
stability augmentation approach, in con- 
junction with adherence to simple design 
criteria and objectives, can eliminate 
both the air and ground resonance instabil- 
ities. 

4. The ground resonance instabili- 
ty which was studied exhaustively in Ref- 
erence 6 is seen to be inherently the same, 
whether conventional oleo shock strut or 
skid type landing gear are used, providing 
the effective stiffness and damping are 
properly represented in the overall system 
design. On the other hand, failure or mal- 
function of a single element of the system 
which destroys the assumed symmetries (e.g. 
a single blade damper on an articulated ro- 
tor system) must be evaluated on an ad hoc. 
basis since the system might then become 
unstable despite the stability of the nor- 
mal system. 

References 

1. Donham, R. E., Cardinale, S. V., and 
Sachs, I. B., "Ground and Air Reson- 
ance Characteristics of a Soft In- 
Plane Rigid-Rotor System," AIAA/AHS 
VTOL Research, Design and Operations 
Meeting, Georgia Institute of Technol- 
ogy, Atlanta, Georgia, February 1969. 

2. Woitsch, W. and Weiss, H., "Dynamic Be- 
havior of a Fiberglass Rotor," AIAA/ 
AHS VTOL Research, Design and Opera- 
tions Meeting, Georgia Institute of 
Technology, Atlanta, Georgia, February 
1969. 

3. Lytwyn, R. T., Miao, W. , and Woitsch, 
W., "Airborne and Ground Resonance of 
Hingeless Rotors," 26th Annual Forum 
of The American Helicopter Society, 
Washington, D.C., June 1970, Preprint 
No. 414. 



4. Young, M. I., "The Dynamics of Blade 
Pitch Control," Journal of Aircraft, 
Vol. 10, No. 7, July 1973. 

5. Ogata, K. , Modern Control Engineering 
Prentice-Hall, Englewood Cliffs, N.J., 
1970, pp. 377-396. 

6. Bailey, D. J., "Automatic Stabiliza- 
tion of Helicopter Ground Instabili- 
ties," University of Delaware, Master 
of Mechanical and Aerospace Engineer- 
ing Thesis, May 1973. 

7. Young, M. I., "A Simplified Theory of 
Hingeless Rotor Helicopters," Proceed- 
ings of the Eighteenth Annual National 
Forum of The American Helicopter Soci- 
ety, Washington, D.C. , May 1962, 

pp. 38-45. 

8. Ward, J. F. , "A Summary of Hingeless 
Rotor Structural Loads and Dynamics Re- 
search, " Journal of Sound and Vibra- 
tions, 1966, Vol. 4, No. 3, pp. 358-377. 

9. Young, M. I. , Hirschbein, M. S., and 
Bailey, D. J. , "Servo-Aeroelastic Pro- 
blems of Hingeless Rotor Helicopters," 
University of Delaware, Department of 
Mechanical and Aerospace Engineering, 
Technical Report No. 155, August 1972, 
(Revised October 1973) . 

10. Kelley, B., "Rigid Rotors vs. Hinged 
Rotors for Helicopters," Annals of The 
New York Academy of Sciences, Vol. 107, 
Article 1, 1964, pp. 40-48. 

11. Marda, R. S., "Bending Vibrations of 
Hingeless Rotor Blades," University of 
Delaware, Master of Mechanical and 
Aerospace Engineering Thesis, April 
1972. 

12. Young, M. I. and Lytwyn, R. T. , "The 
Influence of Blade Flapping Restraint 
on the Dynamic Stability of Low Disk 
Loading Propeller-Rotors," Journal of 
The American Helicopter Society, Vol. 
12, No. 4, October 1967, pp. 38-54. 

13. Hohenemser, K. H. and Sheng-Kuang, Y., 
"Some Applications of Multiblade Co- 
ordinates," Journal of The American 
Helicopter Society, Vol. 17, No. 3, 
July 1972, pp. 3-12. 

14. Seckel, E., Stability and Control of 
Airplanes and Helicopters, Academic 
Press, 1964, pp. 456-457. 

15. Hirschbein, M. S., "Flight Dynamic 
Stability and Control of Hovering Heli- 
copters," University Of Delaware, Mas- 
ter of Mechanical and Aerospace Engi- 
neering Thesis, October 1973. 

16. Huber, H. B. , "Effect of Torsion-Flap- 
Lag Coupling on Hingeless Rotor Sta- 
bility," Preprint No. 731, 29th Annual 
National Forum of the American Heli- 
copter Society, Washington, D. C. , 
May 1973. 



212 




V flapping axis of 

rotation 




STABLE REGION 



2 4 6 8 10 12 14 16 18 20 



Fig. 1 XYZ - coordinate system, x, y, z, u\ t &2 
displacements. 



Fig, 3 Stability boundary as a function of roll 
damping . 




.BLADE PITCH 
CONTROL AXES 



DISTURBANCES 



ira> 



3LAPE PITCH CHANGE 

AND SWASH-PLATE 
DYNAMICS 



FUSELAGE AND ROTOR 
SYSTEM DYNAMICS 



FEEDBACK DEVICE 
CHARACTERISTICS 



80 








X 


1 
1 
1 
1 
1 
1 

1 




40 


UNSTABLE 








1 

1 

s 1 































s 


\. r~~ 


•~~ f 












1 






STA 


3LE 








i_ 




40 






.5 


4/ 


1 




-SO 


1 »« 




2 m tt 


" 


.6 




| 






3 -u 


- 


,7 




1 






AU H 


- 


.8 




1 






5 u u 


s= 


.9 


. 


i 





T/W 



.B 1.0 



Fig. 2 



Fig. 4 The effect of lead-lag natural frequency 
ratio . 



213 




.8 1.0 




Fig. 5 The effect of flapping natural frequency 
ratio and virtual flapping hinge offset. 



Fig. 7 Size effects. 



** 



40 


1 t - .61 

2 t = .75 

3 t - .83 


/ »» \ 


I 
1 
1 
1 




I r x + ffi2 l 


20 










UNSTABLE^ 


1 /" "" 


1 








1 \] 








i > f> 




STOBLE 




1 


-20 






1 


-40 




2 / *■ 


i 




-——_______ 


3 ^/ 


i 
i 


-60 







.2 .4 .6 .8 1.0 
T/W 




Fig. 6 The effect of coupling between lateral 
and rolling motion. 



Fig. 8 The effect of roll position feedback 
control. 



214 



Wt. - 12000 lb. 
6 1/2% 





40 








20 











UNSTABLE 








' 


/ ' % y // 




c 

3 
1 

-1 


5 -20 


1 
STABLE 


I/ y/ 




-40 




f 1 NO FEEDBACK 
da- 

2 M 2-"-3»- 

da, 

3 »2 " 20 -dF 




-60 




da, 
4 M 2 .. BO.jJ 




C 


.2 


.4 


.6 .3 


1. 



Fig. 9 The effect of roll rate feedback control. 



■a 



40 










20 












UNST 


HBLE 




1 






STA 


<L£ 


\ / 


2 / ^^ 




-20 






^X \ 


-40 




' 1 

2 

3 


"^NOFEEDBACK I 
M 2 - -2. « 2 +20. j^- 






4 


'da 
M 2- -*■ "2 +2I> - W 


60 






. 1 



.8 1.0 



Fig. 10 The effect of roll position and roll rate 
feedback control. 



ft 









c/c c 


- 1/2% 








e l 


» .10 


\ 






e 2 


- .10 


■ . 








u £ - 1.15 










u f « 1.20 


.50 


.60 


.78 \ 


v°\ 


.90 1.00 













Fig. 11 The effects of lead-lag natural frequency 
ratio . 



Wt. " 12000 lb 
C/C - 1/2% 



1.25 




e l * * X0 






e" - .10 

ll)j = 1.15 


1.20 






1.15 


S 


^ Stable 

<0^ 


1.10 


Unstable 




1.05 


^ » 6 1/2% ' 

^ - 4 1/2% , 




1.00 







.40 .50 .60 .70 .80 .90 1.00 

Fig. 12 Flapping-in-plane natural frequency 
ratio stability boundary. 




Wt. - 12000 lb. 
^ - 6 1/2% 

" - .10 



Fig. 13 The effect of in-plane blade structural v 
damping. 



215 



wt. 


- 12000 lb. 


■h 


- 4 1/2* 


c/c c 


- 1/2% 




u f - 1.10 








e x - .05 




*- <t 2 ' -IS 




5j - .10 




'-•- - -10 




n^ - 6 1/2% 
0/C o » 1/2% 
= .10 

- .10 

- 1.15 




Wt.- 48000 lb. 



Wt.= 12000 lb. 
Wt.- 5120 lb. 



Fig. 14 The effect of virtual hinge offset. 



Fig. 16 The effect of scaling. 



Wt. - 12000 lb. 
ij, - 6 l/2» 

c/c c - 1/2% 

- .10 




Fig. 15 The effect of changes in the body 
moments of inertia. 




n^ - 6 1/2% 
C/C c = 1/2 S 
= .10 
= .10 



12000 lb. 
Wt. = 8000 lb. 



t$ f.00 



Fig. 17 The effect of off loading without 
changing Ixq and Iyg. 



216 



ft 




Wt. - 12000 lb. 
«fe - 10% 
C/C - 1/2% 




Fig. 18 The effects of pre-cone angle. 



Fig. 20 The effects of a 10% RPM reduction. 





Hjj - 6 1/2% 
C/C c - 1/2% 

.» .10 

- .10 

- 1.15 



■Lower Frequency Instability 
«* - <!-»„) 



Higher Frequency Instability 
o* - (l+o„) 



wt. » 72000 lb. 
Wt. = 48000 lb. 



Fig. 19 The effect of tip weights. 



Fig. 21 Stability of heavy helicopters. 



217 



wt. 

"b 



' 12000 lb. 
■ 4 1/2* 









c/c c 


- 1/2* 

- .10 




4 






e 2 - .10 
u f " 1.15 




3 






M l /o 2 " 1 '° 




2 






B 2 /a 2 =0.0 




1 













*-^— .^6 .60 


.70 


.80. 


.90 


-1 


• 








-2 










-3 


Open Loop — — 








-4 


■ 








-5 


. 










-6 


Closed Loop —~ 










-7 


■ 










-8 


, 











Fig. 22 Comparison of open loop stability to 
closed loop stability with roll 
position feedback. 



wt. 


m 


12000 lb. 


"h 


• 


4 1/2* 


VC r 


« 


1/2* 


•l 


» 


.10 


e 2 


" 


.10 


"f 


. 


1.15 


".. 


« 


.60 



Wt. 
"b 

c/c 




12000 lb 
4 1/2* 
1/2* 


•l 




.10 
.10 




- 


1.15 




« 


.60 





Fig. 24 The effect of roll rate feedback with 
pitch and roll interacting control 
actions. 



Fig. 23 The effect of roll position feedback 
control at maximum instability. 



218 



VERTICAL-PLANE PENDULUM ABSORBERS FOR 
MINIMIZING HELICOPTER VIBRATORY LOADS 



Kenneth B. Amer 
Manager, Technical Department 

James R. Neff 
Chief, Dynamics Analysis 

Hughes Helicopters 
Culver City, California 



Abstract 

This paper discusses the use of pendulum dy- 
namic absorbers mounted on the blade root and 
operating in the vertical plane to minimize heli- 
copter vibratory loads. 

The paper describes qualitatively the concept 
of the dynamic absorbers and presents results of 
analytical studies showing the degree of reduction 
in vibratory loads attainable. Operational experi- 
ence of vertical plane dynamic absorbers on the 
0H-6A helicopter is also discussed. 

Introduction 



In a helicopter it is important to maintain a 
low level of vibration for two reasons; first for 
the comfort of the crew and passengers, and 
secondly to minimize maintenance problems. During 
early flight tests of the 0H-6A helicopter (see 
Figure 1) in 1963, a high level of 4/rev fuselage 
vibration was encountered primarily during ap- 
proach to hover and during high speed flight. 




WKKm 



Figure 1. 0H-6A Helicopter 

Various analytical studies and experimental pro- 
grams were conducted in an effort to alleviate 
this problem. The configuration finally adopted 
was vertical-plane pendulum absorbers mounted at 
the roots of the main rotor blades (see Figure 2) . 
It is the purpose of this paper to describe the 
concept of the vertical-plane pendulum dynamic 
absorber and to present the results of analytical 



studies and flight tests showing the degree of re- 
duction in vibratory loads attained. 




Figure 2. Pendulum Absorbers on 0H-6A 



Over 3 million flight hours of satisfactory 
experience have been obtained with the use of 
vertical-plane pendulum absorbers on the 0H-6A 
helicopter and on its commercial counterpart, the 
Model 500 helicopter. This operational experience 
is also discussed in this paper. 

Sources of Fuselage Vibration 

The 0H-6A helicopter has a 4-bladed main rotor. 
Table I summarizes the sources of 4/rev fuselage 
vibration from the main rotor. It can be seen from 
Table I that vertical shears at the blade root with 
frequencies of 3/rev, 4/rev, and 5/rev can induce 
4/rev vibrations in the fuselage. The 3/rev and 
5/rev blade root shears induce 4/rev fuselage vi- 
brations by producing 4/rev hub moments. The 4/rev 
blade root shear produces a 4/rev hub vertical 
force. With regard to in-plane blade root shears, 
both the 3/rev and the 5/rev component of in-plane 
root shear produce a 4/rev hub horizontal force. A 
further discussion of the mechanism by which rotor 
blades induce vibration in the fuselage can be 
found in Chapter 12 of Reference 1, particularly 
the tables on pages 318 and 319. 



219 



Table I. Sources of 4/Rev Fuselage 
Vibration - 4-Bladed Rotor 



Vertical 




In-Plane 


Shear 


Load Path 


Shear Load Path 


3/ rev 


Hub moment 


3-rev Hub horizon- 
tal force 


4 /rev 


Hub vertical 
force 


- 


5 /rev 


Hub moment 


5/rev Hub horizon- 
tal force 



Table I indicates that there are 5 possible 
sources of excessive fuselage 4/rev vibration in 
the 0H-6A helicopter. The next step was to estab- 
lish which of the 5 possible sources of vibration 
were the most important. Tables II and III pro- 
vide an answer to this question. 



Table II. 0H-6A Main Rotor Blade Natural 
Frequencies (per rev) - 100% 
RPM - Pendulums Off 



Flapwise 


Chordwise (Cyclic Mode) 


2.72 
4.87 


5.14 



In Table II are listed the main rotor blade 
flapwise and chordwise natural frequencies near 
the 3/rev through 5/rev frequency. It can be seen 
from Table II that the two frequencies most likely 
to cause a 4/rev vibration in the fuselage are the 
first and second mode flapwise bending frequencies 
which are very close to 3/rev and 5/rev. The 
blade chordwise natural frequency is also close to 
5/rev (see Table II). However, Table III con- 
firms that the blade flapwise first mode and second 
mode frequencies are the primary source of the 
vibration problem, in that the fuselage vibration 
is much more responsive to hub moments than it is 
to hub vertical or horizontal forces. 



Thus blade vertical bending at a frequency of 4/rev 
and blade chordwise bending at frequencies of 3/rev 
and 5/rev can be ignored and the primary sources of 
vibration can be concluded to be blade flapwise 
bending at 3/rev and at 5/rev. 

Concept of Vertical-Plane Dynamic Absorbers 

Based on the above evaluation, it was con^- 
eluded that it was necessary to reduce the level of 
blade 3/rev and 5/rev flapwise bending. After in- 
vestigating a number of possible approaches,* it 
was decided to pursue the concept of a dynamic vi- 
bration absorber which is discussed in Reference 2 
in the section starting on page 87. 

The concept of a dynamic vibration absorber 
consists of adding a small mass to a large mass. 
The uncoupled natural frequency of the small mass 
(vibration absorber) is chosen to be equal to the 
frequency of the disturbing force. Thus, for the 
OH-6 vibration problem, it was concluded that it 
would be necessary to incorporate two dynamic 
vibration absorbers; one tuned at 3/rev and the 
other tuned at 5/rev. Furthermore, inasmuch as 
rotor speed can vary somewhat, it was necessary 
that the vibration absorbers maintain the proper 
frequency relative to rotor speed. In order to ac- 
complish this, it was decided to use the concept of 
a tuned centrifugal pendulum discussed on page 219 
of Reference 2. This concept has been used for 
many years to minimize the torsional vibrations of 
piston engines. Thus, the final configuration that 
evolved consisted of two pendulums mounted at the 
roots of the main rotor blades; one tuned to a 
natural frequency of 3/rev, the other tuned to a 
natural frequency of 5/rev. Inasmuch as the shear 
force and blade motion which were to be minimized 
were in the vertical plane, the dynamic pendulums 
were oriented to oscillate in the vertical plane. 

Figure 3 shows schematically the pendulum 
motion relative to the blade deflection for the 
case of response to 3/rev excitation. It is evi- 
dent that the centrifugal force from the pendulums 
is directed such as to cancel most of the trans- 
verse shear due to blade modal response. The net 
result is a significant reduction in the 3/rev 
vertical shear force transmitted to the hub. 



Table III. 0H-6A Cockpit Response to Rotor 
Excitation, V = 100 Knots 
(No Pendulums Installed) 





4/Rev 
Vertical 


4/Rev 

Pitching 

Moment 


4/Rev 

Rolling 

Moment 


4/Rev 

Longitudinal 

Shear 


4/Rev 

Lateral 

Shear 


Excitation 


130 


*86 


**112 


10 


35 


Force , lb 












Unit Response 
at Cockpit, 
in/sec/ lb 


.0012 


.00265 


.0106 


.00193 


.0077 


Response at 

Cockpit, 

in/sec 


.16 


.23 


1.19 


.019 


.27 



* Blade vertical shear force causing pitching moment. 
** Blade vertical shear force causing rolling moment. 



* Other approaches evaluated included providing 
control of blade first and second mode natural 
frequencies by means of anti-node weights and by 
use of preloaded internal cables. Flight tests did 
not show these methods to be sufficiently effective. 
Hub-mounted vertical plane pendulums were flown and 
proved to be effective, but considerations of drag 
and weight were unfavorable for this configuration. 
Fuselage-mounted non-rotating dampers were elimi- 
nated because of the difficulty of tuning to a 
sufficiently wide range of frequency. Fuselage- 
mounted centrifugal pendulum dampers were con- 
sidered impractical from the standpoint of space 
requirements and mechanical complexity. 



220 



I 1 1 








U« — 


i FLAPPING HINGE 






















/ 














.. 


' 














v / 














,' 


'^V 






tv. 












1 

















































1.0 
.8 
.6 
.4 
.2 

MODAL 

DEFLECTION _ 2 

-.4 

-.6 

-.8 

-1.0 

20 40 60 80 100 120 140 160 
BLADE STATION - INCHES 

Figure 3. Pendulum Motion Schematic 



Basic Physical Parameters 

The pendulum configuration that was estab- 
lished, flight tested, and put into service has the 
following characteristics : 

3/ rev pendulum 

weight: 1.8 lb 

actual mass ratio: .048 

modal mass ratio: .64 



Analytical Studies 

Analytical studies were conducted to investi- 
gate the effectiveness of vertical plane pendulum 
absorbers in minimizing the blade vertical root 
shears and the fuselage vibration levels. The re- 
sults of these analytical studies are presented in 
Table IV for the 0H-6A at a forward speed of 100 
knots. It can be seen from Table IV that the addi- 
tion of the 3/rev pendulum dynamic absorber reduces 
the 3/rev vertical root shear by 75%. The addition 
of the 5/rev vertical dynamic absorber reduces the 
5/rev vertical root shear by 85%. The net result 
is a 72% reduction in the vibration level in the 
crew compartment. 



Table IV. Effect of Vertical-Plane Pendulum 
Absorbers on Root Shear and 
Cockpit Vibration - 0H-6A 

(Analytical Studies, 100 Knots) 



Configuration 


Root 

Shear 

3/Rev 


5/Rev 


Cockpit 

Vibration, amp. 

in/ sec 


Undamped Blade 
Damped Blade 


91 
23 


42 
6 


1.8 
.5 



5/rev pendulum 

weight: .7 lb 

actual mass ratio: .019 

modal mass ratio: .52 

The pivot axis of both pendulums is located at 
15% of the blade span from the center line of the 
rotor, and 29% of the chord from the leading edge. 
This location was chosen so that existing bolts in 
the blade root fitting could be used, thus pre- 
venting the introduction of stress concentration 
points into critical sections of the blade. 
Analysis indicates that a location further out- 
board would be more favorable, but this has not 
been confirmed by test, because of the structural 
considerations cited above. 

Damping of the pendulums due to friction in 
the pivot bearings is estimated to be equivalent to 
1% of the critical viscous damping ratio for the 
3/rev pendulums at an amplitude of -16 . For the 
5/rev pendulums at the same amplitude the damping 
ratio is 3% of critical. 

The dampers are "bench" tuned, by means of 
shims, to the correct pendular frequency within 
0.5% of the length of the 3/rev pendulums and to 
within 1% of the length of the 5/rev pendulums. 
The effect of mis-tuning has been investigated only 
to the extent of showing that - one shim does not 
have a consistently observable effect on either 
qualitative or measured cockpit vibration. 



The analytical procedure used to achieve the 
results of Table IV is designated SADSAM. This 
analytical procedure is described in Reference 3 
and was conducted in two steps. In the first step, 
SADSAM was used to calculate the blade root shears 
for a forward speed of 100 knots both without and 
with the pendulum absorbers. The analytical model 
of the blade used in this step was a ten station, 
fully coupled representation with aerodynamic ex- 
citation forces obtained from flight measured 
pressure distributions (Reference 4) . In the second 
phase of the analysis, a 41 degree-of-freedom fuse- 
lage mathematical model, adjusted to agree with 
shake test results, was analyzed using SADSAM to 
obtain the effect of the resulting hub moments on 
the response in the crew compartment. 

Flight Test Results 

The favorable analytical results referred to 
above led to a decision to fabricate an experi- 
mental set of pendulum dynamic absorbers. These 
absorbers, similar to those shown in Figure 2, were 
installed on the flight test 0H-6A helicopter. 
Tests were conducted measuring the vibration level 
in the crew compartment, both without and with the 
vertical-plane dynamic absorbers installed. The 
measured vibration levels at the pilot's seat are 
presented in Figure 4. It can be seen that the ad- 
dition of the vertical-plane vibration absorbers 
reduces the vibration level at the pilot's seat 
approximately in half. The qualitative assessment 
by the pilot was also very favorable. Based on 
these results the decision was made to incorporate 
vertical-plane dynamic absorbers in the production 
0H-6A helicopter. 



221 



VIBRATION 

VELOCITY, 

IN./SEC 



2.0 

1.8 

1.6 

1.4 

1.2 

1.0 

.8 

.6 

.4 

.2 



I 1 II 1 














WITHOUT VIBRATION 
















ABSORBERS 


























































































































V 


HTH VIBRATION 
ABSORBERS 
































1 1 1 





70 



80 



90 100 110 
Vi, KNOTS 



120 130 



Figure 4. Measured Vibration Level of 0H-6A 

Without and With Pendulum Absorbers 

Operational Experience on 0H-6A 

The vertical plane pendulum absorbers were 
incorporated on all production 0H-6A helicopters 
and on its commercial counterpart, the Model 500. 
Over 3,000,000 flight hours have been accumulated. 
Up to a service life of between 300 and 600 hours, 
the absorbers did a good job of controlling the 
vibration level of the helicopter. However, after 
approximately 300 to 600 hours of service, the 
bearings and shafts on which the absorbers are 
mounted exhibited excessive wear, resulting in in- 
creased vibration level in the helicopter. Re- 
placement of the bearings and shafts generally 
returned the helicopter to an acceptable level of 
vibration. The premature wearing of the bearings 
and shafts was attributed to the high PV value. 

Laboratory tests were conducted on various 
combinations of bearings and shaft types with the 
objective of selecting a combination that would 
have the desired service life of 1200 hours. It 
was also required that any new shaft and/or bearing 
materials be interchangeable with the initial pro- 
duction bearings and shafts. Thus no change in 
geometry was permitted. 

The results of these laboratory tests showed 
that all combinations of shafts and bearings tests, 
with the exception of one, were inferior to the 
original configuration (which consisted of a 
bearing consisting of a stainless steel outer race 
with a bonded self-lubricating teflon liner, and a 
stainless steel shaft with an 8 RMS finish) . The 
only improved configuration consisted of an Astro 
AM1282 bearing, which was specially made for the 
laboratory test operating on the original shaft. 
This Astro bearing is currently under consideration 
for retrofit. 

Conclusions 

This paper has demonstrated both analytically 
and by operational experience that the use of pen- 
dulum dynamic absorbers, mounted on the blade root 
and operating in the vertical plane, can success- 
fully reduce helicopter vibratory loads. The 
specific application on an 0H-6A helicopter was a 



4-bladed rotor with the pendulums tuned to 3/rev 
and 5/rev. The pendulums reduced the vibration 
level in the cockpit to approximately one half of 
the level that existed prior to the installation of 
the pendulums. 

References 

1. Gessow, Alfred and Myers, Garry C, "Aerodyna- 
mics of the Helicopter," The MacMillan 
Company, New York, 1952. 

2. Den Hartog, J. P., "Mechanical Vibrations," 
Fourth Edition, McGraw-Hill Book Co., New York, 
1956. 

3. Peterson, L., "SADSAM User's Manual," The 
MacNeal-Schwendler Corp., 7442 N. Figueroa St., 
Los Angeles, CA, Report MSR-10, December 1970. 

4. Scheiman, James, "A Tabulation of Helicopter 
Rotor Blade Differential Pressures, Stresses, 
and Motions as Measured in Flight," NASA 
TMX-952, March 1964. 

Acknowledgment 

The contribution of R. A. Wagner and other 
Hughes personnel to the development of the vertical- 
plane pendulum absorbers is hereby acknowledged. 



222 



EVALUATION OF A STALL-FLUTTER SPRIIG-DAMPER 
PUSHHOD IN THE ROTATING CONTROL SYSTEM OF A 
■CH-5^B HELICOPTER 

William E. Nettles 
U.S. Army Air Mobility Research & Development Lab., 
Eustis Directorate, Ft. Eustis, Va. 

William F. Paul and David 0. Adams 

Sikorsky Aircraft, Division of United Aircraft Corp. 

Stratford, Conn. 



Abstract 

This paper presents results of a design 
and flight test program conducted to define the 
effect of rotating pushrod damping on stall- 
flutter induced control loads . The CH-5to hell- 
copter, was chosen as the test aircraft because 
it, exhibited stall-induced control loads . Damp- 
ing was introduced into the CH-5^B control system 
by replacing the standard pushrod with spring- 
damper assemblies . 

Design features of the spring-damper 
are described and the results of a dynamic 
analysis is shown which defined the pushrod stiff- 
ness and damping requirements. Flight test 
measurements taken at ^7,000 lb gross weight with 
and without the damper are presented. 

The results indicate that the spring- 
damper pushrods reduced high-frequency, stall- 
induced rotating control loads by almost 50?. 
Fixed system control loads were reduced by k0% . 
Handling qualities in stall were unchanged, as 
expected. 

The program proved that stall-induced 
high-frequency control loads can be reduced 
significantly by providing a rotating system 
spring-damper. However, further studies and 
tests are needed to define the independent 
contribution of damping and stiffness to the 
overall reduction in control loads. Furthermore, 
the effects of the spring-damper should be 
evaluated over a range of higher speeds and with 
lower-twist blades . 



A0B 
CAS 
C 

C M 

c/c„ 



Notation 

angle of bank 

calibrated airspeed, kt 

damping rate, lb-sec/in. 

blade section pitching moment 
coefficient 

damping ratio 



ERITS equivalent retreating indicated 
tip speed, kt. 



GW 



aircraft gross weight 



Presented at the AHS /MSA- Ames Specialists' 
Meeting on Rotorcraft Dynamics, February 13-15 > 
197 1 *. 



I 
K 



75 



(ll/iJ 



torsional moment of inertia 

spring constant 

damper spring rate, lb/in. 

rotor speed 

blade section angle of attack 

blade angle at 75$ rotor radius 

torsional natural frequency, cycles/sec 

ratio of natural frequency to rotor 
frequency 



Introduction 

Control system loads can limit the 
forward speed and maneuvering capability of high 
performance helicopters. The slope of the con- 
trol load buildup is often so steep {Figure l) 
that it represents a fundamental aeroelastic 
limit of the rotor system. This limit cannot be 
removed by strengthening the entire control 
system without incurring unacceptable weight 
penalties . 



Control 
System 
Vibratory 
Load 



Control System 




Endurance Limit 



Control System 
Airspeed Limit 

Stall Region 



Figure 1. 



Airspeed 
"ontrol Load Characteristic 



Studies of the problem reported in 
Reference 1-7 indicate that the abrupt increase 
in control loads is induced by high-frequency 
stall-induced dynamic loading. This loading 
is attributable to a stall-flutter phenomenon 
which occurs primarily on the retreating side 
of the rotor disc in high advance ratio and/or 
high load factor flight regimes. At the 
relatively high retreating blade angles of 
attack which occur under these conditions, the 
blade section experiences unsteady aerodynamic 



223 



stall and the moment coefficient varies with the 
time-varying angle of attack as shown in Figure 
2. Inspection of the moment hysteresis loops 
exhibited in this figure indicate that positive 
work can be done on the system as the blade 
section oscillates in torsion. This aeroelastie 
mechanism, by which energy is added to' the system, 
can be termed "negative damping" and produces 
pitch oscillations of increasing amplitude at the 
blade/control system natural frequency. The 
rotor system is therefore more responsive to 
rotor loading harmonics which are close to the 
blade torsional frequency, and the end result is 
a rapid buildup of higher harmonic control loads 
during maneuvers and high-speed flight. 



was available to the program. Rotating pushrod 
dampers were used instead of fixed system dampers 
because they provided the required damping 
directly at the blade attachment. The program 
was limited in scope to an analytical and 
experimental feasibility study of the concept, 
and was conducted in four phases . 

(1) Dynamic Analysis 

(2) Functional Design 

(3) Ground Tests 

(k) Flight Test Evaluation 



Blade 
Pitching 
Moment, 0- 

C M 



Damping Area 



Pitch Down 




Constant oc 



Postive Work or 
"negative Damping" Area- 



Figure 2. Pitching Moment Hysteresis Loops. 

The response of the rotor system is 
usually stable, because the blades are moving 
into and out of the negative damping region once 
per revolution. However, during maneuvers in 
which a significant portion of the rotor disc is 
deeply stalled, very large oscillations can exist 
(Reference 7) and the negative damping region can 
increase to a point where blade oscillations can 
continue into the advancing portion of the rotor 
disc . 

Efforts to understand the problem have 
centered on defining unsteady aerodynamic 
characteristics of the blades in stall (References 
h and 6) and on incorporating this data into 
blade aeroelastie computer analyses (References 6 
and 9). Results of the studies are encouraging. 
The buildup of control loads and high-frequency 
stall-induced loads is predicted with reasonable 
accuracy . 

Recognizing that the basic cause of the 
problem was insufficient pitch damping, the 
Eustis Directorate contracted with Sikorsky 
Aircraft to evaluate the effects of pushrod 
spring-dampers on control loads of the CH-5UB 
helicopter. This helicopter was selected for the 
study since it exhibited high-frequency stall- 
induced control loads during maneuvers at 
maximum speeds and U8,000 pounds gross weight and 



Dynamic Analysis 

An aeroelastie analysis of the CH-5te 
rotor was performed to evaluate the effectiveness 
of spring-dampers in reducing the control loads 
associated with retreating blade stall-flutter 
and to evolve design criteria. The primary 
mathematical analysis used was the Normal Modes 
Rotor Aeroelastie Analysis Y200 Computer Program. 
This analysis, which is described in Reference 8, 
represents blade flatwise, edgewise, and torsion- 
al elastic deformation by a summation of normal 
mode responses and performs a time-wise integra- 
tion of the modal equations of motion. This 
analysis can also be used to study blade transient 
response following a control input or disturbance. 
Aerodynamic blade loading is determined from air- 
foil data tabulated as a function of blade 
section angle of attack, Mach number, and first 
and second time derivatives of angle of attack. 
Unsteady aerodynamics and a nondistorted helical 
wake inflow were used throughout this investiga- 
tion. 

The version of the 1200 Program used for 
this study is a single-blade, fixed-hub analysis. 
The assumptions were made that all blades are 
identical and encounter the same loads at given 
azimuthal and radial positions and that blade 
forces and moments do not cause hub motion. Any 
phenomena which are related to nonuniformity 
between blades or to the effect of hub motion on 
blade response are not described by this analysis . 

Free Vibration Characteristics 

For a blade restrained at the root by a 
pushrod, the first step in the aeroelastie 
analysis is the calculation of the undamped 
natural frequencies and modes for a blade 
rotating in a vacuum. In order to analyze the 
spring-damper/blade system using the normal modes 
procedure, the damped free vibration modes and 
frequencies were calculated based on the model 
shown in Figure 3. The torsional system was 
represented by fifteen elastically-connected 
lumped inertias restrained in torsion by a spring- 
damper at the blade root. The eigenvalues and 
eigenvectors of the system response were calcu- 
lated using a Lagrangian formulation of the 
damped free vibration equations . A radial mode 



224 



shape, natural frequency and modal damping were 
calculated and used in the Y200 Program. 



Rotor 




Spr ing-Damper 



Control System 



Figure 3. Schematic of the Spring-Damper Free 
Vibration Problem. 



Spring-Damper Behavior 

The behavior of the CH-5^B spring- 
damper was determined by employing the free 
vibration analysis to determine the general 
relationship between the properties of the damper 
itself and those of the blade first torsional 
mode. Figure k shows the variation of blade first 
torsional natural frequency and percent critical 
damping with changes in the spring and damping 
constants of the spring-damper. 




20 1*0 60 80 100 120 lUO 
Spring-Damper Damping Constant, C , lb-sec/in. 



Figure k. Effect of Spring-Damper Properties on First 
Torsional Mode Frequency and Damping. 

Three trends are evident from this figure: 

1. For a given damper spring constant, K D , 
high levels of damping can increase 
the root dynamic stiffness enough to 
result in torsional natural frequencies 
which are close to those obtained with 
a rigid pushrod. It is clear from 



Figure k that as the damping constant, 
Cjj, is increased, the damper spring is 
effectively bridged so that the 
torsional natural frequency approaches 
the standard pushrod value (7.^ per rev.) 

2. For each spring constant, Kj), a 
specified value of the damping constant, 
Cp, maximizes the modal damping. 
Increasing or decreasing the damping 
constant decreased the percent critical 
damping ratio of the torsional vibra- 
tion. 

3. The variation in the percent critical 
damping parameter with damping constant 
is relatively gradual, so small manu- 
facturing differences between the six 
production dampers will not cause great 
differences in first torsional mode 
damping . 

Rotor System Analysis 

For the initial analytical comparison of 
the control system loading with and without damp- 
ing, prior to design of actual hardware, a repre- 
sentative flight condition was selected for which 
experimental data existed for the conventional 
system. This data was extracted from the 
structural substantiation flight tests of the 
CH-5UB and represents a condition in which stall- 
induced dynamic loading was experienced. The 
specific flight condition used - gross weight 
1+7,000 lb, 100$ Rotor Speed (l85 RPM), sea level 
standard, 30° angle of bank right turn- was 
selected because it was the condition which 
consistently produced stall-induced high-frequency 
loading. The plot of rotating pushrod load 
against azimuth for this condition is shown in 
Figure 5a. 

The pushrod load resulting from the Y200 
Wormal Modes Program for the same flight condition 
is compared with flight test results in Figure 5b. 
To account for the increase in rotor lift ex- 
perienced in the turn, a lift of about 60,000 lb 
and a propulsive force of 3,300 lb was calculated. 
Although the calculated pushrod load shows a 
significantly greater steady nose-down load, the 
vibratory amplitude and frequency content of the 
analytical result match the test reasonably well. 

To study the effectiveness of the 
spring-damper in reducing vibratory control loads, 
the flight condition described above was simu- 
lated using several spring-damper configurations . 
Each of these cases was run with the same control 
settings as the standard case. The results are 
shown in Figure 6. As shown, the combination of 
5000 lb/in. and damping between 50 and 90 lb-sec/ 
in. was about optimum. Referring back to Figure 
k, it is seen that a damping value of 90 lb-sec/ 
in. would provide a frequency of 7P which was the 
same as the standard aircraft. This configura- 
tion was therefore selected because the test 
results could then be used to evaluate the spring- 
damper at the same torsional frequency as the 



225 



standard aircraft. Also it would provide an 
option to reduce the damping in follow-on 
programs to allow an evaluation at 5-5/rev and 
20$- critical damping. 
+ 3000 



+ 2000 

+ 1000 



-1000 

-2000 

-3000 

































/ 


s 
































/ 


V 


























/ 


\ 




/ 
























/ 


A 


/ 


\ 




^ 
























,/ 








/ 


±30001* 




























\ 


/ 


\ 


t 





































result, the overall peak-to-peak control load is 
reduced by only 25%, while the high-frequency 
retreating blade control loads are reduced by 
more than 50%. It is these high-frequency loads 
that cause the 6 per rev control system loads in 
the fixed system. 
1000 



a! 

a 

o 

J! 



kO 80 120 160 200 2l+0 280 320 360 

Azimuth, Degrees 
Figure 5a. Measured Flight Test Result. 




-1000 
-2000 
-3000 
-1+000 
-5000 





/I 


^ 


\ 






















! 


f n 




\ 




\ 


/-* 


v 


/ 


^ 














/ 


f 


, 








\ 


/ 


\ 










j 


\ 


' 




J 














\ 




\ 


1 


\ 


/ 


\ 




1 


















J 


) 


J 


\ 


/ 




I 
































,/ 


Ar 
















- 


3xlh 


3 ] 


.b->-- 









1+0 80 120 160 200 2l+0 280 320 3.60 






t3 

o 

to 
& 



<• 1000 



-1000 

-2000 

-3000 

-l+ooo 

-5000 



z^ v ^ w . 


t ^vA «- rtl 


_j V i . A if 


a 1ii_/\7j _.tt 


3 v \z H :1 t 


u -i 


± 3100 lb-*— *—*■- 






o 



Azimuth, Degrees 
Figure 7a. Conventional Pushrod. 



t- 1000 



-1000 

-2000 

-3000 







A 


\ 












1 1 1 










r 




\ 


-^ 


* 


/ 


\ 










} 










\ 


/ 


\ 












Si 




J 














\ 


rs _^^g: 


V/ N 


















\A 










\J 


' 





1+0 80 120 160 200 2l+0 280 320 360 



1+0 80 120 160 200 2l+0 280 320 360 

Azimuth, Degrees 
Figure 5b. Derived Result. 



Figure 7b. 



Azimuth, Degrees 
Stall-Flutter Spring- Damper, 



K„ 



5000 lb/in., C D = 90 lb-sec/in. 



Figure 5- Comparison of Measured and Derived 
Conventional Pushrod Load - CHS'tB, 
1+7000 lb G.W., Sea Level, 100 KT, 30 c 
A0B Eight Turn. 



1+000 



$ 3000' 



u 

& 



1 



CD 



2000 



1000 




Spring-Damper Damping Constant, CD, lb-sec/in. 

Figure 6. Effect of Spring-Damper Parameters 

on the Amplitude of Vibratory Control 
Loads 

The plots of pushrod load against 
azimuth shown in Figure 7 compare a standard 
pushrod with a spring-damper having a spring rate 
of 5,000 lb/in. and a damping rate of 90 lb-sec/ 
in. For this configuration the free vibration 
analysis gives a torsional frequency of 7 per rev 
and 0.20 critical damping ratio. The Figure 
shows approximately equal amounts of one-per-rev 
variation occurring in the control load time- 
histories since the pushrod spring-dampers do not 
affect the low-frequency torsional motion. As a 



Figure 7 .Comparison of Derived Conventional Pushrod 

Load and Spring-Damper Load - CH-5I+B, 1+7000 lb 
G.W., Sea Level, 100 KT, 30° A0B Right Turn. 

It is clear from this analysis that 
(l) damping at the blade root is effective in 
reducing control loads for a given root stiff- 
ness and (2) reducing root stiffness tends to 
decrease the loads for a given damping constant 
(at least for the ranges investigated). 

Functional Design 



Design Requirements 

The aeroelastic analysis indicated 
that spring and damping introduced at the blade 
root could significantly reduce stall-induced 
loads. The most favorable location for the test 
of a blade root spring-damper is at, the pushrod 
connecting the rotating swashplate to the blade 
horn, since the existing pushrod may be replaced 
easily with the spring-damper. It was determined 
that a spring-damper device could be fabricated 
to replace the conventional pushrod, provided 
that the restrictive size limitations could be 
met. The use of an elastomer as the primary 
structural member met the size and spring rate 
requirements . 

The design requirements, based on the 
aeroelastic analysis and the planned test 
programs, are summarized as follows: 

Replace Conventional Pushrod 

Life - 50 hr 



226 



Load - ±5,000 lb 

Spring Rate - 5.000 lb/in. 

Damping Rate - 90 lb-sec/in. 

Maximum Elastic Deflection - ±1/2 in. 

Adjustable for Blade Tracking 

Fail-Safe Design 

Principles of Operation 

The final configuration of the stall- 
flutter spring-damper pushrod designed to meet 
the above requirements is shown in Figures 8 and 
9. 



Orifice Slot 




5.1 




Figure 8. Stall-Flutter 
Spring-Damper Pushrod Assembly. 



Figure 9. Stall-Flutter Spring-Damper Pushrod. 



The concept consists basically of a piston 
restrained in a cylinder by two natural rubber 
elastomeric bushings which provide the required 
spring rate. Damping is obtained by displacement 
of fluid through orifices . The bushings are 
mounted in parallel, thereby providing a fail-safe 
design. In addition, physical stops are incorpor- 
ated to limit spring-damper deflection to ± 1/2 
inch in the event of overload or complete rubber 
failure. Ho sliding action takes place as the 
spring-damper is deflected. Elastomeric elements 
were chosen because of their high allowable 



227 



strains, integral hydraulic sealing, and compact- 
ness. An integral air-oil accumulator was found 
to "be inadequate and an external accumulator 
system was used in the ground and flight tests . 



Ground Tests 

A comprehensive ground test program was 
conducted to develop the required performance of 
the spring-damper, to demonstrate structural 
adequacy and safety for the flight tests, and to 
evaluate the performance of an installed spring- 
damper system. This was accomplished by the means 
of single unit dynamic performance and fatigue 
tests, flight unit proof and operation tests, and 
an installed system whirl tests utilizing the 
flight test spring-dampers and rotor blades. 



Flight Test Evaluation 

The performance of the stall-flutter 
spring-damper pushrod system installed on a CH-5^B 
helicopter was evaluated in a series of flight 
tests consisting of: (l) base-line flights of 
the CH-5te helicopter in standard configuration, 
and (2) comparison flights with the spring-damper 
system installed. 

The investigation was limited to the 
feasibility of the damper and did not extend to 
an extensive evaluation of the overall effect on 
the CH-5to operating envelope. 

Baseline Flights 

A short series of baseline flights was 
conducted on the instrumented test aircraft in 
standard configuration in order to obtain up-to- 
date performance and control load data. 

Of the several conditions flown, the 
115 kt, 96% rotor speed, level flight point was 
the best stall condition from the standpoint of 
uniformity and repeatability. The maximum pushrod 
vibratory load observed was about ± 2,100 lb. 
This is lower than some stall results observed in 
the past on this aircraft, but the typical stall- 
flutter characteristic was observed in the push- 
rod time histories and was therefore adequate for 
baseline purposes. 

Spring-Damper Pushrod Tests 

The spring-damper pushrods were in- 
stalled on the CH-55b rotor head as shown in 
Figure 10 and 11. Flight test time histories of 
rotating pushrod load for rigid pushrods and for 
the spring-damper pushrods at 1*7,000 lb gross 
weight are shown in Figures 12 and 13. These 
segments of data which depict the time history for 
approximately 1-1/2 revolutions were selected as 
representative samples from oscillograph traces 
in which the waveform was continuously repeated 
for more than 15 revolutions. 




Figure 10. Spring-Damper System Flight 
Aircraft Installation. 




Figure 11. First Flight of the Spring-Damper 
System, February 6, 1973. 



228 



Tension 



Pushrod 
Load, lb 




-Spr ing-Damper 



Rigid Pushrod- 



90 



180 



90 



270 360 
Blade Azimuth, Degrees 



180 



Figure 12. Rotating Pushrod Load Comparison 

110 KT 96% N„ Level Plight, 1*700- lb. 
a 



a) 

3 

o 



60 

« 

-P 
0) 
•P 
O 
K 

S-i 
O 

IS 

U 

•H 
> 




I Spring-Damper Pushrod 
I Rigid Pushrod 



1 2 3 It 5 6 7 8 9 10 



Tension 



Pushrod 
Load, lb 




Harmonic Frequency, Per Rev 



Figure ih. Comparison of Spectral Analyses - 

CH-5to, 1*7000 lb G.W., 115 KT 96$ H 
Level Flight, 2000' Altitude. 

Comparison of Stationary Control Loads 

Flight test time-histories of right 
lateral stationary star load for rigid pushrods 
and for spring-damper pushrods are shown in Figure 
15. These records show the expected dominance of 
the 6 per rev response in a 6-bladed rotor. As 
shown, stationary control loads were reduced by 
k0% for the spring-damper case. 



Blade Azimuth, Degrees 



Figure 13. Rotating Pushroa Load Comparison 

115 KT 96% H E Level Flight, 1*7000 lb GW. 



As shown, the rigid pushrod record ex- 
hibits the high-frequency oscillation beginning on 
the retreating side which is characteristic of the 
stall-flutter phenomenon. This frequency was 
between 7 and 8 per rev and compares well with the 
calculated system torsional natural frequency of 
7.1* per rev. As seen, the high-frequency loads 
were significantly reduced with the spring-damper 
pushrods. The overall reduction was smaller 
because the low-frequency response was not reduced. 
This was expected because the high twist blades 
produce large lp loads and the spring-damper was 
not designed to reduce these loads. As shown, the 
results demonstrate a reduction of almost 50$ in 
high-frequency loads. A spectral analysis of the 
data burst which contains this cycle is shown in 
Figure lit. 



Test Condition; 
1*7,000 lb GW, 115 KT, 
± 3320 lb 



N_, 2000* Alt 

it 



Right 
Lateral 
Stationary 
Star Load 
With Rigid 
Pushrods 



Right 
Lateral 
Stationary- 
Star Load 
With Pushrod 
Spring-Dampers 




Figure 15. Comparison of Stationary Control Loads. 



229 



A plot of stationary control load 
against ERITS (Equivalent Retreating Indicated 
Tip Speed) is shown and defined in Figure 16. 
The sharp increase in load as stall is entered is 
seen to be unchanged by the damper installation, 
but as the aircraft goes deeper into the stall 
region, the loads are reduced. 

CH-S^B Structural Substantiation 

_^^ Flight Test Results 

o Base-line Flight Test Data 
fl - — ~—a Spring-Damper Flight Test Data 



■d 1+000 
o 



o 
u 

13 

o 
o 



3500 



3000 



2500 
2000 



' H 1500 



3 

■p 

m 1000 



o 
■p 

u 



500 

























































°l 




















1° 


, 


» 
















! ** 


U 


















of 

1 


















ft 
/ 1 


i 












|A| 






V o 








—A- 


— -J^ 








C 

















320 



300 



280 



260 



2U0 



Erits-Knots 



Figure 16. Stationary Control Load Against Erits 

Note: Erits - Equivilent Retreating 
Indicated Tip Speed 



Rotating Tip Speed x /Air Density Ratio 



-CAS 



^ 



/ Load Factor x Gross Weight 
37,500 

Comparison of Aircraft Handling Qualities 



The handling qualities of the aircraft 
were unchanged with the spring-dampers installed. 
Pilot's reports state that the aircraft exhibited 
the characteristic increase in vibration, 
difficulty in maintaining airspeed, and forward 
control motion required when approaching a stall 
condition in both the baseline and spring-damper 
flights. The stalled condition of the rotor 
appears unaffected by the installation of the 
spring-damper. Blade stresses and blade motions 
(except for the stall-flutter torsional oscilla- 
tion) are virtually the same in each case. Cock- 
pit vibration levels are unchanged. This was 
expected because the stall was not changed, just 
the local torsional response of the blade was 
changed. 

The effect of the damper on the control 
system can be seen in plots of control positions 
against airspeed (Figure 17). The lateral control 
is unaffected, but as much as 10$ more forward 



longitudinal control is required when flying at 
the 115 kt, 9&% N s reference. stall condition. 

LEGEND 
— «- "^^-^^D Base-Line Flight Data • 



.^& Spring-Damper Flight Data 



100 



t 9 ° 
I 

■2 80 

i 



70 



60 



50 

























fof' 








T7T02 


4& 


r 


fc 


J^t 


J& 

















60 70 80 90 100 110 120 

Calibrated Airspeed, KT 
Figure 17a. 100? Rotor Speed. 



100 



•n 90 



e 



J§ 80 



70 



60 



50 











• 




















2, 


'j© 




-A—" 

























60 



70 



80 



90 



100 



110 120 



Calibrated Airspeed, KT 
Figure 17b. 96% Rotor Speed 

Figure 17 . Longitudinal Control Positions. 

Aeroelastic Analysis of Flight . Test Data 

Following completion of flight testing, 
three additional computer analysis conditions were 
run, using test conditions actually observed in 
the flight tests . The methods used were the same 
as described earlier with the exception that a 
calculated lift higher than* the gross weight 
actually flown was used. The amplitudes of 
pushrod load predicted were much lower than 
observed using the correct lift, and since the 
comparison with and without the spring dampers 
was of primary interest, the calculated lift was 
increased, This shows that improvement in the 



230 



analysis is needed. 

Figure 18 shows pushrod load vs azimuth 
for the 115 Jet, 96% % reference condition for 
conventional pushrods as generated by the aero- 
elastic analysis and as observed in the baseline 
flight. The analysis again shows a good correla- 
tion in wave shape with test result. Based on 
analysis of force-displacement phase shifts seen 
in the flight test results, a damping rate of 70 
lb-sec/in. was determined to be a likely Talue 
actually achieved. Figure 18 also compares the 
analytical result with the flight test result. 



A good correlation in wave shape is 
obtained. However, the sharp reduction in peak- 
to-peak amplitude over the rigid pushrod case as 
predicted by the aeroelastic analysis is again 
not achieved in practice. It should be noted 
that the aeroelastic analysis assumes that all 
blades and spring-dampers are identical, which is 
known not to be case. Difference among spring- 
dampers would at least contribute to the dominant 
one-per-rev component and perhaps the harmonics 
as well. 



Conclusions 



Tension 




It is concluded that: 



Derived Conventional 
Pushrod Load at 
115 KT, 96% N R 

1 (Lif t = 51,925 lb) 



Actual Conventional 
Pushrod Load at 
115 KT, 96% N R 
»(GW = 1*7,000 lb) 



90 180 270 360 90 180 



Tension 




Derived 
Spring-Damper 
Pushrod Load, 
C=70 Ib-sec/in. 
(Lift=l+9,969 Vo) 



Actual Spring-Damper 
Pushrod Load 
(GW=Vf,000 lb) 



90 180 270 360 90 180 
Blade Azimuth, Degrees 

Figure 18. Comparison of Measured and Derived 
Pushrod and Spring-Damper Loads. 



1. 



Stall-flutter spring-damper push- 
rods located in the rotating 
control system effectively reduced 
stall-induced high-frequency 
rotating control loads on the 
CH-5te by almost 50$ and overall 
stationary control loads by more 
than k0%. 

The spring-damper pushrod system 
does not significantly alter the 
performance or handling qualities 
of the CH-5te helicopter. 



Rec ommendations 



The test results were very encouraging, 
but as usual raised more questions than it 
answered. Some of these are stated below: 

1. The combination of a spring and 
damping worked well, but 
quantatively what was the 
contribution of each? 

2. Would lower twist, higher mach 
number and lower frequency provide 
different results? 

3. Would a high-speed aircraft show 
some improvement in performance in 
stall with the spring-damper? 

To help answer these questions, the 
CH-54B rotor system could be installed on an 
H53 helicopter and flown to high speed. Damping, 
torsional frequency, and twist could easily be 
varied to qualify their effects . Plans to 
accomplish this are underway. 



References 

1. Harris, F. D., and Pruyn, R. R., BLADE STALL 
- HALF FACT, HALF FICTI0H, American Helicopter 
Society, 23rd Annual National Forum Proceed- 
ings, AHS Preprint No. 101, May, 1967. 



231 



2. Ham, N. D., ana Garelick, M. S., DYNAMIC 
STALL CONSIDERATIONS IN HELICOPTER ROTORS, 
Journal of the American Helicopter Society , 
Vol. 13, No. 2, April 1968, pp. U°-55. 

3. Ham, N. D., AERODYNAMIC LOADING ON A TWO- 
DIMENSIONAL AIRFOIL DURING DYNAMIC STALL, 
AIAA Journal, Vol. 6, No. 10, October 1968, 
pp 1927-193**. 

k. Liiva, J., et al., TWO-DIMENSIONAL TESTS OP 
AIRFOILS OSCILLATING NEAR STALL, Vol. I, 
Summary and Evaluation of Results, The Boeing 
Company, Vertol Division; USAAVIABS TR 68-13A, 
0. S. Army Aviation Materiel Laboratories, 
Fort Eustis, Virginia, April 1968, AD 670957. 

5. Carta, F. 0., et al. , ANALYTICAL STUDY OF 
HELICOPTER ROTOR STALL FLUTTER, American 
Helicopter Society, 26th Annual National 
Forum, AHS Preprint No. 1+13, June, 1970. 

6. Arcidiacono, P. J., et al., INVESTIGATION OF 
HELICOPTER CONTROL LOADS INDUCED BY STALL 
FLUTTER, United Aircraft Corporation, 
Sikorsky Aircraft Division; USAAVIABS 



Technical Report 70-2, U. S. Army Aviation 
Materiel Laboratories, Fort Eustis, Virginia, 
March 1970, AD 869823. 



Carta, F. 0., and Niebanck, C. F., PREDICTION 
OF ROTOR INSTABILITY AT ffl - TORW © i r "3BS, 
Vol. Ill, Stall Flutter, United Aircraft 
Corporation, Sikorsky Aircraft Division; 
USAAVIABS Technical Report 68-18C, U. S. 
Army Aviation Materiel Laboratories, Fort 
Eustis, Virginia, February 1969, AD 687322., 

Arcidiacono, P. J., STEADY FLIGHT 
DIFFERENTIAL EQUATIONS OF MOTION FOR A 
FLEXIBLE HELICOPTER BLADE WITH CHORDWISE 
MASS UNBALANCE, USAAVIABS TR-68-18A, February 
1969, AD 685860. 

Carta, F. 0., et al., INVESTIGATION OF 
AIRFOIL DYNAMIC STALL AND ITS INFLUENCE ON 
HELICOPTER CONTROL LOADS, USAAVIABS TR72-51, 
Eustis Directorate, U. S. Army Air 
Mobility Research and Development Laboratory, 
Fort Eustis, Virginia, September 1972, 
AD 752917. 



232 



MULTICYCLIC JET-FLAP CONTROL FOR ALLEVIATION OF HELICOPTER 
BLADE STRESSES AND FUSELAGE VIBRATION 

John L. McCloud, III* and Marcel Kretz** 
Ames Research Center, Moffett Field, California 94035 



Abstract 

I Results of wind tunnel tests of a 12-meter- 
diameter rotor utilizing multicyclic jet-flap control 
aef lection are presented. Analyses of these results 
are shown, and experimental transfer functions are 
determined by which optimal control vectors are 
developed. These vectors are calculated to eliminate 
specific harmonic bending stresses, minimize rms 
levels (a measure of the peak-to-peak stresses) , or 
minimize vertical vibratory loads that would be 
transmitted to the fuselage. 

Although the specific results and the ideal 
control vectors presented are for a specific jet-flap 
driven rotor, the method employed for the analyses 
is applicable to similar investigations. A discus- 
sion of possible alternative methods of multicyclic 
control by mechanical flaps or nonpropulsive jet- 
flaps is presented. 



a, b, c, 

b 

c 

ci 

Aci 



CL 

CLR/c 

CXR/a 

Cyr/o 

F l> F 2» F 3 

L 

R 

T 

V 

V *c 

V *s 

X 

Y 

<* s 

6 

6 3p = i tan' 

6 4P = i tan 



Notation 

matrix elements 

number of blades 

chord of blades 

blade section lift coefficient 

increment of blade section lift coeffi- 
cient due to multicyclic jet-flap 
deflection 

rotor average lift coefficient (6Cm/a) 

rotor lift coefficient (L/p(fiR) 2 bcR) 

rotor propulsive force coefficient 
(X/p(SR) 2 bcR) 

rotor side-force coefficient 
CY/p(SR) 2 bcR) 

forces measured below the rotor hub 

rotor lift 

rotor radius 

transfer matrix 

forward flight velocity 

cosine component of the summation of 
forces F for the nth harmonic 

sine component of the summation of 
forces F for the nth harmonic 

rotor propulsive force 

rotor side force 

rotor shaft axis inclination 

jet- flap deflection angle 



-1 



-1 



(<5 /<$ 
3S' 3C 

4s' i»c 



:!} 



azimuth angles for max- 
imum deflection 



blade bending stress (or rotor solidity 

for rotor coefficient definitions) 
air density 
azimuth position 
rotor rotational velocity 



Subscripts 



c 

m 

P 
s 



0, 1, 2, 3 . 
Superscript 



cosine 

variable parts 
primary control 
sine 
.n harmonic number 



transpose of matrix or vector 



(Units are as noted, or such as to produce unitless 
coefficients.) 

Introduction 

To achieve its full potential as the most 
effective VTOL aircraft, the helicopter must dras- 
tically reduce its characteristic vibrations and 
attendant high maintenance costs. As shown in 
Reference 1, helicopter maintenance costs are twice 
those of fixed-wing aircraft of the same empty 
weight. With the same basic elements — engines, 
gear boxes, pumps, propellers, and avionics equip- 
ment — in both aircraft, this difference is 
assuredly traceable to the high vibration environ- 
ment of helicopter components. Coping with this 
environment, helicopter designers are forced to 
provide heavier systems, which result in higher 
ratios of empty weight to payload. These ratios 
combine to yield maintenance costs per unit payload 
that are greater than twice those of fixed-wing 
aircraft. The relationship between oscillating 
loads — hence vibration — and maintenance costs 
has been dramatically demonstrated and reported in 
Reference 2. As shown in that report, the Sikorsky 
bifilar system reduced rotor-induced vibratory loads 
by 54.3%, which in turn reduced failure rates so 
that 48% fewer replacement parts were required, and 
overall maintenance costs were reduced by 38.5%. 

Many vibration suppression systems are being 
investigated by various groups. These systems are 
characterized as either absorption, isolation, or 
active control. The multicyclic jet-flap control 
is an active control system, which controls or 
modulates the oscillating loads at their source, 
that is, on the blades themselves. That we can 
effectively change the loading distribution of a 
helicopter rotor in forward flight so as to reduce 
cyclic blade stress variations, or to reduce vibra- 
tory loads transmitted to the fuselage, has been 
demonstrated by large-scale wind tunnel tests of 
the Giravions Dorand jet-flap rotor at Ames Research 
Center. The rotor, its design, and performance 
characteristics have been reported on in Refer- 
ences 3 and 4. Its supporting wind tunnel test 
equipment and some of the results of the multicyclic 
load alleviation tests were presented in Reference 5. 
Some of that multicyclic test data will be shown 
herein also. 



*Research Scientist 
Ames Research Center, Moffett Field, Calif. 94035 



**Chief Engineer 

Giravions Dorand, 92150 Suresnes, France 



233 



The main purpose of this paper is to show the 
method used to analyze the multivariable data, and 
how it is possible to develop several "ideal" con- 
trol schedules "or vectors to achieve specific blade 
stress and vibratory load reductions. A simplified 
analysis of the results is presented, indicating 
that multicyclic systems that do not employ propul- 
sive jet-flaps may be feasible. 

Rotor and Test Apparatus 

The Dorand Rotor is two-bladed, with a teetering 
hub and offset blade coning hinges, but no feather- 
ing hinges. The rotor is driven in rotation by a 
jet-flap, of the blown mechanical flap type, on the 
outer 30% of the blade radius. The mechanical flaps 
are deflected by a swash-plate and cam system, which 
provided both collective and harmonic control. 
Swash-plate tilt provided the longitudinal and lat- 
eral control, whereas the cams introduced second, 
third and fourth harmonic variations. The rotor is 
shown, mounted in the NASA-Ames 40- by 80-ft wind 
tunnel, in Figure 1. Further details of the rotor 
and test apparatus are given in References 3, 4, 5, 
and 6. 

Results and Analysis 

The wind tunnel tests, their range and the 
modi operandi, are described in Reference 6. The 
tests simulated forward flight conditions at blade- 
loading coefficients Clr/cf somewhat greater than 
conventional rotors employ. 

Figures 2 and 3 (taken from Reference 5) show 
some typical results from the multicyclic tests. 
Figure 2 shows three sets of jet-flap deflection 
angle and blade-bending stresses with and without 
multicyclic control. Some control distortion is 
affecting the "without multicyclic control" in that 
the deflection is not purely sinusoidal. The basic 
bending stresses are predominantly three per revolu- 
tion (3P), typical for a relatively stiff, heavy 
blade. The peak-to-peak stress reductions are 29, 
21, and 36%. Figure 3 shows the effect of the 
multicyclic control on the forces below the hub in 
the nonrotating system: on the left, traces for 
three vertical force transducers for the condition 
of zero multicyclic control; on the right, traces 
for the same transducers for multicyclic control 
applied. 

These tests produced data for a large number of 
flight conditions and multicyclic deflection com- 
binations. More of these data are presented in 
Reference 6, which includes both time histories and 
harmonic coefficients of blade-bending stress, ver- 
tical forces, and jet-flap deflection. 

Blade-Bending Stresses 

As discussed in Reference 5, the relationships 
between the time histories of jet-flap deflections 
and the resulting blade-bending stresses can be 
expressed by a transfer matrix.* The time histories 



"This method of analysis was first suggested and 
developed by Dr. Jean-Noel Aubrun of Giravions 
Dorand. 



of jetrflap deflection and blade-bending stress are 
both expressed as harmonic series. If the harmonic 
coefficients of the stress variation (Eq. 1) are 
related to the jet-flap deflection harmonic coeffi- / 
cients (Eq. 2), as shown in Eq. 3, they can be 
expressed in the matrix form as in Eq. 4. 



a = o Q + oi cos * + 0i sin * ♦ 02 cos 2^ + 02 sin 2<|» + ... lj 

f 
5 = S + 61 cos ij) + 61 sin i|i + 62. cos 2$ + «2 S sin 2$ + ... (2Q 

if 

"ns - («n,)(«o) * (bn s )(«l c ) * («ns) («I„) •♦ KX^) * - ("%„) (3) 

then 



"0 

°lc 
"Is 




ao b c d • • oo 
ai c oi c c, c d, c • • o l( . o 

a 's b ls c "s dl s " ' O1 s 
an s >>n s % s d„ s • • o nso 


« 


«o 

«n 
1 



(4) 



The last term of Eq. 3 and the last column of the 
transfer matrix represent the harmonics of stress, 
which are due to the flight condition. With the 
column matrices or vectors of the harmonic contents 
of jet -flap deflection and blade stresses known for 
several conditions, computer routines can solve for 
the transfer matrix elements, 

A sample result of this method was shown in 
Reference 5, together with correlation plots showing 
very good agreement between stresses calculated 
using the transfer matrix and measured stresses. 
The matrix, based on 15 flight conditions, showed 
large amounts of interharmonic coupling, particularly 
for the third and fourth harmonics of stress. 

It is apparent from Eq. 4 that it is possible 
to determine multicyclic jet-flap deflection ampli- 
tudes that will eliminate the corresponding higher 
harmonic stress coefficients. These higher harmonic 
stress terms are set to zero and the equation is 
then solved for the required jet- flap deflection 
coefficients. These coefficients will be hereinafter 
called the "ideal harmonic control vector." Refer- 
ence 6 presents some of these control vectors. 

Although the objective of zero higher harmonic 
stresses was achieved, the requisite multicyclic 
jet-flap deflections produced different amounts of 
IP stresses and, in some instances, the peak-to- 
peak stresses were increased. The changes in IP 
stresses imply a change in the rotor's thrust and 
inplane forces. (Note that the ideal harmonic 
control vector as determined in Eq. 4 may be consi- 
dered to be for "fixed stick" conditions as existed 
in the wind tunnel tests.) Therefore, a second 
transfer matrix (Eq. 5) was defined as shown below. 



234 



"o 




°>c 




°»s 




°2 C 




»2 S 


■ 


f»e 




*>3 S 




?»c 




K 





<=0 *«, 

lis. "1. 



to 



"1, 



80 ho *<) 



ClrA> 

Cyr/o 



(5) 



«irms = -CT^TJ-lCT/TpJ S p 



(7) 



Notice that the columns of the transfer matrix 
and the elements of the control vector have been 
rearranged. The first column represents stress 
levels for the condition of zero rotor shaft inclin- 
ation, zero rotor force coefficients, and no jet-flap 
deflections. The second through fourth columns 
represent the changes in stress level due to rotor 
angle of attack and the rotor's force coefficients. 
The remaining columns correspond to stress deriva- 
tives with respect to the multicyclic jet-flap 
deflections. The control vector has been realigned 
to reflect the column changes. Note that the matrix 
elements are no longer defined by Eq. 3, but by 
Eq. 5 itself, and the basic "collective" and "IP 
cyclic" terms have now been replaced by the rotor's 
force coefficients, C^r/o, Cxr/u and Cyr/0 (multi- 
plied by 10 3 for numerical convenience) . This can 
be considered the transfer matrix for "fixed flight" 
conditions. Correlations for this matrix are not 
as good as those for the "fixed stick" conditions, 
probably because of the greater scatter in the force 
data. However, for 30 test conditions, the corre- 
lation is very good, comparable to the 15-test con- 
dition correlation shown in Reference 5. 

The matrix, based on 30 flight conditions, is 
shown in Figure 4. Again, it is possible to deter- 
mine multicyclic jet-flap deflections to produce 
zero higher harmonic stresses. These deflections 
also define an ideal harmonic control vector, this 
time for fixed flight conditions. Although the IP 
stresses may still change, and the peak-to-peak 
stress increase, the rotor's force output is 
unchanged, at least to the accuracy of the basic 
methodology. 

While elimination of a particular harmonic, or 
all higher harmonics of stress, may be beneficial, 
it may be more desirable to reduce other stress 
parameters, such as the root -mean-square, or the 
peak-to-peak values. It is difficult to relate 
peak-to-peak values to the harmonic coefficients, 
and the iterative algorithm necessary to affect 
peak-to-peak minimization would be considerably 
more complex, for example, than one to minimize the 
root-mean- square values. The rms value of the 
variable portion of the stresses will be minimized 
when the sum of the squares of the harmonic coeffi- 
cients is also minimized. This sum is given by 



?(\ 



(6) 



where irms indicates an ideal root-mean- square, and 
the matrices and vectors are defined by partitioning 
Eq. 5, as shown below: 



b c do e f go ho io 



'=0 


c °s 


a 'c 


b 


c 


; d 'c • 




J l 


's 






T P 








Tin 




1. 












id» s . 




ii. 



C Y r/<» 



This product will be minimized when the multicyclic 
deflections are given by 



These ideal vectors have also been calculated 
for the 30 cases with resultant rms reductions 
between 40 and 66%. Figure 5 shows a few of these 
cases, with stress calculated for "zero" multicyclic. 
These stresses have been, in effect, extrapolated, 
whereas the data in Figure 2 were measured. As 
indicated on the figure, the ideal rms control also 
reduced peak-to-peak stresses. For the 30 cases 
investigated, the ideal rms control vectors reduced 
peak-to-peak stresses from 39 to 65%. 

The ideal multicyclic vectors given by Eq. 7 
are a function of the flight condition as defined 
by shaft axis inclination, advance ratio, and the 
rotor's lift, propulsive, and side-force coeffi- 
cients. The elements of these ideal rms control 
vectors have been plotted against propulsive force 
coefficient in Figure 6. Different symbols denote 
the corresponding lift coefficient levels. The 
effects of Clr/o and Cxr/o and shaft axis inclin- 
ation are quite apparent. (The range of side-force 
coefficients was insufficient to deduce its effect.) 
The third and fourth harmonics were quite constant 
in phase; hence, only their amplitudes have been 
plotted. Note that these harmonics do not appear 
sensitive to rotor lift coefficient. 

Transmitted Vibration Forces 

The rotor suspension system for the wind tunnel 
tests incorporated a six-component balance and a 
parallelogram support discussed in References 4 
and 5. The parallelogram support absorbed inplane 
vibratory loads very effectively, so that the verti- 
cal vibratory loads were the only ones of interest. 
These loads are due to thrustwise hub shears in 
combination with the motions of the hub due to the 
parallelogram support. For this two-bladed rotor, 
the transmitted loads contained only even-order 
harmonics as shown in Figure 3. These loads may 
also be related to the harmonics of the jet-flap 
deflection by a transfer matrix, as shown by Eq. 8. 

With this transfer matrix it is possible to 
eliminate the second and fourth harmonics of the 
vertical vibratory loads by the same procedures 
used to eliminate the higher harmonic blade-bending 
stresses if two of the harmonic components of the 
control vector are specified. The resulting 



235 



°0 
V 2„„ 



Po 1o r o 
P2. 12 C r 2 



Cxr/° 
Cvr/o 

«2 S 
5 3 s 



(8) 



where 



V A (Fi 



+ F 2 + F3)n r 



Vn s A (Fi + F 2 + F 3 )n s 

deflection harmonic components would define ideal 
vibration control vectors whose elements would depend 
also on the flight condition. Such vectors have been 
calculated for the third harmonic jet-flap deflec- 
tions set to zero and are shown in Reference 6. 
These vectors (calculated for 12 cases) show the 
second and fourth control components to be constant 
in phase, but they are significantly different in 
phase and magnitude from the ideal stress control 
vectors. As might be expected, the lack of third 
harmonic jet-flap deflection, and a large fourth 
harmonic requirement, resulted in very large third 
harmonic blade stresses, when these ideal vibration 
control vectors were input into Eq. 5. 

When ideal rms (stress) control vectors are 
input into Eq. 8, the vibratory loads sometime 
increase. A sample case is shown in Figure 7. 
Shown are the stress and vibratory loads for "zero" 
multicyclic, the actual multicyclic used in the wind 
tunnel test, and the ideal rms control vector. The 
actual peak-to-peak stress reduction is 39% and the 
ideal stress reduction is 47%. The ideal rms con- 
trol vector increased the vibratory loads 78%, while 
the actual control increased them by only 48%. The 
upper portion of the figure shows the actual and 
ideal multicyclic component amplitudes and phases. 
The actual phases are quite close to the ideal 
phases, but the actual third and fourth harmonics 
are too low. It is also apparent, however, that 
these third and fourth harmonics caused the increase 
in vibratory loads. 

It is apparent from the foregoing that some 
sort of combined matrix is needed to effect reduc- 
tions in both stress and vibratory loads. It would 
not be possible to eliminate all of the harmonic 
components since for this test rotor, we only have 
six elements in the control vector, 62 » $z s 
through 6u . It is possible, however? to eliminate 
six of the response elements. For example, one may 
select both harmonics of vibratory loads and the 
third and fourth sine components of stress — the 
largest of the stress components — and construct a 
transfer matrix such as shown below. The multicyclic 
deflections required are determined by the solution 
of this equation for the condition that V2 C , V2 S , 
Vu c , Vu s , 03 s and ai, s are all equal to zero. The 
remainder of the stress coefficients and Vo can be 
determined from Eqs. 5 and 8 after the multicyclic 
control vector has been evaluated. 



V2 c„ 


V2 c 


P2 C 


'«.. 


V2 s 


P2, 


V 'c„ 




P"< 


V *S0 




p» 


° 3 s„ 


°3, 


»3 


■*•,„ 


°»s- 


a* 



q Zc r2 c 
b3 s c 3s 







1 
Clr/° 

CxR/o 
C YR /o 



(9) 



Of course, other ideal control vectors are also 
possible, and these would depend quite obviously on 
the particular rotor and flap control system and the 
number of blades, etc. The blades' natural frequen- 
cies, the position and extent of the flaps will all 
affect the blade stress transfer matrix. The num- 
ber of blades will have a definite effect on the 
harmonics of blade loads transmitted to the nonrota- 
ting system; hence, the compromise between loads 
and stress control would differ in each case. How- 
ever, the basic method for analysis used herein 
can be applied to any such investigation, experi- 
mental or theoretical. 

Multicyclic Lift Requirements 

The results presented here correspond to a 
specific jet-flap driven rotor. The question arises 
to what extent other circulation control means would 
permit a similar reduction of stress levels in the 
blades and of vibratory loads. Such systems as 
mechanical flaps, servo flaps controlling the twist 
of the blades, low-powered jet-flaps, conventional 
rotor blades having multicyclic control in addition 
to swash-plate control may introduce multicyclic 
lift effects and are, at least conceptually, capable 
of producing some amount of stress and vibration 
alleviation. This capability being assumed, the 
problem then becomes one of degree rather than one 
of nature. The systems differ only by their 
unsteady flow characteristics but have to offer the 
similar capability of producing high frequency lift 
inputs up to at least the fourth harmonic of rotor 
frequency. The remaining question is "How much 
incremental lift is needed?" 

There was no instrumentation on the blades to 
determine the local lift variations, and had there 
been, it would not be possible to determine the 
amount due to the multicyclic jet-flap deflection 
directly. However, knowing the jet-flap deflection 
and the average jet momentum coefficient, it is 
possible to calculate an incremental lift coeffi- 
cient, assuming a nonvariant alpha. This has been 
done for several of the wind tunnel test cases and 
the Aci ranged from ±0.12 to +0.68 for the higher 
harmonic components. Figure 8 shows the variation 
of the local blade element coefficient Aci for an 
ideal rms control vector. The corresponding stress 
reduction projected for this case would be 50%. 
(Note that Acj is approximately ±0.68.) The figure 
shows that the highest lift variation occurs on the 
retreating blade, a fact that proves favorable for 
the jet-flap, whose capability increases in low 
Mach-number flows. 



236 



It is believed that these magnitudes of Aci 
are obtainable with low powered jet-flaps. Assuming 
that somewhat lesser incremental lift variations 
would be necessary for softer conventional rotor 
blades, multicyclic mechanical and/or servo-flap 
control appears feasible. Two study contracts 
underway also support this contention. 

The sensitivity of the blade stresses and 
vibration to multicyclic control and our present 
inability to predict harmonic loading, stresses, and 
and vibration, leads to the desirability of com- 
pletely automating multicyclic control such as would 
be attained by feedback control systems. The Gira- 
vions Dorand firm is engaged in a basic research 
program to develop such a feedback system and early 
results are quite encouraging. 

CONCLUDING REMARKS 

Wind tunnel tests of a jet-flap rotor simulat- 
ing forward flight have shown that it is possible 
to modulate the rotor's loading by means of a multi- 
cyclic control system so that rotor blade stresses 
and vibratory loads transmitted to the fuselage 
can be reduced. A method of analyzing the multi- 
variable problem has been presented and several 
"ideal" control schedules are presented. The sched- 
ules themselves are applicable only to the specific 
jet-flap rotor tested, but the method of determining 
the schedules is applicable to similar systems. It 
was shown that it is not possible to eliminate all 
oscillatory blade-bending and vibratory loads with 
a system such as the test rotor, which had only 
three higher harmonics of azimuthal control. Such 
limited systems can, however, be used to eliminate 
specific selected harmonic component stress and 
vibration responses. 



A simplified estimate of the incremental lift 
coefficient being generated multicyclically by the 
test rotor indicates that similar multicyclic 
mechanical or low-powered jet-flaps could also be 
sucessful in reducing blade stresses or vibratory 
loads . 

References 

1. Aronson, R. B. and Jines, R. H., "Helicopter 
Development Reliability Test Requirements, 
Vol. I - Study Results," USAAMRDL TR 71-18A, 
February 1972. 

2. Veca, A. C, "Vibration Effects on Helicopter 
Reliability and Maintainability," USAAMRDL 

TR 73-11, April 1973. 

3. Evans, William T. and McCloud, John L., Ill, 
"An Analytical Investigation of a Rotor Driven 
and Controlled by a Jet-Flap," NASA TN D-3028. 

4. McCloud, John L., Ill, Evans, William T., and 
Biggers, James C, "Performance Characteristics 
of a Jet-Flap Rotor," in Conference on V/STOL 
and STOL Aircraft , Ames Research Center, 

NASA SP-116, 1966, pp. 29-40. 

5. McCloud, John L., Ill, "Studies of a Large-Scale 
Jet-Flap Rotor in the 40- by 80-Foot Wind Tunnel,' 
presented at Mideast Region Symposium A.H.S. 
Status of Testing and Modeling Techniques for 
V/STOL Aircraft, Philadelphia, PA, October 1972. 

6. Kretz, M., Aubrun, J.-N., Larche, M., "March 1971 
Wind-Tunnel Tests of the Dorand DH 2011 Jet-Flap 
Rotor" NASA CRs 114693 and 114694. 





MULTICYCLIC MULTICYCLIC 



Figure 1. Jet-flap rotor in the Ames 40- by 80-Foot Figure 2. Effect of multicyclic jet-flap deflection 
Wind Tunnel. on blade stresses. 



237 



|— I REV— -| 

IVArW 



|—l REV— ) 



/WV\A- 



WITHOUT MULTICYCLIC 
CONTROL 



yWA/^Af 



WITH MULTICYCLIC 
CONTROL 



Figure 3. Effect of multicyclic jet-flap deflection 
on vertical forces below hub. 



-441 


-36 


287 


-t9 


12 


2 


-230 


-12 


-409 


-16 


660 


42 






1 


1 


2 


-5 


3 





5 


-2 


6 


iO 


1 





-3 





-2 


-13 


7 


1 12 


4 


-4 


2 


6 





1 ° 


14 


-2 


1 


10 


-3 


1 " 5 


-13 


32 


-20 


-18 


18 


1 10 


6 


-15 


50 


-52 


32 


1 " 


-5 


18 


27 


-21 


-20 


1 ° 


7 


-7 


5 


59 


-78 



I 

a s 

d.,/0- 

C XR /er 
C VR rtr 

h c 



8 «s 



(T 45 RADIAL STATION 
30 CASES AT V/flR*.4 



Figure 4. Transfer matrix for fixed flight 
conditions . 




Clr/ctxIO 3 
o 110— 121 

a loo— 109 

O 90—99 

' 4 _ A 80 — 89 
2 L k 70 — 79 

-I L- 



NO FLAG 
ONE FLAG 
TWO FLAGS 



a 5 

-10° 
-12° 
-15° 




Figure 6. Ideal rms vector relations. 



MULTICYCLIC 
DEFLECTION PHASES 



8 2 , 
S3' 



BLADE 
STRESS 




A\A\ 



I/O MULTICYCLIC 

comw. 




ACTUAL MULTICYCLIC IDEAL r m s CONTROL 
MIND TNKEL TEST) FOR STRESSES 



10 


r\ a /^ 






8 


-\A A/ 






6 
2 


: vyv 




P-P REDUCTION 
65% 








10 


f\ a A r* 


r"\ 


A 


t 8 


■\A A/ 


-\ 


J V\^ 


X 6 


l-VW V 






'1 4 
z 2 


- V 




51% 








10 


rx /\ A _. 


r^ 


A / 


8 


^\\\r 


. \ 


^/u J 


6 


- \ v 




\y \y 


4 
2 


V 


- 


47% 



WITHOUT MULTICYCLIC WITH "IDEAL rmS CONTROL 

Figure 5. Calculated blade bending stresses using 
equations 5, 6, and 7. 



Figure 7. Calculated blade stresses and vibratory- 
loads using equations 5, 6, 7 and 8. 



Figure 8. 




90 180 270 360 

AZIMUTH ANGLE, * deg 

Variation of the estimated increment of 
blade section lift coefficient due to 
multicyclic jet-flap deflection. 



238 



IDENTIFICATION OF STRUCTURAL PARAMETERS FROM HELICOPTER DYNAMIC TEST DATA 

Nicholas Giansante 
Research Specialist 

William G. Flannelly 
Senior Staff Engineer 

Kaman Aerospace Corporation 
Bloomfield, Connecticut 



Abstract 

A method is presented for obtaining 
the mass, stiffness, and damping param-' 
eters of a linear mathematical model, 
having fewer degrees of freedom than the 
structure it represents, directly from 
dynamic response measurements on the 
actual helicopter without a priori knowl- 
edge of the physical characteristics of 
the fuselage. The only input information 
required in the formulation is the approx- 
imate natural frequency of each mode and 
mobility data measured proximate to these 
frequencies with sinusoidal force excita- 
tion applied at only one point on the 
vehicle. This dynamic response informa- 
tion acquired from impedance testing of 
the actual structure over the frequency 
range of interest yields the second order 
structurally damped linear equations of 
motion . 

The practicality and numerical sound- 
ness of the theoretical development was 
demonstrated through a computer simulation 
of an experimental program. It was shown, 
through approximately 400 computer ex- 
periments, that accurate system identifi- 
cation can be achieved with presently 
available measurement techniques and 
equipment . 



Notation 



C 
d 
f 
f 

g 

i 
J 
K 
m 



Presented at the AHS/NASA-Ames Special- 
ists' Meeting on Rotorcraft Dynamics, 
February 13-15, 1974. 



N 
P 
Q 
R 
S 
Y 

n 
[*] 

Subscripts 

i 

J, k 

( ) 



number of degrees of freedom 

number of forcing frequencies 

number of modes 

residual 

modal mobility ratio 

displacement mobility, 3y/3f 

natural frequency 

matrix of modal vectors 

modal index 

degree of freedom index, 
generalized coordinate index 

a subscripted index in 
parentheses means the index 
is held constant 



Superscripts 
(q) q-th iteration 
* modal parameter 
R real 



influence coefficient 


I 


damping 


T 


force 


-1 


force phasor 


-T 


structural damping coefficient 


+ 


imaginary operator (i = /^T) 


Brackets 


number of generalized coordinates 


[ ], ( ) 


stiffness 


r J 


mass 


{ } 



imaginary 

transpose 

inverse 

transpose of the inverse 

pseudoinverse , generalized 
inverse 

matrix 

diagonal matrix 
column or row vector 



239 



capital letters under matrices indicate 
the number of rows and columns, 
respectively 

a dot over a quantity indicates differen- 
tiation with respect to time 

The success of a helicopter struc- 
tural design is highly dependent on the 
ability to predict and control the 
dynamic response of the fuselage and 
mechanical components. Conventionally/ 
this involves the formulation of intu- 
itively based equations of motion. 
Ideally, this process would reduce the 
physical structure to an analytical 
mathematical model which would predict 
accurately the dynamic response character- 
istics of the actual structure. 
Obviously, the creation of such an 
intuitive abstraction of a complicated 
real structure requires considerable 
expertise and inherently includes a high 
degree of uncertainty. Structural 
dynamic testing is required to substan- 
tiate the analytical results and the 
analysis is modified until successful 
correlation is obtained between the 
analytical predictions and the test 
results . 

Until a prototype vehicle is avail- 
able, intuitive methods are the only 
choice for describing an analytical model. 
However, once the helicopter is built, 
the method of structural dynamic testing 
using impedance techniques can be used to 
define directly a dynamic model which 
correlates with the test data. Such a 
model, synthesized from test data, 
succeeds in unifying theory and test, 
minimizing the intuitive foundation of 
conventional analyses. 

System Identification has been de- 
fined as the process of obtaining the 
linear equations of motion of a structure 
directly from test data. In System 
Identification the objective is the ex- 
traction of the mass, stiffness and 
damping parameters of a simple mathemati- 
cal model directly from dynamic response 
measurements on the actual helicopter 
without a priori knowledge of the physical 
characteristics of the fuselage. Figure 1 
presents a pictorial representation of 
the System Identification process. 

This paper describes the theory of 
System Identification using impedance 
techniques as applied to a mathematical 
model having fewer degrees of freedom 
than the structure it represents. The 
method yields the mass, stiffness and 
damping characteristics of the structure, 
the influence coefficient matrix, the 
orthogonal modes, the exact natural 
frequencies, the generalized parameters 



associated with each mode and dynamic 
response fidelity over the frequency range 
of interest. The only information nec- 
essary to implement the method is the 
approximate natural frequency of each mode 
and mobility data measured proximate to 
these frequencies with the excitation 
applied at a single point on the vehicle. 
This data can be readily obtained from 
impedance type testing of the helicopter 
over the frequency spectrum of interest. 




TBE0B&TZOU. 

DEVELOPMENT 



-MBH*M.H 




Figure 1. System Identification 
Process 

The usefulness and numerical sound- 
ness of the theoretical development was 
demonstrated through a computer simulation 
of an experimental program, including a 
typical and reasonable degree of measure- 
ment error. To test the sensitivity of 
the method to measurement error, a series 
of computer experiments were conducted 
incorporating typical and reasonable 
degree of measurement error. The results 
indicate that accurate identification of 
structural parameters from dynamic test 
data can be achieved with presently 
available measurement techniques and 
equipment . 



240 



Description of the Theory 

Derivation of the Single Point Iteration 
Process 

As presented in References 1 and 2, 
the mobility of a structure at forcing 
frequency, io, is given by 



[y 3 = r*] ty* j r$] T 



(1) 



With excitation at station k, the respon- 
ses at station j, including k, are 
obtained. These provide the k-th column 
of the mobility at a particular forcing' 
frequency o> 1 : 



t{Y j(k)u 1 HY j(k)a) 2 }1 

= Wr^jUY^HY*^}] 

JxN NxN Nx2 

Generally, for p forcing frequencies 
where 1 < p < P, 



tY j( k)p^ - t*ir*kiJ [Y i P ] 



JxP 



JxN NxN NxP 



(4) 



(5) 



{Y 



j (k)u. 






If J > P, Equation (5) is set of more 
equations than unknowns for which there is 
no solution. In this situation, Equation 
(5) can then be written as 



3y J /3f 1 



[Y. ,.. ] = [*] M>, -J [Y. ] + [R. ] 
3 (k)p J L J lT ki^ L lp u jp 



(6) 



J x Y L/ki { * } i = [ ^ {Y L/ki } (2) 



where R. is the residual associated with 

DP 
the j-th station and the p-th forcing 
frequency . 



where 1 £ j £ J and 1 _< i £ N. 

This represents a column of mobility 
response each element of which is the 
response at a generalized coordinate on 
the structure with excitation at station 
k and at forcing frequency u,. Similarly, 
with the exciter remaining at station k, 
the k-th column of the mobility at 
another frequency, w 2 , can be obtained. 

3 Yl /3f 2 ' 
{Y j( k)u, 2 >= {3 y 2 / 3f 2> 



8yj /3f 2 



N 
= 2 



i=1 Y L 2 w*>i = [$]{Y L 2 *ki } < 3) 



The columns of mobility response 
represented by (2) and (3) may be com- 
bined into one matrix 



As described in References 1 and 2, 
the imaginary displacement mobility is 
usually significantly affected by modes 
associated with natural frequencies in 
proximity to the forcing frequency. 
Reference 3 indicates that accurate 
estimates of the modal vectors may be 
obtained by considering only the effects 
of modes . proximate to the forcing fre- 
quency. Therefore, the analysis will 
employ only Q modes, where Q is less than 
N. The imaginary displacement mobility 
may be expressed as: 



[y; 



**■> 



j(k)p 



] = [•] f» ki J [Y ± ;] + [R. p ] (7) 



DP 



Since each column of [Y. ] is 

associated with a particular frequency, 
the dominant element of each row of the 
matrix will be the modal mobility measured 
at the forcing frequency in proximity to 
a particular natural frequency. Nor- 
malizing the rows of the aforementioned 
matrix on the largest element yields 



I V ■ riA in J [Y Ip 3 



(8) 



where Y. is the maximum value of the i-th 
m 

row. Equation (7) may be rewritten, 

incorporating Equation (8) 



241 



[Y j(k) P i - [♦]W kl *2ns lp i+.[H 3p ] 



(9) 



The matrix Equation (9) has no 
solution, however, an approximation to a 
solution may be defined as that which 
makes the Euclidian norm of the matrix of 
residuals a minimum. The modal vector 
matrix with respect to which the Euclidian 
norm of the residuals is a minimum is ob- 
tained through use of the pseudoinverse , 
and is given by 



w = [Y D T (k)p ] [s . p ] + r -i- 



f ki in 



(10) 



where [S. ] is defined as the generalized 

inverse or pseudoinverse of [S. ] and is 
defined by lp 

[S ip ]+= [S ip ]T([S ip ][S ip ]T)_1 (11) 

In Equation (10) each diagonal element of 

[= jf=J simply multiplies the correspon- 
ds .Y. 
Y ki in 

ding column of the modal matrix. Since 
each modal vector is normalized on the 
largest element in the vector, the effect 
of the aforementioned multiplication is 
negated and Equation (10) can be reduced 
to 

M = [Y j(k)pHs ip ] + (12) 



The [S] matrix can be accurately 
estimated from knowledge of only the 
forcing frequencies and the natural fre- 
quencies. Equation (12) will be solved 
utilizing matrix iteration techniques. 
At each successive iteration a solution 
is found that minimizes the Euclidian 
norm of the residual matrix with respect 
to the newly found matrix of either [S] 
or [<{>]. The basic algorithm used in the 
matrix iteration procedure for the q-th 
iteration becomes 

H> (g) ] = tY I ][s (q " 1, ] + 



and 



[S (< J } ] = 



[<f> (g) ] [Y X ] 



(13) 



Determining the Modal Parameters 



The real modal impedance at forcing 
frequency u can be written as 



5 iw 



Y *R 
id) 

5-E ■ - ■ ■■ 

*-d 2 *_ 2 

(Y. ) + (Y.; ) 

10) 10) ' 

p ' p 



(14) 



Substituting the real and imaginary dis- 
placement mobility as given in Reference 1 
yields 



\l =K.(l-a>X 2 ) 
p F 



(15) 



From Equation (15) it is observed > 
that the modal impedance is a linear 
function of the square of the forcing 
frequency. The forcing frequency at which 
the modal impedance becomes zero is, 
therefore, the natural frequency. From a 
least squares analysis of modal impedance 
as a function of forcing frequency 
squared, proximate to the natural fre- 
quency, the generalized stiffness of the 
i-th mode and the natural frequency of the 
i-th mode can be calculated. 

The generalized mass associated with 
the i-th mode is given by 



it it O 

m i - vv 



(16) 



The structural damping coefficient may be 
determined from 



*i = 



2 

(-£- 

1 2 

a* 



i) 



Y. 

10) 



(17) 



Models 



There are two basic types of dynamic 
mathematical models describing structures. 
The first type described as "Complete 
Models" considers as many modes as degrees 
of freedom. The second type labelled 
"Truncated Models" considers fewer modes 
than points of interest on the structure. 
Using the methods described herein, it is 
possible to identify either a complete 
model or a truncated model. 

For the completed model the modal 
matrix [4>] is square. However, in the 
case of the truncated model the modal 
matrix [<H is rectangular having J rows 
corresponding to the points of interest 
and Q columns associated with the mode 
shapes, where J > Q. 



242 



Truncated Models 

Consider a rectangular identified 
modal matrix which has J rows indicating 
the points of interest on the structure 
and Q columns representing the modes being 
considered where J > Q. The influence 
coefficient matrix for the truncated model 
is given by 



[C TR ] = [4,]p4-j[0] T 
K i 



(18) 



The above matrix is singular being of rank 
Q and order J. The mass, stiffness and 
damping matrices for the truncated model 
are 

[m^] = [<f>] rm.jj [<j>] 



[K™! = [<j>] +T tK*J[(|)] + 



"TR 



[d TR ] = [(!>] +T rg i K*4f<J>] + (19) 



The classical modal eigenvalue equation 
has the analogous truncated form 



i c ra"»W { *i>-:rr<*i> 



(20) 



Complete Models 

For the complete model the identified 
modal vector matrix is square, having the 
same number of degrees of freedom as mode 
shapes, thus J = Q. The influence matrix 
is given by 

N T 

[C] = [<!>]n/K*J[(i.] T = t 1/K*{(f>. H(j>.} 
1 i=l 1 1 1 

(21) 

The mass, stiffness and damping 
matrices for the complete model are simi- 
lar to those of Equation (19) , except 
that the matrices are square. 



[m] = Wl^muJEt))]" 1 



[k] = m T tt/K*jr<i>r ;L 



[d] = M -1 ^*.]^]"" 1 



(22) 



Full Mobility Matrix 



The full mobility matrix of either 
complete or truncated models is given by 



[Y] = I*] tY*J [<j>] T 
Computer Test Simulation 



(23) 



The usefulness and numerical sound- 
ness of the theoretical development was 
demonstrated through a computer simulation 
of an experimental program. Approximately 
400 computer experiments were performed in 
the study. A twenty-degree-of-freedom 
lumped mass beam type representation of a 
helicopter supported on its main landing 
gear and tail gear was used to generate 
simulated mobility test data. Each of the 
coordinates was allowed a transverse de- 
gree of freedom. The concentrated mass 
and stiffness parameters of the beam are 
shown in Table I, with EI varying 
linearly between stations and with 5 
percent structural damping. 

Simulated Errors 



System Identification theories of any 
practical engineering significance must be 
functional with a reasonable degree of 
experimental error. Therefore, a typical 
and reasonable degree of measurement error 
ranging to +15% random error uniformly 
distributed and 15% bias error, was incor- 
porated in the simulated test data. Both 
random and bias error were applied to the 
real and imaginary components of the dis- 
placement mobility data. The levels of 
error applied are consistent with those 
inherent in the present state-of-the- 
measurement art. 

Models 



The number of degrees of freedom of 
a physical structure is infinite. There- 
fore, the usefulness of model identifica- 
tion, necessarily with a finite number of 
degrees of freedom, using impedance 
testing techniques , depends on the ability 
to simulate the real structure with a 
small mathematical model. 

Several size models , containing from 
5 to 15 degrees of freedom, were synthesi- 
zed from the simulated test data incor- 
porating the specified experimental error. 
Table II describes the various models 
used in the analysis. The model stations 
used in the models refer to the corres- 
ponding stations in the twenty point 
specimen . 

Identified Models 

Typical generalized mass identifica- 
tions are shown in Tables III, IV and V. 
Table III presents results for several 
different five point models. The model 
designations refer to the descriptions 
presented in Table II. Data are also 



243 



presented for the twenty point specimen 
with zero experimental error. Thus, a 
basis of comparison is established with 
the theoretically exact control model of 
the beam representation of the helicopter. 
It is apparent that no outstanding dif- 
ferences exist among the identified 
generalized masses for the models con- 
sidered for comparison. Table IV presents 
similar data for the nine-point models 
studied. The generalized mass distribu- 
tion associated with each of the models 
is in excellent agreement with the twenty 
point model results. 

Table V describes the results of the 
computer experiments conducted employing 
the twelve point models. The results are 
satisfactory except for the identification 
of the generalized masses of the tenth and 
eleventh modes. However, the generalized 
masses associated with these modes are 



extremely small in comparison with the 
remaining modal generalized masses. An 
examination of the tenth "mode shape re- 
vealed a lack of response at all points 
of interest on the structure other than 
the first station. Therefore, the effect 
of the tenth mode is difficult to evaluate 
in the calculation of the generalized 
parameters. Computer experiment 309 
yielded a negative generalized mass for 
the tenth mode. All computer experiments 
that failed in this respect gave dras- 
tically unrealistic values of generalized 
mass. Ordinarily, in a situation where 
the generalized mass was unrealistic, use 
of different stations for the model 
improved the identification. 



Sta No. 



TABLE I. 



20-POINT SPECIMEN DESCRIPTION 



1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 



Sta (In.) 



60 120 160 200 240 280 320 370 430 

30 100 140 180 220 260 300 340 400 460 



Mass 
(Lb-Sec 2 /In. ) 



.029 3.67 2.18 2.385 2.08 .910 .170 

1.05 3.71 2.18 2.59 1.56 .260 .085 



.070 .095 .210 



.060 .120 .150 



|EI , n .35 .35 1.95 4.37 5.80 4.425 3.07 2.05 .975 .55 
(Lb-In7 x 10 10 ) 

.35 1.20 3.00 5.70 5.60 3.6 2.60 1.60 .65 .50 



Springs to 
Ground (Lb/In.) 



10000 



10000 















TABLE 


II. 


MODEL 


DESCRIPTION 




































Stations 


Used 




















Model 


1 


2 


3 


4 


5 


6 


7 


8 


9 


10 


11 


12 


13 


14 


15 


16 


17 


18 


19 


20 


5A 


x 










X 








X 










X 










X 


5B 


X 










X 










X 










X 








X 


5C 


X 










X 








X 










X 








X 




5D 




X 








X 












X 






X 








X 




9A 


X 


X 








X 


X 






X 






X 






X 




X 




X 


9B 


X 




X 






X 




X 






X 






X 






X 




X 


X 


9C 




X 


X 






X 


X 






X 






X 






X 




X 




X 


12A 




X 


X 


X 


X 


X 




X 




X 




X 




X 




X 




X 




X 


12B 


X 


X 


X 




X 


X 




X 




X 




X 




X 




X 




X 




X 


12F 


X 


X 


X 




X 


X 




X 






X 




X 




X 




X 




X 





244 



TABLE III. 


IDENTIFICATION OF 
5X5 MODEL OF 20 


GENERALIZED MASSES, 
X 20 SPECIMEN 




Model 








5A 


5B 


5C 


5D 




j ** 


Computer 

Experiment 

Number 








296 


297 


292 


295 




_ 


Random Disp. 


Error 




+5% 


+5% 


+5% 


+ 5% 







Bias Disp. Error 






+5% 


+5% 


+ 5% 


+ 5% 







Random Error 


Seed 




13 


13 


13 


13 




- 


Mode 










Generalized Masses 
(Lb-Sec 2 /In.) 






1 








8.544 


8.538 


8.543 


8.568 


8 


534 


2 








4.506 


4.506 


4.619 


4.610 


4 


449 


3 








.494 


.494 


.494 


.49 3 




495 


4 








1.048 


1.047 


1.050 


.994 


1 


.087 


5 








.653 


.653 


.651 


.629 




.630 


** From 20 x 


20 


Sp< 


scimen 















TABLE 


IV. 


IDENTIFICATION 
9X9 MODEL OF 


OF GENERALIZED MASSES 
20 X 20 SPECIMEN 


/ 




Model 








9A 


9B 


9C 


20 Pt 


Computer 

Experiment 

Number 






300 


303 


304 


^** 


Random 


Disp. 


Error 


+ 5% 


+ 5% 


+ 5% 





Bias Disp. Error 




+ 5% 


+ 5% 


+ 5% 





Random 


Error 


Seed 




13 


13 


13 


- 


Mode 










Generalized Masses 
(Lb-Sec 2 /In.) 




1 








9.000 


9.015 


9.043 


8.534 


2 








4.350 


4.335 


4.513 


4.449 


3 








.472 


.472 


.472 


.495 


4 








1.042 


1.042 


1.138 


1.087 


5 








.551 


.549 


.584 


.6 30 


6 








.786 


.783 


.723 


.743 


7 








1.154 


1.243 


1.120 


1. 177 


8 








1.401 


1.411 


1.396 


J.412 


9 








.787 


.708 


.791 


.78(> 


** From 20 x 


20 Specimen 











245 



TABLE V. 


IDENTIFICATION OF GENERALIZED MASSES, 
12 X 12 MODEL OF 20 X 20 SPECIMEN 




Model 








12B 


12F 


12A 20 Pt 


Computer 

Experiment 

Number 








312 


311 


309 1** 


Random Disp. 


Error 






+5% 


+ 5% 


+ 5% 


Bias Disp. Error 






+5% 


+5% 


+ 5% 


Random Error 


Seed 






13 


13 


13 


Mode 










Generalize 
(Lb/Sec 


d Masses 
2 /In.) 


1 








8.474 


8.464 


8.518 8.534 


2 








4.556 


4.510 


4.492 4.449 


3 








.488 


.487 


.487 .495 


4 








1.150 


1.151 


1.103 1.087 


5 








.596 


.597 


.595 .630 


6 








.722 


.724 


.777 .744 


7 








1.182 


1.113 


1.159 1.177 


8 








1.232 


1.242 


1.215 1.412 


9 








.797 


.743 


.789 .786 


10 








1.203 


1.043 


-.564 .043 


11 








,09 3 


.104 


.0103 .172 


12 








1.177 


1.119 


1.147 1.050 


** From 20 x 


20 Specimen 







Response From Identified Model 

One of the most essential requisites 
of relating a discrete parameter system 
to a continuous system is model response 
fidelity over a given frequency range of 
interest. The finite degree of freedom 
model must accurately reproduce the dy- 
namic response of the infinite degree of 
freedom structure over a specific number 
of modes. Figures 2a and 2b show typical 
real and imaginary driving point accel- 
eration response respectively for the 
five point model. The "exact" curve 



represents the simulated experimental 
data for the twenty point structure , 
obtained with zero error. The frequency 
range encompasses the first five elastic 
natural frequencies. Figures 3 and 4 
present similar results for typical nine 
and twelve point models, respectively. 
The computer experiments for which results 
are presented incorporated a +5 percent 
random and a +5 percent bias on the real 
and imaginary displacement mobility data. 
As evidenced by the figures, the various 
models yielded satisfactory reidentifica- 
tion of the twenty point specimen simu- 
lated dynamic response data. 



246 




CASS SEED ERROR 

290 . HO 

292 13 YES 

293 421 YES 




CASS SEED ERROR 

290 SO C 

292 12 YES a 
292 421 ITS ' 



Figure 2a. Effect of Error on Five-Point 
Model Identification of Real 
Acceleration Response; Driving 
Point at Hub 



Figure 2b. Effect of Error on Five-Point 
Model Identification of 
Imaginary Acceleration 
Response; Driving Point at Hub 




CAIE 1SSD ERROR 

299 NO o 

300 13 YES a 

301 421 YES A 

CASES WIOT ERR'.R 
±99 RANDOM, St IZAS 
100 



Figure 3a. 



Effect of Error on Nine-Point 
Model Identification of Real 
Acceleration Response; Driving 
Point at Hub 




300 400 soo 



Figure 3a - Continued 




nm ZOSHTZFUD 



«» RANDOM, St EIAS 




CASE SEED ERROR 
305 HO O 

in 13 yes a 

307 421 YES » 



Figure 3b . 



Effect of Error on Nine-Point 
Model Identification of 
Imaginary Acceleration 
Response; Driving Point at Hub 9 



Figure 4a. 



Effect of Error on Twelve- 
Point Model Identification of 
Real Acceleration Response; 
Driving Point at Hub 



247 




200 300 400 500 600 700 




CMC SEED ERROP 

.305 NO o 
ill D YES a 
307 421 YES A 



Figure 4a - Continued 



Figure 4b. Effect of Error on Twelve- 
Point Model Identification 
of Imaginary Acceleration 
Response; Driving Point at 
Hub 



Conclusions 

Single point excitation of a structure 
yields the necessary mobility data to 
satisfactorily determine the mass, 
stiffness and damping characteristics 
for a mathematical, model having less 
degrees of freedom than the linear 
elastic structure it represents . 

The method does not require an in- 
tuitive mathematical model and uses 
only a minimum amount of impedance 
type test data. 

The eigenvector or mode shape 
associated with each natural 
frequency is also determined. 

Computer experiments using simulated 
test data indicate the method is in- 
sensitive to the level of measurement 
error inherent in the state-of-the- 
measurement art. 



References 

1. USAAMRDL Technical Report 70-6A, 
THEORY OF STRUCTURAL DYNAMIC TESTING 
USING IMPEDANCE TECHNIQUES, 
Flannelly, W.G. , Berman, A. and 
Barnsby, R. M. , U. S. Army Air 
Mobility Research and Development 
Laboratory, Fort Eustis, Virginia, 
June 1970. 

2. USAAMRDL Technical Report 72-63A, 
RESEARCH ON STRUCTURAL DYNAMIC 
TESTING BY IMPEDANCE METHODS - PHASE I 
REPORT, Flannelly, W.G. , Berman, A. 
and Giansante, N. , U. S. Army Air 
Mobility Research and Development 
Laboratory, Fort Eustis, Virginia, 
November 1972. 

3. Stahle, C.V. , Jr., PHASE SEPARATION 
TECHNIQUE FOR GROUND VIBRATION 
TESTING, Aerospace Engineering, July 
1962. 



248 



ENGINE/AIRFRAME INTERFACE DYNAMICS EXPERIENCE 

C. Fredrickson 
Senior Engineer 

Boeing Vertol Company 
Philadelphia, Pa. 



Abstract 

Recent experience has highlighted the 
necessity for improved understanding of 
potential engine/airframe interface 
dynamics problems to avoid costly and time- 
consuming development programs. This 
paper gives some examples of such problems, 
and the manner in which they have been 
resolved. It also discusses a recent pro- 
gram in which contractual engine/airframe 
interface agreements have already proven 
helpful in the timely prediction and 
resolution of potential problems. 

In particular, problems of engine/ 
drive system torsional stability, engine 
and output shaft critical speeds, and 
engine vibration at helicopter rotor order 
frequencies are discussed, and test data 
and analyses presented. Also presented is 
a rotor/drive system dynamics problem not 
directly related to the engine. 

General 

This paper is an attempt to highlight 
some recently encountered problems in the 
area of helicopter engine and drive system 
dynamics. In comparison to the number of 
technical papers published in the area of 
rotor and blade aeroelasticity and 
stability, and fuselage vibration reduc- 
tion schemes, there are relatively few 
indeed dealing with engine/airframe 
dynamics . 

The paper does not present highly 
sophisticated methods of solution for 
these problems . It instead shows that 
solutions were attained by the application 
of basic engineering principles to state- 
of-the-art analytical and test techniques. 
Also, having encountered these problems, 
we are more cognizant of these potential 
"show-stoppers," the manner in which they 
manifest themselves, and the available 
courses of corrective action. It is 
essential that the knowledge gained 
through these programs be judiciously 
applied to new helicopters, and growth 
versions of existing models. 

Engine/Drive System Torsional Stability 

The usual stability requirements that 
dictate fuel control gain limits are com- 
plicated by the flexibility of the heli- 
copter drive system and by the dynamics of 



a gas turbine engine. The interaction of 
the helicopter rotor and drive system, 
engine, and fuel control requires careful 
attention if a good or even workable fuel 
control is to be achieved. In the case of 
the T55-L-11 engine and the CH-47C air- 
craft, these items were growth versions of 
existing components. There was no require- 
ment for new control concepts since opera- 
tion had been successful on previous 
models. However, the fuel control gains 
had to be carefully re-evaluated for the 
new power levels. 

Computer simulation of the CH-47C 
rotor system with the T55-L-11 turbine 
engine was accomplished before initial 
flight tests began. The simulation 
indicated favorable engine/control stabil- 
ity. However, as pointed out in Reference 
(1) , unacceptable oscillations in engine 
shaft torque and rotor RPM were observed 
during initial flight tests (Figure (1) ) . 
These torque oscillations were audible, 
disconcerting to the flight crew, and were 
observed only in hover and on the ground 
(not in forward flight) . The frequency of 
the oscillation was also higher than the 
predicted drive system torsional natural 
frequency . 

Since the torsional instability was 
not predicted by the computer simulation, 
a study of pertinent system parameters was 
undertaken. It was discovered that the 
only parametric change having a significant 
effect on torsional stability was the 
slope of the blade lag damper force- 
velocity curve below the preload force 
level. When this curve was artificially 
"stiffened" beyond its actual limits, as 
shown in Figure (2) , the oscillation was 
reproduced. This fact suggested that by 
"softening" the actual damper preload 
slope, the oscillation might be suppressed. 
Once analytically reproduced, the oscilla- 
tion could be eliminated by simulating a 
fuel control with a reduced steady state 
gain and a slowed time constant. The 
computer analysis, therefore, revealed two 
potential solutions to the torsional 
oscillation problem: a lag damper modifi- 
cation and a fuel control modification. 

Flight tests with a set of lag 
dampers with significantly reduced preload 
slope, together with the original fuel 
controls, were conducted. These tests 
revealed that the torque oscillation was 



249 



apparently suppressed. However, since the 
lag dampers were on the aircraft for 
ground resonance reasons, this significant 
load change reduced damping capacity and 
produced some degradation in the ground 
resonance characteristics of the helicop- 
ter. Therefore, damper modification to 
remedy torque oscillation was rejected. 

Fuel controls with a 30% reduction in 
steady-state gain were flight tested, and 
yielded acceptable torsional stability. 
However, this degraded to marginal insta- 
bility in colder ambient temperatures. 
Controls incorporating a gain reduction 
plus an increase in time constant provided 
acceptable engine torque stability in the 
cold and over the entire engine operating 
envelope. Fuel control frequency response 
curves are shown in Figure (3) . Pilots 
also noted that engine response to input 
power demands was not perceptibly degraded 
with these slowed-down controls. There- 
fore, this fuel control modification was 
considered an acceptable production fix. 

Representation of the lag damper with 
just the force-velocity curve in the 
engine/drive system/fuel control simulation 
had been shown to be insufficient to 
accurately reproduce the torque oscilla- 
tion phenomenon. Therefore, a more 
accurate math model of the damper was 
deemed necessary for further analysis, 
and for a more complete understanding of 
the problem. The derivation of the up- 
graded lag damper math model is shown in 
Reference (1) . Inclusion of this lag 
damper math model into the torsional 
stability computer simulation accurately 
reproduced the torque oscillation with 
the original fuel controls. Frequency of 
oscillation, phasing and magnitude of 
damper force, shaft torque oscillation, 
and fuel flow fluctuation were now sim- 
ulated accurately. Final simulation may 
be seen in Figure (4) . The primary 
difference between this damper simulation 
and the earlier version is that the new 
model included the hydraulic spring 
effect of the damper. 

The reduced gain-increased time con- 
stant fuel control fix has provided 
satisfactory torsional stability for the 
CH-47C production fleet. However, 
several early production aircraft reported 
instances of a "pseudo- torque oscillation". 
This phenomenon is a torque split, 
followed by a low amplitude torque oscil- 
lation of the high torque engine. The 
problem was traced to high levels of 
vibration affecting the internal workings 
of the fuel control. Vibration at cross 
shaft frequency caused an instantaneous 
increase in the effective gain of the 
control, increasing its torque output 
with respect to the other engine, and 



making it susceptible to torsional 
instability. The problem was resolved by 
closely monitoring cross shaft vibration, 
and with minor fuel control component 
modifications . 

During the latter part of the torque 
oscillation program, it became apparent 
that the engine and airframe manufacturers 
can easily coordinate their efforts to 
prevent this type of incompatibility. 
Lycoming has now provided Vertol with a 
mathematical model of the engine and fuel 
control system, so that rotor/drive system 
design changes may be evaluated for their 
effect on torsional stability. It is 
equally important that as accurate a 
representation as possible of the rotor 
and drive system be given to the engine 
manufacturer. 

There has been some mention in recent 
years about the possibility of using a 
zero torsional stiffness coupling 
(Reference (2)) to effectively isolate the 
engine from the rotor drive system, there- 
by precluding torque oscillation. At this 
time, potentially high developmental costs, 
uncertainty of transient behavior, and 
added weight to the drive system seem to 
rule out the z.t.s. coupling. However, 
continued research may yield an acceptable 
concept that may be the design solution 
for torsional instability for the next 
generation of increasingly larger, faster 
and more complex VTOL rotorcraft. 

Engine Vibration at 
Helicopter Rotor Frequencies 

The CH-47/T55 engine installation is 
"hard-mounted", as shown in Figure (5). 
It employs two front mounts on a yoke at 
the engine inlet housing, and an aft 
vertical support link at the engine 
diffuser flange. The outboard yoke air- 
frame point is connected to take out high 
f ore-af t maneuver loads . Engine vibration 
had rarely been a problem on the CH-47A 
and B models with this type installation. 

However, field service reports 
indicated an increase in engine, engine 
component and engine mount vibration- 
related problems with the installation of 
the T55-L-11 and -11A engines in the 
CH-47C helicopter. These problems led to 
a full scale engine and strain survey, the 
purpose of which was to determine the 
dynamic characteristics of the engine 
installation, especially the vibration/ 
strain relationships. The engine survey 
(Reference (3)) provided a wealth of 
information concerning the CH-47C engine/ 
airframe interface dynamic characteristics. 
In particular, the survey identified 
rotor 3/rev as the predominant excitation 
frequency in the engine mounting system. 



250 



Also, inlet housing stresses and drag strut 
load increased significantly with frequency 
(rotor speed) , as if approaching a res- 
onance, as shown in Figure (6) • As a 
result of this discovery, a ground shake 
test was recommended to define the char- 
acteristics of the apparent engine/air- 
frame mode being excited by rotor 3/rev. 
The shake test setup is shown in Figure 
(7). 

The CH-47C/T55-L-11 engine shake test 
revealed a 14.2 H z rigid body yaw mode. 
Installation of -11A engines (an addition- 
al 40 lbs.) caused a .4 H z downward shift 
in modal frequency, and a twofold increase 
in 3/rev inlet housing strains. Addi- 
tional testing showed that reducing drag 
strut bolt torque could lower the engine 
yaw mode frequency into the CH-47 operat- 
ing range (11.5 to 12.5 H z ) . Complete 
elimination of the drag strut lowered the 
mode to 7.5 H z , well below the CH-47C 
operating range. Shake test frequency 
sweeps are shown in Figure (8) . Removal 
of the drag strut, however, is not a 
practical solution. It is needed to 
assure acceptable cross shaft alignment 
under high maneuver G and jet thrust 
loads. The solution, therefore, was to 
retain the drag strut, but slot one end to 
eliminate dynamic stiffness for small 
amplitude motions, resulting in a struc- 
turally detuned installation. 

Flight evaluation of the slotted drag 
strut was desired, and the Model 347 
research helicopter was available as a 
testbed. The 14 H yaw mode fell within 
the operating n/rev frequency range (14-16 
H z ) of the four-bladed Model 347 and, 
therefore, it would be possible to verify 
the inflight placement of the mode. How- 
ever, rotor speed sweeps of from 210 to 
240 RPM with the standard strut failed to 
show a peak inlet housing stress response 
in the expected frequency range. Reducing 
rotor RPM still further finally located 
the engine yaw mode at 13.2 H . 

Installing the slotted drag strut on 
one engine completely eliminated the 13 H z 
peaks, and resulting 4/rev inlet housing 
stresses were reduced by as much as 75%. 
Lateral 4/rev vibration at the engine 
diffuser showed as much as an 85% reduc- 
tion. These load and vibration reductions 
are illustrated in Figure (9) . 

It is noteworthy that analytical 
efforts to predict the installation 
dynamic characteristics met with limited 
success. This analysis first made use of 
assumed values of fuselage backup struc- 
ture stiffness, and later used values 
calculated from a finite element struc- 
tural model of the entire fuselage. 
However, the accuracy of these stiffness 



values is a function of idealization 
accuracy and validity, and end condition 
assumptions. The analytical predictions 
began to resemble the actual test results 
only when static load-deflection test data 
at the engine support points was used in 
the analysis. It is important here to 
point out two other factors that con- 
tributed to the CH-47C engine vibration 
stress problem; the increase in normal 
rotor RPM from the A to C model to improve 
the flight envelope resulted in a higher 
forcing frequency, and the increasing 
engine weight and inertia of the more 
powerful engine moved the resonant 
frequency downward. 

Engine bending was not a contributing 
factor in this installation. In engine 
installations where it is a factor, the 
analysis becomes much more complex. Close 
coordination between engine and airframe 
manufacturers, through engine/airframe 
interface agreements, will be necessary to 
accurately describe the installed engine 
dynamics in this case. 

In the overall design of an engine 
installation, it is imperative to choose 
the engine dynamic characteristics 
(isolated, detuned or hard mounted) such 
that output shaft alignment is not 
jeopardized. Or, conversely, output shaft 
couplings must be tailored to the vibra- 
tory environment of the engine. In an 
isolated engine installation (where most 
engine modes are placed well below pre- 
dominant forcing frequency) , output shaft 
couplings with high misalignment capa- 
bility must be employed. In a hard-mounted 
or detuned installation, low misalignment 
couplings, such as the Thomas coupling, 
may be utilized. 

Rotor/Drive System B/Rev 
Torsional Resonance 

The Boeing Vertol Model 347 research 
helicopter is a derivative of the CH-47C 
Chinook helicopter, the primary differences 
being a 30 inch higher aft pylon, a 100 
inch longer fuselage, and an increase in 
rotor blades from 3 to 4 per rotor 
(Reference (4) ) . A Holzer torsional 
analysis of the CH-47C revealed natural 
modes at roughly .3/rev, .9/rev, 4.1 and 
4.2/rev; therefore, the Chinook was con- 
sidered to be free from b/rev torsional 
resonance (3/rev in this case) . A similar 
analysis on the Model 347 revealed almost 
identical non-dimensional torsional 
frequencies, despite a lengthened aft 
rotor shaft and forward synchronizing 
shaft, and a reduction in rotor RPM. 
There was some concern about the proximity 
of the third and fourth torsional modes to 
b/rev (4/rev in this case) . However, it 
was believed that forcing levels and 



251 



phasing would not be sufficient to excite 
these modes. The Model 347 drive system 
torsional modes are shown in Figure (10) . 

The Model 347 program was flown 
successfully, until the aircraft was flown 
at high gross weights. Here, high 4/rev 
blade chordwise bending moments in transi- 
tion and high speed forward flight became 
a structurally limiting factor. Examina- 
tion of flight test data revealed that the 
chordwise bending moments of all four 
blades on each hub were exactly in phase. 
Data also revealed substantial rotor. shaft 
4/rev torque fluctuations, with the for- 
ward and aft rotor systems opposing each 
other as shown in Figure (11) , and 4/rev 
chordwise bending moments increasing 
sharply with RPM, as if approaching a 
resonance (Figure 12) . 

Analytical parametric studies were 
conducted to evaluate the effect of various 
system modifications on the apparent 4/rev 
resonance. Modifications such as forward 
and aft rotor shaft stiffness changes, 
synchronizing shaft stiffness changes and 
effective lag spring stiffening were all 
found to be effective to some extent. 
However, these changes were rejected due to 
the magnitude of change required to move 
the resonance and sensitivity to RPM 
changes. A much more acceptable modifica- 
tion was found to be raising the blade 
uncoupled chordwise bending natural 
frequency. On the CH-47C, this blade 
frequency was just above 5/rev; conse- 
quently, the largest blade bending loads 
are at 5/rev. However, with these same 
blades on the Model 347, the largest blade 
bending loads were at 4/rev, indicating 
the blade/drive system coupling effect. 

Both blade softening and stiffening 
were investigated. It was found that 
decreasing the blade chordwise bending 
frequency was more effective in moving the 
drive system resonance than the same per- 
centage increase, as shown by the Figure 
(13) analysis. But it was felt that this 
blade softening would present too great a 
structural degradation problem in the 
blade. Hence, raising the blade chordwise 
frequency, and with it the coupled blade/ 
drive system torsional resonance, was the 
design goal. Analysis revealed that a 4 H 2 
increase in blade natural frequency would 
result in satisfactory detuning of the 
blade/drive system resonance. 

The most effective location to attempt 
a chordwise frequency increase is at the 
trailing edge. It was necessary in this 
case to add on a material of high stiffness 
and minimum weight, such that chordwise 
balance and CF loads are not grossly 
affected. The design selected consisted of 
top and bottom boron fiber doublers bonded 



to the stainless steel trailing edge from 
30% to 70% span, and boron skins applied 
to several blade boxes. The benefit of 
the boron stiffening is twofold, for in 
addition to increasing the chordwise 
frequency to avoid resonance, strength is 
increased. 

The addition of boron stiffening moved 
the blade uncoupled flexible chordwise 
frequency from 5.26/rev to over 6/rev. 
This resulted in a shift in the blade/ 
drive system natural frequency to over 
4.2/rev (at 235 RPM) or to 4.3/rev (at 220 
RPM) . This was sufficient to preclude 
high 4/rev amplification, since blade 
chordwise trailing edge loads are now 
highest at 6/rev (the uncoupled blade 
frequency) . 

This problem does not fall strictly 
into the category of engine/airframe inter- 
face dynamics. However, the influence of 
the engine in the drive system dynamics, 
and the potential impact of such a problem 
on the engine cannot be ignored. For 
example, to accurately predict drive 
system modes, the power turbine inertia 
must be accurately known. 

Engine Output Shaft 
Critical Speed Analysis 

The Boeing Vertol Heavy Lift Helicop- 
ter prototype will incorporate three 
Detroit Diesel Allison XT701-AD-700 turbo- 
shaft engines. These engines have been 
developed from the Allison 501-M62B as 
part of a program to procure representative 
engines for the HLH Advanced Technology 
Component (ATC) dynamic systems test rig. 
Many helicopters built in the past were 
designed around existing engines. However, 
in the case of the HLH, initial development 
of the engine is to be fully coordinated 
by the prime contractor; hence, development 
of both engine and airframe will- be in 
parallel. The HLH engine program is dis- 
cussed in Reference (5) . 

A development problem was encountered 
during the program which involved the 
engine/airframe output drive shaft inter- 
face. The original design of the engine 
output shaft was a short splined shaft 
with the torquesensor mounted within the 
main frame of the engine. Based on more 
detailed engine nacelle design, it was 
requested that the splined shaft interface 
be moved forward to reduce inlet blockage 
and to facilitate inspection of the shaft 
coupling. This change was agreed upon, 
and the drive shaft connection was moved 
to a point 17 inches forward of the front 
face of the engine. The torquesensor 
was also housed in the resulting engine 
"nose". A cutaway view of the torque- 
sensor and housing is shown in Figure (14) . 



252 



The original shafting concept on the 
HLH was to drive into the main transmis- 
sion directly, without right angle gear- 
boxes, resulting in a substantial weight 
savings. A layout of the original HLH 
engine/ntixbox shaft configuration is 
shown in Figure (15) . The original 
engine-to-mixbox shafting consisted of two 
7.25 inch diameter sections of equal 
length with a single bearing support 
point. However, in an attempt to further 
reduce inlet blockage and reduce weight, 
the shaft diameter was reduced to 6 inches. 
This decision was based on preliminary 
analytical trade studies which used an 
initial estimate of engine flexibility. 
Critical speed placement was analyzed to 
be more than 25% above normal operating 
speed (11,500 RPM) . 

As the detailed design of the engine 
progressed and was included in the critical 
speed analysis, it became apparent that the 
anticipated critical speed margin would not 
be realized. The analysis was expanded to 
include the torque sensor, its housing, 
bearings, and effective engine radial and 
moment flexibility. This more detailed 
analysis, performed at Detroit Diesel 
Allison and confirmed by Boeing Vertol, 
revealed the shaft/ torque sensor whirl 
mode in the area of 12,500 - 13,000 RPM, 
or only about 10% above normal operating 
speed. The analytical mode shapes and 
frequencies are shown in Figure (16) . 

Working together, both companies 
conducted parametric analyses to evaluate 
various potential fixes. Prime candidates 
were inlet housing and torquesensor 
housing stiffness increases, a shorter 
engine nose, auxiliary support struts, 
stiffened torque sensors, plus combina- 
tions; however, when they were analyzed in 
combination with a complete engine dynamic 
model, none proved satisfactory. In fact, 
with the complete engine model, the 
critical speed of the original configura- 
tion was around 10,200 RPM, below normal 
operating speed. The mode involved sub- 
stantial whirl of the torquemeter housing, 
some shaft bending and some case bending, 
and was very sensitive to output shaft 
coupling weight and unbalance. 

This analysis revealed that the only 
practical solution was a drastic shorten- 
ing of the torquesensor and housing, such 
that the shaft adapter is an integral part 
of the engine output shaft, and the 
flexible coupling is now only 5.3 inches 
from the front face of the engine. Due to 
the increased distance between the engine 
and combining transmission, the output 
shaft was changed to a 3-section configura- 
tion. This also reduced the amount of 
weight hung off the engine. Analysis of 
this configuration placed the natural mode 
at about 14,200 RPM, which was basically 



power turbine conical whirl interacting to 
some extent with the torquesensor shafting. 
Another mode at about 17,200 RPM showed 
compressor conical whirl with rotor, power 
turbine and case participation. Forced 
response analysis showed both these modes 
were only mildly responsive to mass unbal- 
ance at the output shaft coupling, as 
shown in Figure (17) . This indicates that 
the desired shaft/engine dynamic decoupling 
has been accomplished. 

It is interesting to note how design 
decisions not directly related to engine 
shaft dynamics provided constraints to the 
solution of the interface problem. For 
example, the decision to move the shaft 
interface well forward of the engine front 
face led to the long torquesensor housing 
design, which brought about the shaft/ 
torquesensor whirl problem in the first 
place. Also, the engine/shaft interface 
could not be moved very much closer to the 
engine front face without shortening the 
torquesensor. Since torquesensor accuracy 
is a function of length, the decision to 
drastically shorten the torquesensor and 
housing was made with reluctance, since 
torquesensor accuracy had to be compromised 
to some extent. 

Another interesting aspect of this 
problem is the fact that the critical speed 
of the engine-to-mixbox shafting could not 
be accurately analyzed until the complete 
engine dynamics were included. This is 
where the engine/airframe interface agree- 
ment in effect between Boeing Vertol and 
Detroit Diesel Allison has been instru- 
mental. It has led to excellent working 
agreements between the companies that have 
helped to reveal, analyze and solve this 
potential problem before it reached the 
hardware stage. Preliminary shaft critical 
speed work was done at Boeing Vertol. How- 
ever, when it became apparent that engine 
dynamics must be included to accurately 
predict the critical speeds, all work was 
done jointly with Allison. 

Conclusions 

(1) Helicopter engine/drive system 
torsional instability may be pre- 
vented if care is taken to accurately 
represent both engine and rotor 
systems in the analysis, including 
such effects as hydraulic compress- 
ibility of the blade lag damper. 

(2) Accurate analysis and/or shake testing 
of all engine installations, whether 
hard mounted, detuned, or isolated, is 
required to determine potential engine 
vibration and stress problem areas. 

(3) Helicopter rotor blades and drive 
systems must be designed such that 
blade lag flexibility does not couple 



253 



with drive system torsional flexibil- 
ity to produce a resonance at the 
number of rotor blade's frequency 
(b/rev) . 

(4) Formal engine/airframe interface 
agreements have already proven 
beneficial in the timely resolution 
of potential interface dynamics 
problems . 

References 



1. Fredrickson, C, Rumford, K. and 
Stephenson, C, FACTORS AFFECTING FUEL 
CONTROL STABILITY OF A TURBINE ENGINE/ 
HELICOPTER ROTOR DRIVE SYSTEM, 27th 
National American Helicopter Society 
Forum, Washington, D.C., May 1971. 

2. Vance, J. M. and Gomez, J., VIBRATORY 
COMPATIBILITY OF ROTARY-WING AIRCRAFT 
PROPULSION COMPONENTS, 29th National 
American Helicopter Society Forum, 
Washington, D.C., May 1973. 

3. Boeing Vertol Company, D210-10348-1, 
CH-47C/T55-L-11 ENGINE VIBRATION AND 
STRAIN SURVEY, Rumpel, M. , October 
1971. 

4. Hooper, W. E. and Duke, E., THE MODEL 
347 ADVANCED TECHNOLOGY HELICOPTER, 
27th National American Helicopter 
Society Forum, Washington, D.C., May 
1971. 

5. Woodley, D. and Castle, W. , HEAVY LIFT 
HELICOPTER MAIN ENGINES, SAE Technical 
Paper 730920, October 1973. 



254 



4.1 HZ 



FWD, ROTOR 
SHAFT TORQUE 

IV 



AFT ROTOR 
SHAFT TORQUE 



FUEL FLOW 
ENG. TORQUE 




Alt. ± 116K in- lb. 
(12.4% of Max.) 



A A It A 

Alt. ± 10 3K in- lb. 
(11.0% of Max.) 



WVK 



AA/WW\M 



DAMPER FORCE 




± .455 GPM 
(8.95%) 

Steady 80.1% 
Alt. t 7.6% 



± 2800 lb. 



Original Fuel controls 
Standard Lag Dampers 



Figure 1. Torque Oscillation Plight Test 
Data 



5000 



. 4000 



<u 
o 

£ 3000 

En 

u 
<u 
& 



2000 



n) 



1000 



Damper "Stiffened" to 
Reproduce Oscillation 
on Computer 



t ' Damper "Softened" to 
Eliminate Oscillation 
in Test 




1.0 2.0 3.0 

Peak Damper Velocity, In/Sec 

Figure 2. Lag Damper Force-Veiocity Curves 



+74K IN-LB. 



ROTOR SHAFT 
TORQUE 



+10K IN-LB. 



& 

s 

B 



i.O 
.8 

.7 
.6 

.5 

.4 

.a 



■2 



/ORIGINAL PROD. 
(.0 3 SEC. T/C) 
-30% REDUCED GAIN 



'" ■ » ■ -,' —-^ jo % 

---Ox 

*•.. \ ^-sINCRE 




LAG DAMPER 
LOAD 



+ 400 LB. 



J 



INCREASED TIME 
CONST. (.10 SEC.) 



FINAL FIX-^ \ \ X s 
(.10 SEC. T/C, \ \ \ 
30% RED. GAIN) \ \ \ 

• Si ■ - v ■ v ■ 2" 

2 3 A 5 6 7S5IO to 80 



FREQUENCY, CPS 
Figure 3. Fuel Control Frequency Response 



+2500 LB. 



y\/\/w 

+.33 GPM 



ORIGINAL 
FUEL CONTROLS 



FUEL FLOW 



REDUCED GAIN AND 
"SLOWED DOWN" 
FUEL CONTROLS 



Figure 4. Final Torque Oscillation Simulation 



255 




2500 




230 240 250 
ROTOR SPEED - RPM 



Figure 6. Inlet Housing Stress & 

Drag Link Load vs . RPM 



256 



^ W3 BP* : ¥'^ s *^ 1 Si 




■Figure 7. CH-4 



W 

3 

EH 
TO 

O 

a 

H 
CO 
D 
O 

En 
H 



2000 



1500 



1000 



500 







3/REV 










/ 


^STANDARD 
STRUT 




NO STRUT 
/ 




1 \ 








%. 




-— -^S^K 



10 15 

FREQUENCY - Hz 



20 



Figure 8. CH-47C/T55-L-11 Engine Shake Test Results 



257 



800 



INLET HOUSING 4/REV STRESS 



LATERAL DIFFUSER 4/REV VIBRATION 




180 



200 apjyj 220 



240 



12 



13 



14 
Hz 



15 



16 




ROTOR SPEED - RPM 



180 



200 



220 



240; 



Figure 9. Vibration and Stress Reductions with Slotted Drag Strut 



Fwd. 



Aft 




Modal Deflections Normalized to 
1 Degree @ Fwd. Rotor Hub 



Figure 10, Model 347 Drive System 
Torsional Modes 



Fwd. Blade 
Chordwise 
Bending 
-In-Lb. - 



Fwd. Lag 
Damper Load 
-Lb.- 



;ip FWD 




ULUl 100,000 t 

Torque J yVVV> ftftfl^ /WV 

'*" -ioo^oo Idsivino I I 

Dtor "•••«• r 

Torque *f- 
b.- -ioo.oooIc 



Fwd. Rotor m,ooo 
Shaft 
-In-Lb 

-IDO^OO «D8IVIN0 

Aft Rotor «*-•»><>% 

Shaft Torque °r — i^^V^/V j s^/lvlvH/VV^ 

"In-Lb.- -100,400 1 DRIVISS 
SOOOf TENSION 

Aft Lag 
Damper Load 
-Lb.~ 



Af t Blade 
Chordwise 

Bending 
-In-Lb • — -401000 




vvv^ 



1 ROTOR REVOLUTION 



0,3. 0.4 0,6 ad 

SIMS -MM-- 



Figure 11. Model 347 Rotor/Drive System 
Flight Test Data 



258 



i 40,000 

H 
+1 



§ 30,000 

H 
W 

a 

En 

H 


@ 20,000 

g 

Hi 
H 



^ 



4/REV 



GW 32,100 LBS 
TAS 150 KNOTS 



10,000 






o 

lis 



A 



200 



210 



220 
ROTOR RPM 



230 



240 



Figure 12. 



Model 347 Blade Chordwise 
Bending Moment vs. RPM 



5.0 




m 
u 

H 

■a 

S 
d) 
-P 
to 
>i 

> 

(!) 0) 
> M 
■H 

u u 
a a) 

■d 

H 

m 
■d 

<D 
H 

§• 

o 
o 



Uncoupled Blade Ghordwise Bending Freq. 
{per rev) 

Figure 13. Effect of Chordwise Blade Bending 
Frequency on Model 347 Drive 
System Modes 



Output Drive Shaft 
Torquesensor Shaft - 
Extended Housing — 

B/V Shaft 




Bearings 
Figure 14. Original 501-M62B Torquesensor Configuration 



259 




Figure 15. Original HLH Engine to Combiner 
Box Shafting 



Engxne 
11500 EPM t 1 ? 




215. Hz (12900 RPM) 
i 



Figure 16. Preliminary Engine/Shaft Dynamic 

Analysis showing Torquesensor/Shaf 
Conical Whirl Mode 



0) 

o 
a 
<d 

a 

D 
N 

o 
I 

a 



< 
a 



.20 



.15 



.10 



• 5.5" Overhang 

• Torquesensor in 
Inlet Housing 

• 1 In-Oz Unbalance 
@ Thomas Coupling 







h HP Rotor 
LP Rotor 
V Torquesensor 
© B/V Shafting 



8000 



12000 16000 20000 



Engine Speed - RPM 

Figure 17. Final HLH Engine/Shaft Analysis 
Response to Unbalance 



260 



fflNGELESS ROTOR THEORY AND EXPERIMENT 

ON VIBRATION REDUCTION BY PERIODIC VARIATION 

OF CONVENTIONAL CONTROLS 

G. J. Sissingh and R. E. Donham 

Lockheed-California Company 

Burbank, California 



Abstract 

The reduction of the n per rev. pitch-, roll- and vertical 
vibrations of an n-bladed rotor by n per rev. sinusoidal variations 
of the collective and cyclic controls is investigated. The 
numerical results presented refer to a four-bladed, 7.5-foot model 
and are based on frequency response tests conducted under an 
Army-sponsored research program. The following subjects are 
treated: 

• Extraction of the rotor transfer functions (.073R hub 
flapping and model thrust versus servo valve command, 
amplitude and phase) 

• Calculation of servo commands (volts) required to 
compensate .073R hub flapping (3P and 5P) and 
model thrust (4P) 

• Evaluation of the effect of the vibratory control inputs 
on blade loads 

• Theoretical prediction of the root flapbending 
moments generated by o to 5P perturbations of the 
feathering angle and rotor angle of attack. 

Five operating conditions are investigated covering advance 
ratios from approximately 0.2 to 0.85. The feasibility of vibra- 
tion reduction by periodic variation on conventional controls is 
evaluated. 

Summary 

For several operating conditions covering advance ratios 
from approximately 0.2 to 0.85, the control inputs required to 
counteract the existing 4P pitch, roll and vertical vibrations are 
calculated. The investigations are based on experimental vibra- 
tion and response data. As the tests were part of and added on to 
a larger hingeless rotor research program, only a few operating 
conditions with essentially zero tip path plane tilt were investi- 
gated because of limited tunnel time. At the test rotor speed (500 
rpm) the rotor blade mode frequencies were 1 .34P, first flapping, 
6.3P, second flapping, and 3.6P, first inplane. 



This work was conducted under the sponsorship of the 
Ames Directorate of the U. S. Army Air Mobility R&D Lab- 
oratory under Contract NAS2-7245.'The authors gratefully 
acknowledge the assistance of Mr. David Sharpe, the AMRDL 
Project Engineer, and Messrs. R. London and G. Watts of 
Lockheed in conducting the experimental portion of this work. 



It should be noted that there was no instrumentation to 
measure the vibratory pitching and rolling moments. These 
moments were obtained by properly adding up the flap-bending 
moments of the four blades at 3.3 in. (0.073R) which were 
measured separately. This means, the effects of the inplane 
forces, vertical shear forces and blade torsion have been ignored. 
These are important influences in current hingeless rotor designs. 
The inplane 3P and 5P shear forces are of particular 
interest. However, the experimental data obtained for a model 
hingeless rotor system provides the beginning of at least a partial 
data base for the investigation of vibration attenuation of such 
systems through periodic variation of conventional controls. 

Generally speaking, the control inputs required for flapping 
(hub moment) sourced vibration elimination are smaller or about 
of the same magnitude as those used for the frequency response 
tests. Their amplitudes lie, depending on flight condition and 
advance ratio, between 0.2 and 3 degrees. With the exception of 
the m = 0.85 1 case, for which the results are somewhat in doubt 
(the response tests to lateral cyclic pitch and the corresponding 
baseline data were inadvertently run with 0.3-degree collective 
pitch differential), the control inputs required for vibration re- 
duction drastically reduce the 3 and 5P, and have only a minor 
effect on the 2P flexure flap-bending moments. Chord-bending 
moments and blade torsion generally increase. 



The theoretical predictions mentioned refer to forced- 
response influence coefficients. They are based on the first two 
flapping modes. The blade root flap-bending moments (OP 
through 5P) which result from unit perturbations of blade 
feathering angle and rotor angle of attack have been calculated. 
The solution provides for intermode coupling through the 17th 
harmonic by analytic solution of the two-degree-of-freedom 
system, utilizing constant coefficient and loading descriptions 
over ten-degree azimuth sectors. In each solution case, the rotor 
reached steady-state motion in eight revolutions. In that time the 
least converging second mode flapping motion converged to a 
minimum of four significant figures. 

Evaluation of the test data reveals two types of short- 
comings, which should be avoided in future tests. First, the data 
given are based on a single test and have not been verified. 
Second, in some cases, the baseline and frequency response tests 
were not run successively. 

From the data available, the approach is promising, 
especially for the low and medium advance ratio range. At higher 
advance ratios (n~ 0.8), the control inputs required for vibration 
reduction may become prohibitive. 

Notation 



Presented at the AHS/NASA-Ames Specialists' Meeting on A, B 

Rotorcraft Dynamics, February 13-15, 1974. 



quantities describing cos 4^ and sin 4^ components 
of actuator input for frequency response tests, volt, 
see Table II and Equation (1) 



261 



C, D quantities describing responses to A and B, in.-lb 

and lb, respectively, see Equation (1) 

E, F, G, H blade loads due to unit actuator input, in.-lb/ volt, 
see Equation (13) 

Kj . . . Kjg gains of rotor response, see Table I 

m calculated flapbending moment at 3.3 in., in.-lb, 



m = m + 2m ns sin n# + 2m nc cos n^ 



M, L, T 4P vibratory pitching moments, rolling moments 
and thrust variations, in.-lb and lb, respectively; 
subscript e denotes existing vibrations to be com- 
pensated, subscript control describes effects of 
oscillatory control inputs. 



M e = Mg sin Aii + M c cos 4^ 



L e = L s sin 4^ + L C cos Aii 



T e = T s sin Aii + T c cos Ai> 



^nominal nominal collective pitch, degrees 

O , 6 S , C oscillator inputs for collective, longitudinal 
and lateral cyclic pitch,*volt 



6 o = e os sin 4 * + e oc cos 4 * 



S = SS sin Aii + 8 SC cos Aii 



C = 0™ sin Aii + 6 rr cos Aii 



T l • • • T l 8 * a 8 aa ^ es °f response, degrees, see Table I 



SI 



rotor angular velocity, sec' 



■1 



azimuth position of master blade, rad 



C„„ Blade Root Moment, STA (o) 

RM 

a<T ttR 3 p(«R) 2 aa 



where 

a = 5.73 

p = 0.002378 slugs/ft 3 

a = 0.127 

"Compensating Control Inputs" define those which reduce 
the existing 4P pitching moments, rolling moments and vertical 
forces of a given flight condition to zero. 

The analysis deals with the concept of vibration reduction 
by oscillatory collective and cyclic control applications. Several 
related aspects of this problem are treated. The foremost are the 
determination of the proper control inputs and their effect on 



the vibratory blade loads. These studies are based on frequency 
response tests conducted on a 7.5 foot-diameter, four-bladed, 
hingeless rotor model, the results of which are published in 
Appendixes C and D of Reference 1. The subject matter covered, 
apart from the items listed below, is an abridged version of these 
appendixes. 

Other subjects treated are (a) the calculation of blade loads, 
based on test data, due to vibratory control command applica- 
tions; (b) the theoretically determined eigenvalues, at 10-degree 
azimuth intervals, of the first and second flapping modes, at 
M = 0. 1 9 1 , 0.45 and 0.85 1 ; (c) the computed single-blade root 
flap-bending moment, Sta 0, harmonic influence coefficients 
at m = 0. 1 9 1 , 0.45 and 0.851; and (d) a limited comparison of the 
theoretical loads with experiments. 

The general case of vibration control will include the effects 
of lateral and fore-and-aft shear forces at blade passage frequency. 
These forces can be as influential as the pitch and roll moment 
and thrust oscillations in causing fuselage vibrations. Thus, in 
general, five rotor vibratory inputs are to be controlled by mani- 
pulation of three controls. Although the five vibratory inputs 
cannot be nulled individually with three controls, their combined 
contribution to the fuselage vibration can be controlled. Thus, 
the general application will involve control of fuselage vibration 
at three points; say two vertical vibrations and one roll angular 
vibration. This general application implies the use of adaptive 
feedback controls. Although the present paper is limited to the 
more simple case outlined herein, the general application to the 
control of any three suitable quantities will be apparent. 

Although prior investigations of the use of higher harmonic 
pitch control on teetering and offset hinge rotors have been con- 
ducted to investigate improved system performance and also for 
vibration attentuation (References 2 , 3 and 4 ), this is believed 
to be the first experimental and theoretical hingeless rotor study 
of the use of periodic variation of conventional controls for 
vibration attentuation. The use of 2P feathering to improve rotor 
performance is not included as part of this work. 

Transfer Functions Involved 

As a distinction must be made between control applications 
in phase with sin Aii and cos 4^, there are six control quantities 
available, i.e., QS , QC , &s , d sc , 8 CS and CC , to monitor the 
pitching moments, rolling moments and vertical forces. This 
means the dynamic system investigated, which consists of rotor, 
control mechanism and oscillators used, is characterized by 18 
gains Kp and lag angles r p . The subscripts p (p = 1 through 1 8) 
are defined by Table I. 



TABLE I 

GAINS AND LAG ANGLES OF RESPONSE 

TO OSCILLATORY CONTROL APPLICATIONS 





e os 


Ooc 


fl ss 


9 sc 


9 cs 


fl cc 


M 


K m 


% T 2 


K 3 T 3 


K 4 r 4 


K 5 '5 


K 6 '6 


L 


K 7 r 7 


K 8 T 8 


K 9 r 9 


K 10 r 10 


K ll r ll 


K 12 T 12 


T 


K 13 T 13 


K 14 r 14 


K 15 T 15 


K 16 T 16 


K 17 T 17 


K 18 T 18 



262 



As indicated, Kg is defined as the amplitude ratio M/0 SS and 
t 3 is the lag angle of M with respect to 8 SS . For convenience, the 
dimensions used are identical with those of the computer output, 
i.e., oscillator voltage for input, in.-lb for M and L, lb for the 
thrust variation T. This means the dimensions of Kp are 



Kj through Kj 2 
Kj3 through Kjg 



in.-Ib/volt 
lb/volt 



See also Figure 1 which shows the oscillatory pitching moments 
due to combined 8 SS and SC control applications. The moments 
generated are presented by rotating vectors where cos 40 is posi- 
tive to the right and sin 40 positive down. This means, the vector 
positions shown refer to = 0. By definition, the quantities Ry 
characterize the responses in phase with the excitation and Iy 
those out of phase. The latter are positive if the response leads. 
As indicated, there are altogether four responses involved which 
are combined to the resultant M. 



The phase angles r p are given in degrees, r p is positive if the re- 
sponse lags. 

Although the investigations deal exclusively with 4P control 
variations, some general remarks may be in order. The general 
case involves sinusoidal collective and cyclic control variations 
with the frequency n£2 where n can be any positive number. 

If n is an integer, the rotor excitations repeat themselves 
after each rotor revolution which means that the responses of 
each revolution are identical. This is true for any number of rotor 
blades but does not necessarily mean that all blades execute 
identical flapping motions. The latter is true only if n equals the 
number of rotor blades or is a multiple of the blade number. 
Only for these cases does a truly time independent response with 
invariable amplitude ratios K and lag angles r exist. 

Extraction of Gains and Lag 
Angles from Experiments 

As for all response tests conducted, the oscillator input con- 
tained both sin 40 and cos 40-components; always two amplitude 
ratios K and two lag angles t sis involved. Therefore, each time a 
set of two tests must be evaluated. According to Table II, the 
input is characterized by the quantities Aj B j A2 B2 and the 
response by Cj Dj C2 D2. 

If the rotor responds to cos 40 excitations with the gain K; 
and the lag angle t-. (j = even number) and to sin 40 excitations 
with Kj and r- x (i = odd number), input and output are related by 
the equations 

Aj Kj cos (40 - rp + Bj Kj sin (4* - t { ) = C j cos 40 

+ Disin40 



A 2 Kj cos (4* - Tj) + B 2 K } sin (40 - Tj) = C 2 cos 40 

+ D 2 sin 4^ 

TABLE II 
INPUT AND OUTPUT NOTATIONS 



(1) 



Test 


Input 


Response 


#1 
#2 


Aj cos40 + Bj sin 4* 
A 2 cos 40 + B 2 sin 40 


q eos40 + Dj sin 40 
C 2 cos 40 + D 2 sin 40 



To calculate the unknowns Kj Kj Tj and v., a component 
analysis is used. The gains Kj Kj are expressed as 



Kj = (R? + 1?) 1/2 
Kj - (R?+ I. 2 ) 1/2 



(2) 




SS'3 



SIN 40 

Figure 1. Vector Diagram Showing Pitching Moment 
Due to SS and 9 SC Control Applications 

Inserting Equation (2) into Equation (1) leads to 



A,D 2 -A 2 D 1 



AiBj-AsBj 

A 1 C 2 -A 2 C 1 
A 1 B 2" A 2 B 1 



tan?;= llj/Rj I 0<Ti<ff/2 



(3) 



and 



Rj = 



Y 



CiB 2 -Bj[C 2 
AjB^B! 

BiD 2 -B 2 Pl 
AiB 2 -A 2 B, 



tan?j= llj/Rjl 0<Tj<jr/2 



(4) 



263 



In both cases 



r=+T for R>0 I<0 

= -t R>0 I>0 

= 7T+? R<0 I>0 

= rr-r R<0 I<0 



Check of Calculated Kj K: Tj and r.- Values 

If so desired, Equation (1) can be used to check the calcu- 
lated values of Kj Kj Tj and' t-.. Splitting up these equations into 
sin 4# and cos 4v components leads to the following four 
expressions which must be satisfied 



Aj Kj cost: -Bj Kj sin Tj = Cj 
Aj Kj sin Tj + B j Kj cos-tj = Dj 
A2 K: cos Tj - B2 Kj sin tj = C2 

A2 & Sin T: + B2 Kj COS Tj = D2 

Oscillatory Control Inputs Required 



(5) 



The six oscillator inputs available have to be selected so that 
their responses satisfy the requirements, whatever they may be. 
By definition, the vibratory control inputs result in the following 
pitching moments, rolling moments and vertical forces (n = 4): 

M control = + e os K l sin ( n *- T l) 
+ e oc K 2 cos(n^-T 2 ) 
+ 9 ss K 3 sin(n*-r 3 ) 
+ sc K 4 cos(n<J'- r 4 ) 
+ cs K 5 sin(n*-r 5 ) 

+.fl cc K 6 cos(n*-r 6 ) (6) 

L control =+e s K 7 sin ( n *- T 7) 
+ 9 oc K 8 cos(ni//-Tg) 

+ ss K 9 sin(n^- T g ) 

+ e sc K 10 cos(n*-T 10 ) 

+ cs K 11 sin(n*-T 11 ) 

+ cc K 12 cos(n>/'-r 12 ) (7) 



T control = + e os K 13 sin ( n *" r 13> 

+ e oc K 14 cos ( n *- T 14) 
+ ss K 15 sin(n*-r 15 ) 

+ e sc K 16 cos(n*-r 16 ) 

+ 6» cs K 17 sin(n'/'- t 17 ) 

+ cc K 18 cos(n</'- t 18 ) 

M control = ^ sin 4 * " M c cos 4 * 
L control = " L s sin 4 * " L c cos 4 * 
T control =-T s sin44'-T c cos4* 



(8) 



(9) 



To reduce the existing vibrations, the moments and forces 
generated must counteract M e , L e and T e , i.e., 

Equations 6 through 9 lead to six linear equations, (10), 
for the unknowns OS , OC , 8 SS , SC , CS and 6 CC . 

Effect on Blade Loads 



An objective of the investigations is to determine the effect 
of the compensating control input on the blade loads, i.e., on the 
following measured quantities: 

• flapbending at 3.3 in. 

• flapbending at 1 3. 1 5 in. 

• chordbending at 2.4 in. 

• torsion at 9.28 in. 

In all cases the 2 to 5P content of the loads is of interest. 
The first task is to determine from the response tests the contri- 
bution of each of the six possible 4P control inputs to these 
loads. Again, two sets of data are required. The vibratory control 
applications used and the resulting n" 1 harmonic of the load con- 
sidered are written as follows : 



Test 



Input 



Resulting Load (in.-lb) 



#1 Aj cos 4^ + B j sin 4^ C n j cos n^ + D n j sin nf 
#2 A 2 cos 44* + B 2 sin Ai> C n2 cos n* + D n2 sin n* 



(11) 



+Kj cos Tj +K 2 sin t 2 +K3 cos T3 +K4 sin t 4 +K 5 cos r 5 +Kg sin Tg 

-Kjsin Tj +K 2 cos t 2 -K3 sin T3 +K4 cos t 4 -Kg sin tj +Kg cos Tg 

+K7 cos Tj +Kg sin Tg +Kg cos 19 +Kjq sin Tjq +Kj j cos tj j +Kj2 sin t j2 

-K7 sin Tj +Kg cos Tg -K9 sin Tg +K jq cos Tjq -Kj j sin tj j +Kj2 cos r ^ 

+Kj3Cosfj3 +Kj4sinrj4 +Kj5COsrjj +KjgsinTjg +KJ7C0STJ7 +Kjg sin Tjg 

-Kj3sinTj3 +KJ4COSTJ4 -KjgsinTjj +Kjg cos Tjg -Kj^sinrj^ +Kjg cos Tjg 

264 





#OS 




-M s 




9 oc 




-Mc 




e ss 




-h 




9 SC 




"L c 




CS 




" T s 




_0 C c. 




3. 



(10) 



If nonlinear effects are excluded, the n per rev load vari- 
ation due to unit control application in phase with 

(a) cos # amounts to (E n cos ni^ + F n sin n# ) 



(b) sin 4* (G n cosn<HH n sinniJ>) 

In these expressions 



(12) 



These moments were obtained by properly adding up the 
flap-bending moments of the four blades at 3.3 in. which were 
measured separately. This means, the effects of the in-plane 
forces, vertical shear forces and blade torsion have been ignored. 

TABLE IV 

VIBRATORY MOMENTS AND FORCES 

TO BE COMPENSATED 



E n = 



B 2 C nl- B l C n2 



% = 



H„ 



A 1 B 2" 


A 2 B 1 


B 2 D nl 


- B l D n2 


A 1 B 2" 


A 2 B 1 


A l c n2 


" A 2C„1 


A 1 B 2" 


A 2 B 1 


A lDn2 


- A 2^1 



A 1 B 2 -A 2 B 1 



(13) 



If 0£ S , 6t c (t, = o, s, c) denote the vibratory control inputs 
used, the increments of the n tn harmonic of the load considered 



(14) 



(Aload) n = (0g c E n + 0t s G n ) cos n^ 

+ (0| C F n + £s H n )sinn* 
Evaluation of Experiments 
Flight Conditions Investigated 



The methods outlined in the previous sections are applied to 
the following five operating conditions for which test data are 
available: 

TABLE HI 
OPERATING CONDITIONS INVESTIGATED 



M 


^nominal 


a 


Cj/ff 


0.191 


12° 


-5° 


0.102 


0.239 


4 


-5 


0.028 


0.443 


4 


-5 


0.011 


0.849 


10 


-5 


-0.005 


0.851 


4 


-5 


-0.013 



In all cases the shaft angle of attack is a = -5° and the rotor is 
trimmed so that essentially a| =b| =0. As can be seen, the tests 
cover the advance ratio range from approximately fi = 0.2 to 
H= 0.85. The case u= 0.191 is characterized by nomuia i = 12° 
and C T /ff = 0.102, the latter figure indicates a relatively high 
specific loading. In contrast, at the advance ratios ji = 0.849 
and 0.851 the rotor is practically unloaded, i.e., no steady lift- 
ing force is generated. The 4P vibrations associated with the 
various test conditions are listed in Table IV. The moments are 
given in inch-pounds and the vibratory forces in pounds. 



V 


0.191 


0.239 


0.443 


0.849 


0.851 


Ms 


0.3805 


- 1.7207 


2.6149 


20.0483 


3.5349 


M c 


- 0.5301 


-0.4113 


-0.5208 


-4.5724 


-8.4341 


h 


12.2080 


1.3725 


-6.7626 


9.4647 


-10.5154 


L c 


2.2180 


-1.9145 


- 3.7399 


-31.1214 


-17.2626 


T s 


0.1979 


-0.1089 


0.0304 


1.9247 


0.8838 


T c 


-0.2013 


- 0.0865 


0.0556 


- 0.0048 


- 0.8626 



Gains and Lag Angles 

The rotor response characteristics are calculated by applying 
equations (2, 3,4) to the test data available. The results 
available are listed in Table V. As pointed out previously, the 
values given include the effect of the actuator used. Some 
general statements can be made. It is obvious that for n = 0. 
the gain and lag angle of the responses to sin 4V- and cos 
4^-type control applications must be the same. For p. j= this is 
no longer true, and one would expect that the spread between 
KjKj and tjTj (see equations (3), (4)) widens with increasing 
advance ratio. Further, according to classical rotor theory which 
neglects blade stall, the nominal collective pitch setting has no 
effect on the frequency response characteristics. 

Generally speaking, the KjK: and Tj t: values given in 
Table V differ very little. It appears however, that at higher 
advance ratios (compare columns for n - 0.849 and 0.851) the 
collective pitch has a larger effect than anticipated. It is also 
possible that the error of the baseline data described in the sum- 
mary may play a role. 

Oscillator Inputs Required 

Equation (10) is used to calculate the inputs required to 

(a) generate unit amplitudes of pure pitching moments, 
rolling moments and vertical forces and 

(b) compensate the existing vibrations 

The results are given in Tables VI and VII. They show that, as 
to be expected, the oscillatory.inputs required for vibration 
reductions generally increase with increasing advance ratio. 
Surprisingly, the rotor collective pitch setting seems to play a 
larger role than the steady lift generated. See also Table VIII 
which summarizes the results obtained and lists the operating 
conditions investigated in the order of decreasing vibrations. 
The first column shows the relative magnitude of the vibratory 
moments generated and the last column the approximate 
amplitude of the blade pitch variation required to compensate 
the vibrations. The amplitude of the pitch variation produced 
per volt oscillator input changes with the control loads and 



265 



TABLE V 
GAINS AND LAG ANGLES DERIVED FROM EXPERIMENTS 











CKp- 


in.-lb/volt, T p - degrees) 










p 


M = 0.191 


H = 0.239 


H = 0.443 


/u= 0.849 


M= 0.851 


K p 


T P 


K p 


T P 


K p 


T P 


K p 


T P 


K p 


T P 


1 


5.617 


42.3 


1.099 


125.6 


2.236 


120.5 


4.798 


72.0 


4.094 


116.5 


2 


6.126 


44.0 


1.141 


149.1 


2.791 


129.3 


4.787 


72.6 


3.487 


135.6 


3 


17.571 


-9.6 


52.416 


-30.1 


42.237 


-28.7 


18.537 


-19.8 


43.319 


-5.1 


4 


26.019 


-45.4 


47.991 


-37.3 


40:073 


-30.1 


20.329 


-41.5 


37.081 


12.7 


5 


30.696 


155.7 


59.416 


182.9 


45.186 


188.4 


33.002 


183.4 


26.170 


214.2 


6 


32.505 


181.7 


77.408 


193.2 


61.144 


180.8 


21.085 


180.0 


38.661 


184.5 


7 


2.856 


136.0 


4.246 


81.9 


8.166 


86.5 


2.472 


102.1 


10.097 


93.1 


8 


1.507 


98.4 


5.083 


67.1 


8.077 


66.9 


3.412 


144.7 


7.979 


62.9 


9 


35.384 


213.4 


59.420 


198.8 


43.846 


181.4 


44.506 


200.5 


48.081 


176.2 


10 


41.674 


185.8 


51.280 


198.6 


39.383 


195.7 


48.473 


201.0 


40.850 


187.7 


11 


45.953 


116.6 


76.875 


108.3 


78.512 


101.8 


67.268 


134.4 


88.540 


94,17 


12 


61.589 


131.5 


86.361 


99.3 


80.995 


95.7 


61.288 


141.5 


90.934 


95t3 


13 


6.879 


45.6 


5.420 


51.4 


8.928 


39.2 


8.188 


35.8 


9.340 


38.5 


14 


7.211 


43.7 


6.195 


46.4 


8.999 


35.9 


8.906 


36.1 


9.651 


35.6 


15 


6.635 


245.2 


4.275 


205.9 


2.571 


195.2 


5.976 


215.0 


3.623 


184.0 


16 


6.033 


218.3 


3.962 


208.1 


3.123 


188.7 


4.775 


229.5 


1.977 


185.4 


17 


13.000 


127.3 


7.596 


94.3 


7.632 


76.7 


13.261 


133.1 


11.188 


86.9 


18 


10.057 


128.6 


8.176 


97.4 


8.381 


92.2 


7.953 


126.3 


11.101 


90.7 



the type of control (0 , 6 s , C ) used. Therefore, the con- 
version factor varies and the last column of Table VIII is 
given only to indicate the approximate amplitudes involved. 

With one exception, the vibratory control applications re- 
quired were smaller than those used for the frequency response 
tests. The exception is the case with the highest vibration level 
encountered for which the compensating controls required were 
approximately 1 5 to 20% higher than the inputs used for the 4P 
frequency response tests. 

Blade Loads 

The calculation of the effect of the compensating control 
inputs on the blade loads is based on Equations ( 1 3 ) and ( 1 4). The 
first step is to calculate, for each specific case, the quantities E n 
through H n (n = 2, 3, 4, 5). See Table IX which refers to n = 0.849 
and lists the sin n 4* and cos n^ components of the various 
loads due to unit control (volt) application. The table shows, for 
instance, that at the advance ratio n = 0.849, a ±1 volt variation 
of „, produces 3P chordwise bending moments of the magnitude 



(-91.77 sin 3^+7.15 cos 3*) in.-lb 



As the control inputs required for vibration reduction have been 
previously calculated, their effects on the blade loads can be de- 
termined by adding up the various contributions. The reader is 
referred to Table X which applies to the flapbending moment 
at 3.3 in.for the case m= 0.849. Given are the original loads 
without vibratory control application, the individual contribu- 
tions and the sum. The last column shows the amplitudes with- 
out and with compensating control input. A summary of the 



loads is represented in Table XI. Generally speaking, chord- 
bending, blade torsion and the 4P flap-bending moments of the 
root flexure increase with increasing advance ratio. The 3 and 5P 
flap-bending moments of the flexure are, by nature, reduced and 
the 2P flap-bending moments are least affected. From the limited 
data available, it appears that the 4P chordwise- and 5P torsion 
moments may be the critical load for this configuration, 
inasmuch as the natural frequencies are close to these 
values. 

As mentioned previously, it is assumed here that 
the pitching and rolling moments are solely caused by the flap- 
bending moments of the root flexure which were individually 
measured and properly combined by a sin-cos potentiometer. 
This means, the only source for the troublesome 4P moments in 
the nonrotating system are the 3 and 5P flap-bending moments at 
3.3 in. For four identical blades, it follows that elimination of 
the 4P pitching and rolling moments requires that the sin 3^, cos 
3^, sin Si> and cos 5^ components of the flap-bending moments 
at 3.3 in. are reduced to zero. As the four blades behave dif- 
ferently, this ideal condition will practically never be fulfilled. 

In the preceding paragraphs the flapbending moment of a 
specific blade, with consideration of the compensating control 
input, was calculated. To a certain extent, these predicted loads 
can be used as an independent check. As an example, the case 
H - 0.849 is treated. According to Table IV the amplitudes of 
the 4P pitching and rolling moments to be compensated are 



M= 20.56 in.-lb 
L = 32.52 in.-lb 



(15) 



266 



The calculated 3 and SP flap-bending moments with considera- The amplitudes of the resulting 4P pitching and rolling moments 

tion of the compensating control input amount to (see Table VII), are 



m 3s = 0.6233 in.-lb 
m 3c = -1.1833 
m 5s = -1.9266 
m 5c = 0.3099 



(16) 



M= 3.14in.-lb 
L = 5.91 in.-lb 



(17) 



TABLE VI 

OSCILLATOR INPUTS REQUIRED (VOLT) TO GENERATE PURE sin #- AND cos #- COMPONENTS 

OF PITCHING MOMENTS, ROLLING MOMENTS AND VERTICAL FORCES 



f 


"■control 


e os 


fl oc 


9 ss 


e sc 


9 cs 


9 cc 


0.191 


"s, control = ' 


+0.0143 


- 0.0485 


+0.0508 


+0.0290 


- 0.0296 


+0.0241 




**c, control = ' 


+0.0117 


-0.0123 


- 0.0055 


+0.0283 


-0.0219 


-0.0098 




*% control = * 


-0.0177 


- 0.0236 


-0.0113 


+0.0052 


-0.0169 


+0.0073 




^c, control = * 


+0.0042 


-0.0071 


-0.0209 


-0.0200 


+0.0003 


-0.0147 




* s, control = * 


+0.0922 


+0.1380 


- 0.0490 


-0.0302 


+0.0252 


-0.0232 




^c, control = ' 


-0.1044 


+0.1164 


+0.0123 


-0.0210 


+0.0235 


+0.0081 


0.239 


™s, control = ' 


+0.0028 


-0.0069 


+0.0299 


+0.0219 


-0.0111 


+0.0211 




"*c, control ~ ' 


+0.0109 


+0.0028 


-0.0096 


+0.0206 


-0.0154 


- 0.0070 




^s, control = ' 


- 0.0023 


-0.0108 


- 0.0056 


+0.0203 


-0.0167 


+0.0078 




^c, control = ' 


+0.0128 


-0.0029 


- 0.0245 


-0.0243 


- 0.0008 


-0.0210 




l s, control * 


+0.1356 


+0.1337 


-0.0053 


-0.0155 


+0.0072 


-0.0128 




* c, control = * 


-0.1436 


+0.1085 


+0.0168 


+0.0070 


+0.0091 


+0.0100 


0.443 


™s, control = ' 


-0.0019 


- 0.0053 


+0.0255 


+0.0116 


- 0.0069 


+0.0145 




"*c, control = * 


+0.0053 


+0.0011 


-0.0023 


+0.0331 


-0.0168 


-0.0004 




^s, control = ' 


- 0.0057 


- 0.0067 


-0.0021 


+0.0253 


-0.0135 


+0.0126 




*% control ~ ' 


+0.0120 


-0.0028 


-0.0155 


-0.0084 


- 0.0093 


-0.0112 




*s, control - * 


+0.1020 


+0.0732 


-0.0088 


-0.0094 


-0.0018 


-0.0138 




* c, control = * 


-0.0714 


+0.0941 


+0.0071 


-0.0108 


+0.0171 


-0.0024 


0.849 


^s, control - ' 


+0.0049 


-0.0240 


+0.0338 


+0.0179 


- 0.0229 


+0.0182 




"*c, control = * 


+0.0149 


-0.0149 


-0.0109 


+0.0487 


-0.0271 


- 0.0222 




*% control ~ * 


-0.0124 


-0.0137 


-0.0120 


+0.0074 


-0.0118 


+0.0024 




**c, control = * 


+0.0052 


-0.0056 


-0.0072 


-0.0121 


+0.0006 


-0.0123 




l s, control * 


+0.1050 


+0.0698 


-0.0211 


+0.0037 


+0.0017 


-0.0214 




*c, control = * 


-0.0772 


+0.1079 


-0.0034 


-0.0305 


+0.0221 


+0.0031 


0.851 


™s, control = * 


+0.0001 


-0.0081 


+0.0191 


+0.0122 


- 0.0077 


+0.0109 




™c, control = ' 


+0.0082 


-0.0055 


-0.0126 


+0.0290 


-0.0135 


- 0.0055 




*% control = * 


- 0.0080 


-0.0107 


-0.0043 


+0.0117 


- 0.0057 


+0.0098 




""c, control = ' 


+0.0113 


-0.0102 


- 0.0069 


+0.0028 


-0.0137 


-0.0037 




* s, control = * 


+0.1016 


+0.0599 


-0.0087 


+0.0109 


-0.0130 


-0.0091 


T — 1 

1 c, control ~ ' 


-0.0682 


+0.0998 


+0.0034 


-0.0143 


+0.0189 


- 0.0058 



! in.-lb 



267 



TABLE VII 

OSCILLATOR INPUTS REQUIRED (VOLT) 

TO COMPENSATE EXISTING 4P- VIBRATIONS 



TABLE VIII 
VIBRATION SUMMARY 



V- 


0.191 


0.239 


0.443 


0.849 


0.851 


6 OS 


0.1683 


0.0394 


0.0146 


0.0457 


0.0300 


e oc 


0.3121 


0.0224 


-0.0490 


0.2354 


-0.2726 


9 SS 


0.1746 


0.0090 


-0.1400 


-0.7980 


-0.3275 


e sc 


-0.0133 


-0.0293 


0.1273 


-0.5881 


0.3498 


9 CS 


0.2052 


-0.0026 


-0.1176 


0.4610 


-0.3549 


9 CC 


-0.0651 


-0.0180 


0.0056 


-0.8308 


-0.0428 



Rel. Vibration 
Level 


M 


flnomi- 
nal 


C T /ff 


Ampl. of Pitch 
Variation 


1 


0.849 


10° 


-0.005 


-3.0° 


0.58 


0.851 


4 


-0.013 


2.0 


0.32 


0.191 


12 


0.102 


0.8 


0.21 


0.443 


4 


0.011 


0.5 


0.08 


0.239 


4 


0.028 


0.2 



Decreasing 
Vibration 
Level ! 



TABLE IX 

EFFECTS OF UNIT 4P OSCILLATOR INPUT ON BLADE BENDING 

AND TORSION MOMENTS (in-lb). ju = 0.849 



M= 0.849 


Input 


sin 2ii 


cos 2^ 


sin 3* 


cos 3^ 


sin 4^ 


cos Ai> 


sin 5* 


cos 5* 




"os 


0.3815 


- 2.6028 


- 1.1212 


+ 1 .9467 


+ 0.0022 


1.6252 


- 0.4640 


+ 0.2286 




#oc 


- 0.7265 


- 0.7428 


- 2.1170 


- 0.9082 


- 1.7646 


0.1744 


+ 0.4336 


- 0.2014 




e ss 


-20.1796 


- 7.1252 


0.4843 


10.9746 


9.2290 


- 1.4705 


-12.1221 


-16.4408 


Flapbending 


9 SC 


1.4455 


- 18.6069 


-11.8793 


0.8771 


1.9670 


9.2946 


+18.4710 


-13.1116 


3.3 in. 


#cs 


-15.0717 


19.2091 


- 1.7568 


+ 13.3006 


4.4390 


13.0827 


24.2022 


-18.4700 




e cc 


-11.0041 


- 12.5052 


-12.2451 


- 3.9250 


-11.5818 


6.8481 


17.1269 


+18.2863 




#os 


- 3.1446 


0.01156 


+ 0.0644 


- 6.4289 


0.5673 


- 5.5966 


- 2.6912 


- 5.2806 




#oc 


0.4488 


.- 3.3139 


+ 5.7587 


- 0.6033 


7.2213 


1.7289 


4.4109 


- 1.9638 


Flapbending 


*ss 


-13.1131 


- 1.6401 


- 9.4439 


11.4718 


2.7493 


1.6368 


20.3552 


30.4485 


13.15 in. 


e sc 


- 3.1093 


- 10.4663 


-13.7168 


- 7.3647 


- 0.7250 


4.6008 


-31.6355 


23.4534 




#CS 


-15.3541 


3.9011 


- 20.8842 


- 14.1583 


- 4.0272 


- 4.6816 


-53.1766 


36.9531 




e cc 


- 3.7738 


- 10.2279 


7.2742 


- 11.8491 


2.4534 


1.0036 


-30.9619 


-33.3918 




e os 


- 5.2318 


5.1653 


18.4997 


- 66.4765 


8.5046 


- 2.0555 


6.0027 


8.6689 




#oc 


- 0.3311 


2.6008 


55.9170 


15.3823 


8.5503 


12.5308 


-10.1401 


4.6381 




0ss 


- 23.2604 


3.6649 


-91.7693 


7.1537 


-12.9172 


- 5.1116 


-13.8450 


7.4174 


Chordbending 


9 SC 


4.7043 


- 8.0015 


-37.9514 


- 71.7419 


6.5301 


-16.8130 


- 4.2184 


-12.8505 


2.4 in. 


0cs 


-25.0714 


15.3009 


- 59.7492 


- 177.5673 


41.5059 


-80.7110 


- 5.8153 


- 27.4052 




e cc 


- 2.0059 


- 7.7253 


77.1483 


- 7.0902 


68.5358 


26.5134 


7.7451 


-28.5566 




60s 


0.1891 


0.0544 


- 0.2460 


0.5652 


- 1.0733 


0.2665 


0.1925 


0.0465 




0OC 


0.0788 


- 0.1531 


- 0.1960 


- 0.2328 


- 0.6076 


- 1.0110 


0.0102 


0.01822 




" d ss 


0.4975 


0.2685 


- 0.9271 


- 1.5838 


- 0.0498 


1.4606 


15.6374 


13.1496 


Torsion 


e sc 


- 0.6976 


- 0.7498 


3.0700 


- 1.4345 


- 1.0039 


0.9952 


-11.8807 


15.1709 


9.28 in. 


#cs 


0.8756 


- 0.0250 


- 1.5421 


- 0.9968 


- 1.9762 


1.0423 


- 14.6088 


21.3914 




#CC 


- 0.8745 


- 0.9375 


2.1226 


- 2.5792 


- 1.3255 


- 1.2713 


- 17.9657 


-13.8937 



268 



Comparison of Equations (15) and (17) shows that the vibratory 
pitching moment is reduced to approximately 15 percent and the 
rolling moment to approximately 18 percent of its original value. 
This indicates that the various blades behave differently and that 
the goal of zero 4P pitch-roll and vertical vibrations is achieved 
by cancellation of the effects of the four blades. 



Analytical Formulation and 
Calculated Results 

The aeromechanical characteristics of the High Advance 
Ratio Model (HARM) has been analytically described in 2 
degrees of freedom. These are based on the first and second flap- 
ping modes which have been approximated by polynomial fits of 



finite element determined mode shapes. The first and second 
mode shape approximations used are given by 



and 



where 



#2 = 2.292x 2 - 1.292x 3 



2 = -10.21x 2 + 20.78x 3 -9.57x 4 



x = r^ the non-dimensional radial station. 
R 



The aerodynamics are based on classical quasi-steady incom- 
pressible strip theory. The reverse flow region is fully accounted 
for, but stall effects have been neglected, as described in Refer- 
ence 5. 



TABLE X 

EFFECT OF VIBRATION COMPENSATION ON FLAPBENDING 

MOMENT (in-lb) AT 3.3 in. y. = 0.849 



n 




cos n 


sin n 


Amplitude 


2 


W/O Vibration Control 


-92.7652 


17.2338 


94.35 




Contribution of O 


- 0.0559 


- 0.1536 






h 


16.6165 


15.2507 






0c 
TOTAL 


19.2393 


2.2002 




- 56.9653 


34.5311 


66.61 


3 


W/O Vibration Control 


- 1.1732 


- 14.7883 


14.83 




Contribution of Q 


- 0.1248 


- 0.5496 






h 


- 9.2715 


6.5928 






6c 
TOTAL 


9.3862 


9.3684 




- 1.1833 


0.6233 


1.34 


4 


W/O Vibration Control 


- 0.1403 


- 3.5448 


3.55 




Contribution of g Q 


0.1153 


- 0.4152 






9s 


- 4.2868 


- 8.5191 






TOTAL 


0.3317 


11.6713 




- 3.9801 


- 0.8078 


4.06 


5 


W/O Vibration Control 


3.2312 


2.2658 


3.95 




Contribution of d s 


- 0.0370 


0.0809 






<?s 


20.8199 


- 1.1807 






»c 
TOTAL 


- 23.7042 


- 3.0926 




0.3099 


- 1.9266 


1.95 



269 



TABLE XI 

SUMMARY OF OSCILLATORY BLADE LOADS (IN.-LB) 

WITHOUT AND WITH VIBRATION COMPENSATION 



Operating 
Condition 


M 


Flapbending at 3.3 in. 


Flapbending at 13.15 in. 


Chordbending at 2.4 in. 


Torsion at 9.28 in. 




0.191 
0.239 
0.443 
0.849 
0.851 


n=2 


n = 3 


n = 4 


n = 5 


n = 2 


n = 3 


n = 4 


n = 5 


n = 2 


n = 3 


n = 4 


n = 5 


n = 2 


n = 3 


n = 4 


n = 5 


Without 
Oscillatory 
Control 
Input 


r 
i 


30.1 
10.5 
16.4 
94.4 
18.9 


4.4 
0.6 
2.7 
14.8 
8.6 


1.6 
0.2 
0.1 
3.6 
1.5 


3.5 
0.9 
1.6 
4.0 
3.1 


16.0 

5.3 

9.2 

55.9 

17.7 


1.9 
1.7 
3.2 
3.6 
4.6 


3.0 
0.9 
0.4 
9.5 

3.4 


4.3 
1.2 
3.5 
5.9 
5.8 


21.0 

4.6 

9.4 

31.5 

17.4 


2.2 

2.0 

1.7 

31.4 

10.9 


8.3 
11.0 
10.5 
13.1 
18.9 


19.4 

2.6 

7.7 

14.6 

10.7 


1.2 
0.5 
0.9 
6.8 
3.3 


0.7 
0.2 
0.6 
4.1 
2.4 


0.4 
0.3 
0.3 
0.9 
0.7 


0.6 
0.2 
0.2 
0.3 
0.4 


With 

Oscillatory 
Control 
Input 


•- 


0.191 
0.239 
0.443 
0.849 
0.851 


29.6 
10.3 
12.3 
66.6 
20.2 


1.1 
0.4 
1.3 
1.3 

2.4 


2.9 
0.3 
1.3 
4.1 
6.5 


0.4 
0.7 
1.1 
2.0 
4.1 


16.1 

5.3 

7.5 

41.7 

16.5 


4.4 
1.9 
2.7 
1.3 
3.8 


5.0 
0.8 
0.5 
2.4 
7.0 


3.0 
1.5 
1.3 
2.1 
2.5 


19.2 

4.7 

7.7 

15.7 

17.5 


22.7 

3.0 

3.5 

68.8 

13.6 


10.9 
11.5 
13.6 
38.9 
75.6 


3.9 
2.1 

8.7 

22.3 

7.0 


1.0 
0.5 
0.8 
6.5 

2.3 


1.2 
0.3 
1.4 
4.4 
4.0 


0.4 
0.3 
1.4 
0.8 
0.7 


4.5 
1.2 
1.7 
3.1 
3.5 



The method of solution provides for intermode harmonic 
coupling through the 1 7th harmonic. This is accomplished by 
obtaining transient solutions of the 2-degree-of-freedom descrip- 
tion of the rotor system described as constant coefficient linear 
differential equations over 1 0-degree sectors of the rotor 
azimuth. 

The values of the coefficients for the system of differential 
equations evaluated in this work have been determined at the 
center of the sectors i. e., at 5°, 15°, 25°, etc. 

The basis for the analytical formulation is founded on 
Shannon's sampling theorem which says that the discrete signal is 
equivalent to the continuous signal, provided that all frequency 
components of the latter are less than 1/2T cycles per second, T 
being the time between instants at which the signal is defined, 
(References 6 and 7 ). Since the solution also provides for a com- 
pletely general transient solution, it can be used to calculate a 
Floquet solution by specializing the initial conditions. This has 
been done for the square spring oscillator case studied by M. A. 
Gockel and reported in the AHS Journal in January 1972. The 
problem statement which is exactly describable by this theo- 
retical method was shown to yield the identical Floquet solutions 
as those reported. It is important to note that should the system 
be unstable, the harmonic balance method of solution would not 
directly reveal this instability. 

Briefly, the initial conditions at the beginning of a sector are 
determined by calculating the terminal conditions for the pre- 
vious sector which are then used to initialize the new sector. It 
has been found that essentially arbitrary conditions can be used 
to start the solution and that excellent steady-state conditions 
have been obtained for the conditions examined in six rotor revo- 
lutions. For each solution case presented, the rotor has been 
solved for eight revolutions to ensure that the second flapping 
mode contribution to the response has converged to a steady- 



state value accurate to at least four significant figures. The pro- 
gram is used to calculate closed-form analytic solutions over each 
10-degree sector and therefore is not dependent on a particular 
method of numerical integration. (See Appendix A.) The 
method, however, when applied to the analysis of steady-state 
conditions, does require that sufficient solution time be calcu- 
lated so that initial transients are dissippated to ensure that 
steady-state equilibrium is achieved (Reference 8). 

The test configuration experimentally examined with re- 
spect to IP flapbending distributions sX\i = 0, including center- 
line measurements, has been compared with this analysis 
procedure on Figure 2, utilizing the two-mode description. This 
is a limited use of the analysis technique to establish test/analysis 
correlation. It is believed that the absence of time-dependent 
aerodynamics quasi-steady, largely accounts for the phase error 
in response. The centerline shaft moment measured was 0.75 of 
the calculated (a = 5.73). This may be due to the relatively low 
inflow of the test condition. 



In general this correlation, including the spanwise distribu- 
tion, appears reasonable. 

The eigenvalues of each 10-degree sector are evaluated as 
part of the method. These are summarized in Tables XII, XIII, and 
XIV versus azimuth the ju = 0.191, 0.45, and 0.85 where the real 
and imaginary parts of the eigenvalues have been normalized by 
the noted natural-mode frequencies. The negative aerodynamic 
spring effects over azimuth 90 < *< 270 as well as the positive 
stiffening from 270 < %<90 are as expected more pronounced 
on the first mode frequency. The effects of reduced aerodynamic 
spring and damping are also seen on the retreating side. These 
results show that both damping as well as frequency variations 
occur around the azimuth which influence the rotor response 
with harmonic excitations. 



270 



3.0 



§2.0 
cc 



ui 

< 

x 
°- 1.0 













) o- 




— o| 




"C 


t •- 










NOTE: PHASE MEASURED IN 

DIRECTION OF ROTATION 
FROM 4/ = 0° 

1 1 1 



TABLE XII 

NORMALIZED EIGENVALUES* AT EACH 10-DEGREE 

AZIMUTHAL SECTOR FOR *i= 0.191 




0.2 0.3 

x = r/R 



Figure 2. One-Per-Rev Blade Radial Flap-Bending Moment 
Distribution at h-=0. 

The rotating frequencies and properties of the flapping 
modes noted in Tables XII, XIII, and XIV analytically describe 
the 7.5-ft-diameter rotor, configuration (5), 500-rotor-rpm con- 
dition for which all harmonic feathering tests were conducted. 
In an effort to further improve analytic correspondence with 
test data the slight change of the second flapping mode fre- 
quency resulted from matching collective blade angle selection 
at the test conditions. Details of the test model are given in 
References 9, 10 and 11. 

The harmonic components of the blade root flap-bending 
moment (OP through 5P) were calculated for these advance ratios 
for unit perturbation of blade feathering angle at 6j c , 0j s , $2v 

e 2s> e 3c< 9 3s» 9 4c> "4s> e 5c> 9 5s> as we . U as for unit chan 8 e in 
6 and a 



The single non-dimensional blade root, centerline flap-bending 
moment harmonic influence coefficients resulting from harmonic 
feathering are summarized in matrix form in Tables XV, XVI, and 
XVII for ft = 0.191, 0.45, and 0.85. These are based on har- 
monic analysis of the moment at each condition for 36 equally 
spaced (10-degrees apart) azimuth intervals. Single-blade 







P = 


1.34 


P = 


6.38 


SECTOR 


xJrO 


R l 


h 


R.o 


h 


1 


5 


-.204 


1.024 


-.155 


1.002 


2 


15 


-.212 


1.022 


-.163 


1.002 


3 


25 


-.220 


1.019 


-.170 


1.002 


4 


35 


-.227 


1.014 


-.177 


1.002 


5 


45 


-.233 


1.007 


-.183 


1.001 


6 


55 


-.238 


.999 


-.187 


1.001 


7 


65 


-.242 


.990 


-.190 


1.000 


8 


75 


-.244 


.980 


-.192 


1.000 


9 


85 


-.245 


.970 


-.193 


.999 


10 


95 


-.244 


.960 


-.192 


.998 


11 


:105 


-.242 


.951 


-.190 


.998 


12 


115 


-.239 


.943 


-.186 


.997 


13 


125 


-.234 


.937 


-.182 


.997 


14 


135 


-.228 


.933 


-.176 


.997 


15 


145 


-.221 


.930 


-.169 


.996 


16 


155 


-.213 


.930 


-.162 


.996 


17 


165 


-.205 


.932 


-.154 


.996 


18 


175 


-.197 


.935 


-.146 


.997 


19 


185 


-.188 


.940 


-.138 


.997 


20 


195 


-.180 


.945 


-.130 


.997 


21 


205 


-.172 


.952 


-.123 


.997 


22 


215 


-.165 


.958 


-.116 


.998 


23 


225 


-.159 


.965 


-.111 


.998 


24 


235 


-.154 


.971 


-.106 


.998 


25 


245 


-.150 


.978 


-.103 


.999 


26 


255 


-.148 


.984 


-.101 


.999 


27 


265 


-.147 


.989 


-.100 


1.000 


28 


■ 275 


-.148 


.995 


-.101 


1.000 


29 


285 


-.150 


.999 


-.103 


1.000 


30 


295 


-.154 


1.005 


-.107 


1.001 


31 


305 


-.158 


1.010 


-.111 


1.001 


32 


315 


-.164 


1.014 


-.117 


1.001 


33 


325 


-.171 


1.018 


-.124 


1.002 


34 


335 


-.179 


1.021 


-.131 


1.002 


35 


345 


-.187 


1.023 


-.139 


1.002 


36 


355 


-.195 


1.025 


-.147 


1.002 



♦SECTOR EIGENVALUES ARE GIVEN BY: 
(Rj+Ij i) (1.34fl) 
AND (R 2 + I 2 i) (6.38S2 ) 

computed root flap-bending moment influence coefficients 
at n = 0.45 are compared with experimental 0.O73R 
single-blade data, in parentheses, from Reference 1 and 1 2. 
in Table XVIII. ' 

These appear reasonable when shear effects are considered. 

It is important that the general character of these influence 
coefficients be established in future tests. These tests should be 
. structured to permit measurement to confirm these distributions. 



271 



TABLE XIII 

NORMALIZED EIGENVALUES* AT EACH 10-DEGREE 

AZIMUTHAL SECTOR FOR n= .45 



TABLE XIV 

NORMALIZED EIGENVALUES* AT EACH 10-DEGREE 

AZIMUTHAL SECTOR FOR M = .85 







P = 


1.34 


P = 


6.2 


SECTOR 


#° 


Ri 


h 


R 2 


h 


1 


5 


-.215 


1.087 


-.167 


1.007 


2 


15 


-.234 


1.088 


-.186 


1.007 


3 


25 


-.252 


1.084 


-.203 


1.007 


4 


35 


-.269 


1.075 


-.218 


1.006 


5 


45 


-.283 


1.059 


-.232 


1.005 


6 


55 


-.295 


1.037 


-.242 


1.004 


7 


65 


-.303 


1.011 


-.250 


1.002 


8 


75 


-.309 


.982 


-.255 


1.000 


9 


85 


-.311 


.951 


-.256 


.998 


10 


95 


-.310 


.920 


-.254 


.996 


11 


105 


-.305 


.891 


-.249 


.995 


12 


115 


-.297 


.867 


-.240 


.993 


13 


125 


-.285 


.850 


-.229 


.992 


14 


135 


-.271 


.839 


-.215 


.991 


15 


145 


-.255 


.837 


-.200 


.991 


16 


155 


-.237 


.842 


-.182 


.991 


17 


165 


-.218 


.854 


-.164 


.992 


18 


175 


-.197 


.870 


-.145 


.992 


19 


185 


-.177 


.889 


-.126 


.993 


20 


195 


-.158 


.909 


-.108 


.994 


21 


205 


-.139 


.928 


-.092 


.995 


22 


215 


-.123 


.945 


-.078 


.996 


23 


225 


-.109 


.960 


-.068 


.997 


24 


235 


-.098 


.972 


-.061 


.998 


25 


245 


-.089 


.982 


-.057 


.998 


26 


255 


-.085 


.990 


-.056 


.999 


27 


265 


-.083 


.997 


-.055 


1.000 


28 


275 


-.084 


1.003 


-.056 


1.001 


29 


285 


-.089 


1.011 


-.058 


1.001 


30 


295 


-.078 


1.018 


-.062 


1.002 


31 


305 


-.108 


1.027 


-.069 


1.003 


32 


315 


-.122 


1.038 


-.079 


1.003 


33 


325 


-.138 


1.049 


-.093 


1.004 


34 


335 


-.156 


1.061 


-.110 


1.005 


35 


345 


-.175 


1.072 


-.129 


1.006 


36 


355 


-.195 


1.081 


-.148 


1.006 



♦SECTOR EIGENVALUES ARE GIVEN BY: 
(Rj+Ij i) (1.340) 
AND (R 2 +I 2 i) (6.20 U) 







P 1 = 


1.34 


P 2 = 


6.20 


SECTOR 


\]/0 


*i 


II 


R 2 


h 


1 


5 


-.231 


1.192 


-.040 


1.014 


2 


15 


-.267 


1.209 


-.048 


1.015 


3 


25 


-.301 


1.212 


-.055 


1.016 


4 


35 


-.332 


1.200 


r.061 


1.015 


5 


45 


-.360 


1.171 


-.067 


1.013 


6 


55 


-.382 


1.126 


-.071 


1.010 


7 


65 


-.399 


1.065 


-.074 


1.006 


8 


75 


-.409 


.992 


-.076 


1.002 


9 


85 


-.413 


.911 


-.076 


.997 


10 


95 


-.411 


.826 


-.076 


.992 


11 


105 


-.402 


.745 


-.073 


.988 


12 


115 


-.387 


.675 


-.070 


.984 


13 


125 


-.366 


. .625 


-.065 


.982 


14 


135 


-.339 


.603 


-.060 


.980 


15 


145 


-.308 


.611 


-.053 


.980 


16 


155 


-.274 


.645 


-.040 


.981 


17 


165 


-.237 


.698 


-.039 


.983 


18 


175 


-.199 


.759 


-,031 


.986 


19 


185 


-.160 


.822 


-.023 


.989 


20 


195 


-.123 


.879 


-.016 


.992 


21 


205 


-.090 


.925 


-.012 


.993 


22 


215 


-.062 


.954 


-.011 


.994 


23 


225 


-.043 


.970 


-.012 


.996 


24 


235 


-.034 


.977 


-.014 


.997 


25 


245 


-.032 


.983 


-.015 


.998 


26 


255 


-.032 


.990 


-.015 


.999 


27 


265 


-.033 


1.000 


-.016 


1.000 


28 


275 


-.032 


1.009 


-.015 


1.000 


29 


285 


-.032 


1.016 


-.015 


1.001 


30 


295 


-.034 


1.021 


-.014 


1.002 


31 


305 


-.043 


1.028 


-.012 


1.004 


32 


315 


-.061 


1.040 


-.011 


1.006 


33 


325 


-.088 


1.063 


-.013 


1.007 


34 


335 


-.120 


1.094 


-.017 


1.008 


35 


345 


-.156 


1.130 


-.024 


1.010 


36 


355 


-.193 


1.165 


-.032 


1.012 



*SECTOR EIGENVALUES ARE GIVEN BY: 
(Rj± Ij i) (1.34S2) 
AND (R 2 ±I 2 fi) (6.20 £2) 



272 



TABLE XV 



:EM. _ BLADE ROOT (STA 0) BENDING MOMENT INFLUENCE COEFFICIENT MATRIX FOR n = 0.191 
a " • (Pj = 1.34, P 2 = 6.38) 





cpo 


cp lc 


c Pis 


CP 2 C 


C P2S 


C P3C 


C P3S 


C P4C 


cp 4 s 


C P5C 


C P5S 


Aa 


.0034 


.0009 


-.0016 


-.0001 


-.0001 




















A6 


.0132 


.0049 


-.0111 


-.0007 


-.0005 


-.0001 

















A61S 


.0036 


.0057 


-.0225 


-.0018 


-.0003 





.0002 














AG1C 


-.0005 


-.0213 


-.0045 


-.0001 


.0018 


.0002 

















A62S 





-.0056 


-.0004 


.0018 


.0031 


.0014 


-.0009 


.0002 


.0003 








A62C 


-.0003 


-.0005 


.0057 


.0031 


-.0018 


-.0009 


-.0014 


.0003 


-.0002 








A63S 





-.0001 


.0004 


.0001 


.0002 


-.0046 


-.0012 


.0015 


-.0014 


.0001 


.0003 


A63C 





.0004 





.0002 


-.0001 


-.0012 


.0046 


-.0014 


-.0015 


.0003 


-.0001 


A94S 














.0001 


-.0010 


.0017 


-.0076 


-.0026 


.0017 


-.0020 


A64C 











.0001 


-.0001 


.0017 


.0009 


-.0026 


.0076 


-.0020 


-.0017 


A65S 


.0119 














.0002 


.0002 


-.0015 


.0024 


-.0101 


-.0033 


A65C 

















.0002 


-.0002 


.0024 


.0015 


-.0033 


.0101 



TABLE XVI 
C RM _ BLADE ROOT (STA 0) BENDING MOMENT INFLUENCE COEFFICIENT MATRIX FOR n =.45 
aff (Pj = 1.34, P 2 = 6.20) 





CPo 


cp lc 


CP 1S 


cp 2C 


C P2S 


C P3C 


C P3S 


cp 4C 


CP4S 


CP5C 


■ c Pss 


Aa 


.0085 


.0053 


-.0089 


-.0010 


-.0012 


-.0004 


-.0004 


-.0001 


-.0001 








A6 


.0160 


.0135 


-.0276 


-.0038 


-.0029 


-.0009 


-.0004 


-.0001 


-.0002 








A61S 


.0087 


.0102 


-.0292 


-.0048 


-.0016 


-.0004 


.0007 


-.0006 











A61C 


-.0011 


-.0226 


-.0034 


-.0004 


.0046 


.0011 


.0002 


.0001 


.0003 








^G2S 





-.0130 


-.0006 


.0006 


.0046 


.0036 


-.0020 


.0009 


.0015 


-.0002 


.0002 


A62C 


-.0015 


-.0024 


.0142 


.0043 


.0001 


-.0020 


-.0035 


.0015 


-.0009 


.0001 


.0002 


A63S 








.0023 


.0004 


.0010 


-.0057 


-.0032 


.0038 


-.0036 


.0008 


.0016 


A63C 


-.0002 


.0022 


-.0002 


.0007 


-.0001 


-.0033 


.0057 


-.0036 


-.0038 


.0016 


-.0008 


A64S 





-.0001 





.0003 


.0007 


-.0026 


.0042 


-.0090 


-.0046 


.0043 


-.0048 


A64C 











.0004 





.0042 


.0026 


-.0046 


.0090 


-.0048 


-.0043 


A65S 














.0003 


.0008 


.0008 


-.0038 


.0058 


-.0118 


-.0055 


A65C 














.0003 


.0008 


-.0009 


.0058 


.0038 


-.0054 


.0118 



TABLE XVII 
d|M. - BLADE ROOT (STA 0) BENDING MOMENT INFLUENCE COEFFICIENT MATRIX FOR ju = .85 

(Pi = 1.34, P 2 = 6.20) 





CP 


C PlC 


C PlS 


C P2C 


C P2S 


cp 3C 


C P 3 S 


C(3 4C 


CP4S 


C P5C 


C P5S 


Ace 


.0201 


.0227 


-.0296 


-.0056 


-.0102 


-.0037 


-.0039 





-.0021 





-.0015 


A6 


.0253 


.0378 


-.0598 


-.0141 


-.0155 


-.0061 


-.0039 





-.0032 


-.0003 


-.0018 


A61S 


.0192 


.0278 


-.0490 


-.0117 


-.0114 


-.0036 


-.0004 


-.0019 


-.0014 


-.0005 


-.0001 


A81C 


-.0024 


-.0258 


-.0015 


-.0012 


.0085 


.0035 


.0008 


.0003 


.0022 





.0009 


Afi2S 


-.0006 


-.0229 


-.0007 


-.0026 


.0079 


.0081 


-.0034 


.0024 


.0054 


-.0019 


.0016 


A62C 


-.0056 


-.0110 


.0308 


.0084 


.0067 


-.0026 


-.0064 


.0052 


-.0019 


.0013 


.0023 


A63S 


-.0003 


-.0014 


.0076 


.0010 


.0035 


-.0088 


-.0082 


.0100 


-.0084 


.0033 


.0049 


A83C 


-.0009 


.0060 





.0029 


-.0009 


-.0084 


.0089 


-.0084 


-.0100 


.0048 


-.0033 


A64S 


-.0005 


-.0004 


.0003 


.0013 


.0008 


-.0067 


.0087 


-.0135 


-.0100 


.0112 


-.0100 


A94C 


-.0004 


.0002 


-.0002 


.0007 


-.0014 


.0087 


.0067 


-.0100 


.0134 


-.0101 


-.0112 


A65S 





.0001 


.0006 


.0002 


-.0003 


.0029 


.0023 


-.0088 


.0124 


-.0167 


-.0115 


A85C 





.0006 


-.0002 


-.0003 


-.0002 


.0023 


-.0029 


.0124 


.0088 


-.0116 


.0167 



273 



TABLE XVIII 
BLADE ROOT (STA 0) BENDING MOMENT (IN-LB)/DEG INFLUENCE MATRIX FOR n •■ 

( £2= 52.36, Pj = 1 .34, P 2 = 6.20) 



.45 





Po 


Pic 


Pis 


P2C 


P2S 


P3C 


Pas 


P4C 


P4S 


P5C 


P5S 


LIFT 


Aa 


19 


12 


-20 


-2 


-3 


-1(1) 


-1 (-2) 








0(1) 


0(-D 


6 


A6 


36 


31 


-62 


-9 


-7 


-2(1) 


-KD 








0(1) 


0(0) 


10 


A61S 


20 


23 


-66 


-11 


-4 


-KD 


2 (-2) 


-1 





0(1) 


0(0) 


6 


A61C 


-2 


-51 


-8 


-1 


10 


3(0) 


0(1) 





1 


0(0) 


0(0) 





A92S 





-29 


-1 


1 


10 


8 


-5 


2 


3 


-1 








A02C 


-3 


-5 


32 


10 





-5 


-8 


3 


-2 








-1 


A63S 








5 


1 


2 


-13 


-7 


9 


-8 


2 


4 





A63C 





5 


-1 


2 





-7 


13 


-8 


-9 


4 


-2 





A64S 











1 


2 


-6(0) 


9 {6) 


-20 (-8) 


-10 (-5) 


10 (-5) 


-11 (-1) 





A64C 











1 





9(6) 


6 (-2) 


10 (-4) 


20(7) 


-11 (-6) 


-10(3) 





A65S 














1 


2 


2 


-9 


13 


-27 


-12 





A65C 














1 


2 


-2 


13 


9 


-12 


27 






Full-Scale Control Loads 

The feasibility of active vibration attenuation depends on 
the capability of the rotor to generate cancelling shaft moments 
and shears while control forces and displacements remain within 
acceptable limits. 

Since full-scale data are the most relevant from the stand- 
point of hardware test background, the CL 840/AMCS 
(Advanced Mechanical Control System) Cheyenne rotor 
configuration, at a gross weight of 20,000 and with a rotor shaft 



moment of 100,000 in.-lb, was analyzed for hovering flight to 
gain a numerical measure of how loads compare with limits. In 
this analysis three higher harmonic blade-feathering excitations, 
3P, 4P and 5P, were examined to determine the relationships 
among control loads, shaft moments and shear forces. The 
Lockheed Rotor Blade Loads Prediction Model was used for 
this analysis; 68 finite elements were used to describe the system. 
The calculated results, based on 1 -degree excitation levels, are 
summarized in Table XIX. 



TABLE XIX 

CL 840 ANALYSIS - 

SHAFT AND BLADE LOADS DUE TO ONE-DEGREE 

OF HIGHER HARMONIC BLADE-FEATHERING MOTIONS 





FEATHERING FREQUENCY 




3V 


4V 


5* 


Endurance 
Limit, in; -lb 




Amplitude 


Phase 


Amplitude 


Phase 


Amplitude 


Phase 


Shaft Forces 

4P H-force 

4P Y-force 

4P Pitching Moment 

4P Rolling Moment 

4P Thrust 


3801b 
3801b 

22,000 in.-lb 

22,000 in.-lb 




61° 
84° 
83° 
16° 


401b 
401b 




3000 lb 


59° 
83° 

40° 


3101b 
3101b 

108,000 in.-lb 

108,000 in.-lb 




34° 

12° 

8° 

76° 


\ 325,000 


Blade Root Torsion * 
Harmonic 

Steady 
IP 
2P 
3P 
4P 
5P 


-3800 in.-lb 
210 in.-lb 
80 in.-lb 
1500 in.-lb 
130 in.-lb 
20 in.-lb 


11° 
49° 
15° 
47° 
27° 


-4000 in.-lb 
210 in.-lb 
50 in.-lb 
70 in.-lb 
13,300 in.-lb 
80 in.-lb 


11° 
42° 
82° 
88° 
35° 


-3900 in.-lb 
220 in.-lb 
50 in.-lb 
40 in.-lb 
400 in.-lb 
7800 in.-lb 


11° 
39° 
84° 
57° 
10° 


\ 15,500 



Pitch link forces are internal loads between the blade and 
swashplate and therefore self-cancelling. 



274 



The calculated root torsion moments shown in the table 
reflect both the feathering moments at the primary exciting 
frequency and the interharmonic coupling terms; as expected, 
the latter are considerably less. Pitch link loads can be 
determined by multiplying the root torsion moment by 0.1 (to 
account for all applicable geometry); endurance limit of the 
pitch link load is 1550 pounds. 

The 7.5-foot hingeless rotor model data showed that 0.2 to 
0.6-degree cyclic angle excitation levels were required. Study of 
CL 840 test data indicates that similar blade excitation would 
be expected with a full-scale, four-bladed rotor. The CL 840 
data are not yet published in documents that can be referenced, 
however, this material is expected to be published during 1974. 

In summary, full-scale data founded on endurance limit 
considerations indicate that internal blade loads and control 
loads will not limit the trim flight use of periodic variation of 
conventional controls for vibration attenuation. 



Conclusions 

The present report is a preliminary evaluation of the con- 
cept of vibration reduction by properly selected oscillatory col- 
lective and cyclic control applications. The investigations are 
based on experimental frequency response data covering advance 
ratios from approximately 0.2 to 0.85. 

Because there was no instrumentation for the measurement 
of the pitch and roll vibrations, these values were obtained by 
properly adding up the flap-bending moments at 3.3 inches. Any 
other quantity representing pitch/roll vibrations can be 
compensated for in the same fashion. 

The calculated control inputs required for vibration reduc- 
tion stay within acceptable limits. For four of the five conditions 
tested they are smaller than the values used for the frequency 
response tests. The blade pitch variations required for vibration 
alleviation vary, depending on the advance ratio, less than 1 ° for 
.2 < m < -45 and ~ 3° for /u = .85. 

As to be expected, the compensating controls greatly affect 
the blade loads, i.e., torsion, flap- and chordwise bending. With 
regard to flap-bending at 3.3 inches (root flexure), the following 
statements can be made : 

• 3 and 5P flap moments were, by command, 
drastically reduced 

• 2P flap moments were least affected. These 
were the largest oscillatory loads. 

• 4P flap moment increments generally increased 
with increasing advance ratio, but were small 
relative to the 2P flap moments. 

As a general rule, chordwise bending and blade torsion 
increments also increase with the advance ratio. At lower \x 
values the loads are not critical. It is concluded that the 
concept investigated is primarily suited for low and medium 
advance ratios, i.e., for the speed-range of present day 
rotary wing aircraft. The latter application appears promising 
and further studies and tests are suggested. Instrumentation 



to determine rotor vertical and inplane shear forces should 
be incorporated in such future tests. Also a system with a 
first inplane frequency in the vicinity of 1 .5P in combination 
with a flapping frequency of 1.1 P should be tested at con- 
ventional advance ratios to provide experimental data 
representative of current designs. 



References 



1 . London, R. J., Watts, G. A., and Sissingh, G. J., EXPERI- 
MENTAL HINGELESS ROTOR CHARACTERISTICS AT 
LOW ADVANCE RATIO, NASA CR-1 148A, December 
1973. 

2. USAAVLABS Technical Report 69-39, SUPPRESSION OF 
TRANSMITTED HARMONIC VERTICAL AND INPLANE 
ROTOR LOADS BY BLADE PITCH CONTROL, Balcerak, 
J. C, and Erickson, J. C, Jr., Ft Eustis, Virginia, July 1969. 

3. ASRL - Technical Reference 150-1, HIGHER HARMONIC 
BLADE PITCH CONTROL FOR HELICOPTER, Shaw, 
John Jr., Massachusetts, December 1968. 

4. USAAVLABS Technical Reference 70-58, WIND TUNNEL 
INVESTIGATION OF A QUARTER-SCALE TWO- 
BLADED HIGH-PERFORMANCE ROTOR IN A FREON 
ATMOSPHERE, Lee, Charles; Charles, Bruce, and Kidd, 
David, Ft Eustis, Virginia, February 1971. 

5. Sissingh, G. J., DYNAMICS OF ROTORS OPERATING AT 
HIGH ADVANCE RATIOS J. American Helicopter Society, 
13(3) July 1968. 

6. C. E. Shannon, COMMUNICATION IN THE PRESENCE 
OF NOISE, Proc. IRE, Vol. 37, January 1949, p.ll. 

7. DeRusso, P. M., Roy, R. J., and Close, C. M., STATE VARI- 
ABLES FOR ENGINEERS, New York, John Wiley and 
Sons, Inc. 1967, p. 6-9. 

8. Donham, R. E., Subsection titled "RESPONSE OF HELI- 
COPTER ROTOR BLADES TO GUST ENVIRONMENTS" 
in NUCLEAR HARDENING SURVIVABILITY DESIGN 
GUIDE FOR ARMY AIRCRAFT. This report is being pre- 
pared by the B-l Division of Rockwell International under 
Army Contract DAAJ02-73-C-0032. 

9. Kuczynski, W. A., Sissingh, G. J., RESEARCH PROGRAM 
TO DETERMINE ROTOR RESPONSE CHARACTER- 
ISTICS AT HIGH ADVANCE RATIOS, LR 24122, 
February 1971, prepared under Contract NAS 2-5419 for 
U. S. Army Air Mobility Research and Development Lab- 
oratory,Ames Directorate, Moffet Field, California. 

10. Kuczynski, W. A., Sissingh, G. J., CHARACTERISTICS OF 
HINGELESS ROTORS WITH HUB MOMENT FEEDBACK 
CONTROLS INCLUDING EXPERIMENTAL ROTOR 
FREQUENCY RESPONSE, LR 25048, January 1972. pre- 
pared under Contract NAS 2-5419 for U. S. Army Air 
Mobility Research and Development Laboratory,Ames 
Directorate, Moffet Field, California. (Volumes I and 11). 



275 



1 1 . Kuczynski, W. A., EXPERIMENTAL HINGELESS ROTOR where cr { , o- 2 , a„ are all distinct, this yields 

CHARACTERISTICS AT FULL SCALE FIRST FLAP 

MODE FREQUENCIES LR 25491, October 1972, prepared 

under Contract NAS 2-541 9 for U. S. Army Air Mobility p(t) 

Research and Development Laboratory,Ames Directorate, 

Moffet Field, California. 



n P(o- k ) 

w W 6 



<*kt 



(4) 



1 2. Watts, G. A. and London, R. J., VIBRATION AND LOADS 
IN HINGELESS ROTORS, Vol. I and II, NASA 
CR-1 14562, September 1972. 



Appendix A 

The transient response solution of a system described by 
constant coefficient linear differential equations is developed 
in this appendix. The single-degree-of-freedorn case with arbi- 
trary initial conditions and solution of the general case for an 
nth order system with both zero and nonzero initial conditions 
is reported. 

Given the single degree of freedom: 



In the case cited 

Q(s) = A(s)(s-or)(s- Y ) 



where 



«j = 



o/2 ~ ot 
<>3 = V 



and 

P(s) 

Therefore 



L + p(0) A s l + p(0)B + p(0)A s 



A 14+ B ^r + C P = F W 



(i) 



df 



dt 



where A, B, and C are constants, then 

AX^f) +BX(§) + CX(p) = £(F(t)) 

where X. is the Laplace transform operator. This yields 
(As 2 + Bs + C) p(s) = F(s) + p(0)(As + B) + p(0)A 



(2) 



B(s) = F(s) + P(0) (As + B) + p(0)A 
As 2 + Bs + C 

If a positive constant step load of magnitude + L is the form of 
F(t), then 

£(F(t)) = F(s) = ±i 



and 



p(s) 



p(0)As 



A(s) (s - a) (s - -y) A(s - a) (s - y) 

, P(0)B + p(0)A 
A(s- a)(s-y) 

Where p(0) and p(0) are the values of the variable p at time 
t = and c, y are the roots of s 2 + Bs/A + C/A, p(s) trans- 
formed back into the time plane is accomplished through use 
of the inverse Laplace transform of the form P(s) 

Q(s) 

where 

P(s) = polynomial of degree less than n 



and 



Q(s) = (s-a 1 )(s-a 2 ) (s-of n ) 



p(t) = ^ W e 



p(t) = 



A(-a)(- Y ) 



t [L + [p(0)A] a 2 + [p(0)B + p(0)Al a\ at (5) 
[ A(+ o)(+ o, - y ) J 6 



| [L + [p(0) A] V 2 + [p(0)B + p(0)Al y 



f 



(A)(+y)(y-«) 



^ 



Extension to the general case is accomplished as follows. 
Given the general determinantal equation: 



Js 2 [a] + s[b] + [c]J jp(s)j = (f(s) 



(6) 



Where the elements of matrix A, B, and C are constants, 
using Cramer's Rule: 



PjGO 



Denominator with 
Column i replaced by F(s) 



(7) 



s 2 A + sB 



[c] 



' ' Expanding 
yields 
where 



| S 2[A] + s[B] + [C]| 
AqCs-ojXs-^) • • • (s-a n ) 



(8) 



Coefficient of highest power term 



cxj (i = 1 . . . n) are the eigen values (roots) of 
the determinantal equation 



276 



Case 1 - Zero Initial Conditions 



Case 2 - Nonzero Initial Conditions 



Assume p^O) and (3^0) for all i are both zero and that a 
positive unit load acts on p e and that the response of Pf is to 
be determined. Then 



+ 1/s in row e with all other 
rows equal 0_ 



Defining 



F(s) 

s 2 [a] + s[b] + [c] 



(e,fj 



(9) 



as the original determinantal equation with Row e and Column 
f removed and all the remaining rows and columns moved up 
and to the left, respectively, this forms a determinantal equation 
of one less order. 

Based on the earlier development in the s-plane 

X ( Pf(t) ) -J2. + _!L. + 4_ + ... + j2_ 

*"- v 1 / s s-aj s-<*2 s ' a n 



and in the time plane 



«it 



0">t 



o„t 



where 



3f(t) = a + aje ' + a 2 e l + . . . + a n e n 



(.I)(e + f)D(o) ej f 



and 



l o - 


n 




A o n«i 

i=l 




(-l)(e + f)D( 0j ) e;f 


l 3 


n 




ot: A n («; -a,) 

i=l 



ir J 



A is determined by the relationship 



D(o) = A Q n^i 
i=l 



(10) 



The general form of F(s) now becomes: 



F(s)j =j^[ + j 4 A ) + [B][ jpjCO) 
+j[A]jp>) 



(11) 



where Lj are the forces applied at each coordinate pj and pj(0) 
and pj(0) are the positions and rates of the coordinates at time 
zero (initiations of the solution). In this case place the column 
s |F(s)j into the column location of the coordinate for which the 
response is desired without reduction of the order. Then 



P(s) 



Column i 



s F( S ) 



(12) 



where all other terms are 



and 



s 2 [a] + s[b] + [c] 



Q(S) = A (s-«o)(s-ai).,.(s-« n ) 



(13) 



where the a's are the eigenvalues of the determinantal 
equation 

s|s 2 [a] + s[b] + [c] |j --= 



Then 






(14) 



D(o) e f and D(<a:) e j- are formed from the original determinantal 

equation with Row e and Column f removed and all the remaining where s = and the remaining eigenvalues of the general deter? 

minantal equation form the set of aj-'s, and Aq is determined 

by the relationship 



rows and columns moved up and to the left, respectively, 
evaluated at o and a:. The a: are the roots of the original deter- 
minantal equation before Row e and Column f were removed. 
These roots are assumed distinct, an unimportant limitation for 
most physical systems. Note that this solution does not preclude 
instability either aperiodic or oscillatory. 

In practice the eigenvalues are obtained prior to the forma- 
tion of the coefficients and are examined to verify the distinct 
character of the eigenvalues. 

Scalar multiplication of this solution provides the result 
for the nonunit loading case. Summation of solutions obtained 
for loadings at each coordinate can be used to provide the 
general solution for this case where Pj(0) and Pj(0) for all i 
are both zero, i.e., that the initial conditions at time zero 
are all zero. 

In most applications the restriction that the initial condi- 
tions are zero is an unacceptable constraint and this condition 
has been relaxed; the solution follows. 



D(0) = A n c4 
i=l 



(15) 



as given in Case 1 . 



277 



FOREWORD TO THE SUPPLEMENT 



This supplement includes questions and answers following the papers of Sessions I 
through IV and all of the material of Session V. Questions and answers, as well as panel members' 
prepared comments were transcribed from tape recordings. This material has been carefully 
checked and minor changes have been made for clarification. Where the meaning may have been 
ambiguous, editorial comments are bracketed. Panel members have checked their comments for 
accuracy and made minor corrections in the transcript where required. 

R. A. Ormiston 
Technical Chairman 



279 



WELCOME 

Clarence A. Syvertson 

Deputy Director 

Ames Research Center, NASA 

Dr. Mark is out of town, so I have been asked to substitute. I guess most of you have been here at Ames before but in spite of that I 
would like to welcome you to the Ames Research Center. I'm very pleased, the whole center is very pleased, that the specialists in this 
field of rotor dynamics picked Ames for the site of this meeting. I'm certainly not an expert in your field but from all I can see the 
field of rotor dynamics represents one of the most technically difficult and most challenging fields in modern aeronautics. I looked 
over the papers that will be presented and it seems, in spite of all the difficulties, that some real progress is being made in the field. I 
think that is very encouraging. I also noticed that you have papers by representatives of virtually every major rotorcraft manufacturer 
in the country, by representatives from the Ames and Langley Research Center, and by representatives from the Ames, Langley, and 
Eustis Directorates of the U. S. Army Air Mobility R&D Laboratory. I think that it's this broad representation from all the 
organizations throughout the country concerned with these problems that really makes meetings like this especially fruitful. I hope 
you find the papers interesting and the meeting productive. And again, in spite of the fact that you've probably been here before, I'd 
like to welcome you to Ames Research Center. Thank you. 

OPENING REMARKS 

E. S. Carter, Meeting General Chairman 

Chief of Aeromechanics 

Sikorsky Aircraft 

First, I would like to acknowledge the indispensable contributions from NASA that have made this meeting possible. The AHS is 
not an affluent organization and NASA has provided not only the facilities, but the printed brochure, the manpower to staff the 
registration tables, the bus service for the tours, and our hardworking administrative chairman, Jim Biggers. When the meeting was first 
conceived by Bob Wood's AHS Dynamics Committee, it was hoped that the Army could also co-sponsor this meeting. A significant 
factor in selecting Ames as a location was the presence here of the rotary wing dynamics research team in the Army Ames Directorate 
which has probably the most sharply focused rotorcraft dynamics program to be found anywhere within the government research 
agencies. It developed that the Army cannot officially co-sponsor a meeting such as this, but they have provided the technical 
chairman, Dr. Bob Ormiston. To Bob must go the credit, not only for having a good bit to do with initiating the meeting in the first 
place, but for following through with the excellent technical program which you are about to hear. 

The rotorcraft dynamics problem, which this meeting addresses, is perhaps the most challenging, most complex and technically 
sophisticated challenge that can be found short of the biological sciences. The problem is well illustrated by Slide 1 which I have 
borrowed from a paper by Bob Tapscott at the Civil Transport meeting at the Langley Center in November 1971. It also illustrates all 
of the ingredients of the problem that we will be addressing in the next two days: the air mass dynamics problem , which because of its 
four dimensional (time variant) characteristics, virtually defies visualization; the lifting surface problem with its skewed flow, unsteady 
effects, and centrifugally pumped boundary layer which can't possibly be reproduced in two dimensional wind tunnel tests; the blade 
dynamics, complicated not only by the centrifugal field, but by the difficult coupling effects introduced by built-in or elastic twist; 
the fixed to rotating coordinate transformation problem that immensely complicates the airframe and rotor interactions; and finally, 
the aeroelastic characteristics of the body itself with its large concentrated masses, unsymmetrical offsets and very large cutouts. 



DYNAMICS COMPLEXITY 



IMPACT LOADS 




FLEXIBLE BLADES 
COUPLED MODES 



STRUCTURAL MODES 



For the next two days we'll be assessing the state of the art 
and our ability to handle each of these problems and on Friday 
we will have a chance to back off and overview the whole situa- 
tion. Bob Ormiston's paper in this final session is, as far as I 
know, unique in the comparison it makes between all of our 
competitive methods addressed to a single problem. Finally, in 
the panel sessions, our ultimates customers, the designers and the 
service users, will be given an opportunity to tell us what we're 
doing wrong. 



^^ BLACK BOXES 



Slide 1. 



280 



DINNER ADDRESS 

WHAT CAN THE DYNAMICIST DO FOR FUTURE ARMY AIRCRAFT? 

PaulF. Yaggy 

Director, U. S. Army Air Mobility R&D Laboratory 

Moffett Field, California 

The terms "rotary wing aircraft" and "dynamics" are synonomous. All dynamics are not rotary wing, but all rotary wings are 
dynamic. Every consideration of the rotorcraft structure includes dynamic phenomena in some form. One cannot talk of the utility 
and economy of rotorcraft without considering the impact of dynamics on structural vibrations, passenger comfort and ride quality, 
pilot handling qualities, safety, and wearout and life cycle of critical components. Some have been so derisive in their comments as to 
say that the rotorcraft in many respects is an effective inherent fatigue testing machine. 

Now you may object to an analogy as harsh as that, but you would be hard pressed not to admit that there is much semblance in 
fact to support it. For this beautifully sophisticated and integrated machine with its vertical flight capability has a nasty attribute of 
creating its own hostile environment as it attempts translation flight. We speak of the rotor operating through the vortices shed from 
its rotating blades, producing large amplitude nonsteady loads; not unlike a wheeled vehicle bouncing over a corduroy road. This 
would be problem enough, but these first order loads produce highly interactive, coupled phenomena throughout the structure from 
the rotor through the shaft to the drive system, the fuselage, the control system, the instruments, and even the pilot's posterior resting 
on his seat. All of these motions are important, but their relative importance is not easily determined. Ability to adequately account 
for dynamic phenomena has had a pronounced influence on the development of rotorcraft, particularly on its most prevalent 
derivative, the helicopter. 

It is both interesting and revealing to consider the role of dynamics in the history and development of rotary wing aircraft. For 
convenience, let us consider them in five decades from that before 1 940 to the current decade of the '70's. 

Prior to 1940, development efforts were more of a novelty than an orderly plan. The Berliner, Focke, Flettner, Sikorsky, and 
others made brilliant achievements in vertical flight, while Cierva identified some of the basic problems and restrictions of rotary wings 
with his autogyro. The immediate goal of this period was simply the demonstration of some type of successful, controlled flight, an 
elusive goal which many failed to attain. It is interesting that many of the dynamic problems we wrestle with today were readily 
identified in that period. Rotor vibration, high blade vibratory stresses, "ground resonance," air resonance, blade flutter, stick 
vibration, and short life of critical components were all readily apparent. Lacking sophisticated analytical tools and methods, the 
stalwart pioneers of the day turned to empirical approaches which further proved their genius as they incorporated dampers, blade 
balancing techniques, and other modifications to improve system dynamics of their marvelous machines. 

Based on these early efforts, the decade from 1940 to 1950 witnessed a great acceleration of more orderly development for the 
helicopter. So many advances were made in this period that it is almost startling to consider them in retrospect. The full range of rotor 
configurations was investigated; single, coaxial, tandem, and jet driven rotors were experimented with by industrial groups such as 
Bendix, Platt-Lepage, McDonnell, Aeronautical Products, Piasecki, Hiller, Kaman, Bristol, Cierva, and others. Some that reached 
production included Flettner, Focke, Sikorsky, and Bell. 

Novelty gave way to utility in this decade, even though much was exploratory in nature. Included were wire laying, shipboard 
operations, courier duty, observation and air ambulance operations. Utility was limited by the reciprocating engine with its awkward 
volume envelope and high specific weight. This was but one of the limitations with which the designer had to cope and which 
restricted operational vehicles to maximum capabilities of 80 knots, 6 passengers and significantly reduced altitude and hot day 
performance. This decade, too, was plagued by the now all too familiar dynamic problems of high vibration levels, rotor instabilities, 
blade weaving, blade flutter, and a very limited life of dynamic components. Typical lifetimes ran about 75 hours before removal for 
discard or overhaul; a high price to pay for the unique vertical flight capability. 

Significant research and development efforts were undertaken in this decade to cope with blade structural and dynamic analyses 
and rotor airflow and wakeflow analyses. However, the complexity of the mathematical models overtaxed the computational capabil- 
ities of the day and necessary linearization of the problem masked many of the important characteristics. 

In the decade from 1950 to 1960, the number of variants in helicopter design began to diminish as the most optimal designs 
began to be apparent. Utility was increased by applying new found technology and methodology for design. A new source of power, 
the turbine engine was appearing which would give a more optimum volume envelope and eliminate the vibration input of the 
reciprocating engine. This transition to turbine power, with its improved specific weight and fuel consumption, was destined to 
revolutionize helicopter utility and capability. 

The helicopters of the '50's were produced in large quantities under the impetus of the Korean War and their new found military 
utility. Payload capability increased to 20 passengers. Performance reached the ability to hover at 6000 feet and 75-degrees F. 
Allowable vibration criteria were quantified and defined by military specification. Bold new life goals for dynamic components were 

281 



set at 1000 hours before removal. Although some reached these levels, many still barely attained 250 hours. Utility of the vehicle was 
the primary gain of this period, but the plaguing restrictions resulting from dynamic phenomena still produced high costs and 
restricted performance. 

In the decade just past, 1960-1970, the gas turbine completely supplanted the reciprocating engine in helicopters with a resulting 
increase in payload and range. Payload exceeded 5000 pounds, carried at speeds in excess of 200 miles per hour in test flights. Hover 
was attained at 6000 feet and 95-degrees F. Life of dynamic components was somewhat improved, even to on-condition removal for 
some parts, but the spread of lifetimes for dynamic components still ranged from 250 to 1000 hours. Although the established 
vibration criteria was met in part by some helicopters, vibration problems still plagued the designer. Unpredicted rotor instabilities and 
coupled phenomena still occurred with surprising regularity, quite often with disastrous impact to vehicle and development program. 
Flight speeds were increased by improved power ratios. However, this only served to further increase the dynamics problem, since at 
these higher speeds stall induced loads, stresses and vibrations became the key limiting factors in determining critical speeds and 
maneuvers. Use of the helicopter for combat, with requirements for greater agility, further sharpened the awareness of these limita- 
tions and focused greater attention on the dynamic constraints. 

Now what of this decade in which we reside. Our research and development efforts, based on our newly acquired computational 
capability and advanced technology, have given us promise that we can achieve greatly increased performance, utility, and agility from 
our new helicopter system developments. Driven by military requirements, we have accepted the challenge to also survive in the hostile 
environment of the battlefield. To do this, reliability, maintainability, availability, and detectability all must reach new levels. Among 
these criteria, nonsteady phenomena become most critical factors. Improvements in capability and cost now become the challenge 
primarily of the dynamicist and he will determine success or failure. 

Then where are we now in making these advances? Early helicopters had accepted low component life, high vibration, and 
marginal performance. Predictive techniques at that time were based on relatively simple analytical and experimental models. Techno- 
logical advance was slow, based on the empiricism of rudimentary experiment. We have exploited to the maximum that past 
technology with its simple representations of rotor dynamics and flow fields. The old barrel of empiricism is bare. The greater demands 
for agility, longer life, and lower vibration demand new advances in dynamics from basic and applied research and development. A 
rededication to innovative methodology for aerodynamics, rotor flows, dynamics, and their coupling in interdisciplinary systems is 
necessary to meet the demands of the forthcoming generations of helicopters. Only significant advances in the comprehension of 
dynamic phenomena will meet the need for a technological base for desired future growth of capability. 

Three specific areas of emphasis suggest themselves in considering the scope and direction for future technological development. 

First, intensive effort is required in both the evolution of global analytical models which describe the dynamics of the helicopter 
and in the more specific interpretive models which describe the physics of the phenomena in more detail. Results of both of these 
modelling efforts must be verified continually by experiments with both model and full-scale tests. The improved comprehension of 
physical phenomena from interpretive models must be integrated into the global models in a timely manner as they are verified. 

Second, the adequacy of the forcing functions, which are the inputs for the foregoing models, must be improved. These forces 
result from aerodynamic loads generated on the rotor blades, which are often highly nonlinear in nature. Unlike fixed wing aircraft, in 
helicopter flight these regions of nonlinear loading are penetrated deeply and periodically by the rotor blade. The generation of these 
loads results from the complex flow field in which the rotor operates. Here again, adequate mathematical models must be generated to 
predict these dynamic loadings and the models must be verified by experiment. When correlation is attained, we must assure timely 
efforts to include these descriptions into the inputs of the global and interpretive dynamic analyses. 

Last, but by no means least, we must demonstrate true professional acumen to sensitize our efforts to the gain to be expected. 
Expressed in other terms, we must assess the gain to be made by more accurate models against the cost to obtain that gain. The last 
few percent of accuracy may well not be worth the cost. Our goal should be to produce predictive techniques for the designer to 
assure sufficient accomplishment of performance goals without the surprises of instabilities and shortened component life which have 
plagued us for now five decades, but without the cost of even one degree of sophistication more than required for that goal. 

The discipline of dynamics shares the preeminence with that of structures as offering the greatest potential for advanced 
helicopter capability in this decade and the next. The assemblage at this symposium is the hope of that achievement. May you rise to 
the challenge and show its worth! 



282 



SESSION V 

APPLICATION OF DYNAMICS TECHNOLOGY TO HELICOPTER DESIGN 

Panel 1 : Prediction of Rotor and Control System Loads 



Peter J. Arcidiacono, Moderator 
William D. Anderson 
Richard L. Bennett 
Wayne Johnson 
Andrew Z. Lemnios 
Richard H. MacNeal 
Robert A. Ormiston 
Frank J. Tarzanin, Jr. 
Richard P. White. Jr. 



Panel Members 

Head of Rotor Systems Design and Development, Sikorsky Aircraft 

Research and Development Engineer, Lockheed California Company 

Assistant Group Engineer, Aeromechanics, Bell Helicopter Company 

Research Scientist, USAAMRDL, Ames Directorate 

Chief Research Engineer, Kaman Aerospace Corporation 

President, The MaeNeal-Schwendler Corporation 

Research Scientist, USAAMRDL, Ames Directorate 

Chief, Rotor Loads Unit, Boeing Vertol Company 

Executive Vice President and Director of Engineering, Rochester Applied Science 
Associates, Inc. 



283 



COMPARISON OF SEVERAL METHODS FOR PREDICTING LOADS ON A HYPOTHETICAL HELICOPTER ROTOR 

Robert A. Ormiston 
Research Scientist 

Ames Directorate 

U. S. Army Air Mobility R&D Laboratory 

Moffett Field, California 94035 

ABSTRACT 

Several state-of-the-art methods for predicting helicopter rotor loads were used to calculate rotor blade loads including airloads, 
bending moments, vibratory hub shears, and other parameters for a hypothetical helicopter rotor. Three different advance ratios were 
treated: n = 0.1 , 0.2, and 0.33. Comparisons of results from the various methods indicate significant differences for certain parameters 
and flight conditions. Trim parameters and flapwise bending moments show the smallest variations, while chordwise bending moments, 
torsional moments, and vibratory shears show moderate to large differences. Torsional moment variations were most sensitive to 
advance ratio. Analysis of the results indicates that the differences can be attributed to all three fundamental parts of the problem: 
numerical solution methods, structural dynamics, and aerodynamics. 

INTRODUCTION 

The prediction of rotor loads is one of the most difficult analytical problems in rotary wing technology since it involves a highly 
nonlinear aeroelastic response problem. Rotor loads, however, are basic to helicopter design because vibratory forces and moments 
from the rotor largely determine fatigue life, reliability, flight envelope limits, and ride comfort. Ultimately, rotor loads have a large 
impact on the cost of the vehicle. Much effort has been devoted to the development of sophisticated methods for calculating rotor 
loads, but these methods necessarily depend on empiricisms and approximations because the aerodynamic and structural phenomena 
involved are not completely understood. 

Needed progress in the development and refinement of these methods is often hindered for several reasons - the inherent 
difficulties of the problem, the specialization of different methods to treat different rotor types, the scarcity of reliable experimental 
data, and the difficulty in transferring experience gained by different investigators using different methods. As a result, it is difficult to 
accurately assess the state-of-the-art or to reach a consensus on the areas that require special attention. One proposal to partly 
overcome these difficulties is to specify a standard problem for calculating and comparing results of several loads prediction methods. 
This approach would focus attention on common, as well as individual, problem areas, permit sensitizing or "calibrating" different 
methods with respect to one another, and provide a new and broader basis for transferring experience. 

Several rotor loads specialists jointly agreed to undertake a project of this type for presentation at the AHS/NASA-Ames 
Specialist's Meeting on Rotorcraft Dynamics. This paper summarizes the main results. The project was made possible only by the 
enthusiastic cooperation of these specialists and the support of helicopter manufacturers and other organizations. The principal 
individual contributors were Wayne Johnson, USAAMRDL, Ames Directorate; Richard L. Bennett, Bell Helicopter Co.; Frank J. 
Tarzanin, Jr., Boeing Vertol Co.; James R. Neff, Hughes Helicopters; A. Z. Lemnios, Kaman Aerospace Corp.; John Gaidelis, Lockheed 
California Co.; Michael P. Scully, M.I.T.; J. J. Costes, O.N.E.R.A.; Peter J. Arcidiacono, Sikorsky Aircraft; and A. J. Landgrebe, United 
Aircraft Research Laboratories. 

Experimental data are not available for establishing the accuracy of any one loads prediction method; therefore, all 
interpretations and conclusions are based solely on relative comparisons of the results. 

STANDARD PROBLEM SPECIFICATION 

The standard problem was defined on the basis of inputs from all contributors. The basic philosophy was to emphasize 
aerodynamic phenomena by choosing a simple structural configuration. Most of the analytical difficulties are associated with 
aerodynamics, and interpretations of the results can be made simpler by removing unnecessary structural details. 



Presented at the AHS/NASA-Ames Specialist's Meeting on Rotorcraft Dynamics, Moffett Field, California, February 13-15, 1974. 

284 



A conventional articulated rotor was chosen, with three rectangular blades having 10° twist, an NACA 0012 airfoil section, and 
coincident flapping and lead-lag hinges with 4% offset. The blade is uniform in stiffness with coincident mass center, aerodynamic 
center, shear center, and feathering axis to minimize aeroelastic coupling effects. Complete details are given in the Appendix. A tip 
speed of 750 ft/sec (M = 0.672) and rotor lift coefficient C L /a = 0.0897 were chosen for three basic flight speeds: Case A at 
250 ft/sec 0* = 0.333) to emphasize retreating blade stall flutter, Case B at 150 ft/sec Qi = 0.20) for a typical unstalled flight 
condition, and Case C at 75 ft/sec (p. = 0.10) in the transition flight regime to emphasize vortex wake-induced blade loads. 

In addition to the three basic cases, several additional specialized cases were defined. These include selective combinations of rigid 
blade motion, linear aerodynamics, and uniform downwash in contrast to the most general case that includes elastic blade response, 
nonlinear aerodynamics, and nonuniform downwash. The particular combinations treated are listed in Table 1. Nonlinear 
aerodynamics is defined here to include the effects of unsteady stall, compressibility, and spanwise flow on the airfoil lift, drag, and 
moment coefficients. Linear aerodynamics is defined somewhat arbitrarily as inviscid, incompressible, steady-state, two-dimensional 
airfoil theory with a skin friction drag contribution (cj = 2hol, c^ = c^ , c m = 0). Nonuniform downwash is based on either a linear 

rigid wake geometry or a nonlinear distorted wake geometry determined by self induction. 

CONTRIBUTORS 

Results were contributed by the organizations listed in Table 2. A code is assigned to each one to identify results in the figures 
that follow. Where applicable, the name of the computer program is also included. These programs encompass a rather wide variation 
of sophistication and capability. Several are multi-use "global" programs that include rotor loads prediction as one of several 
capabilities. Others were developed primarily for predicting rotor loads. Three methods (noted by asterisks) are specialized programs, 
not intended for routine rotor loads predictions, but are included here to broaden the comparisons. 

Only the main results from the large body of contributed data are presented. The complete results include azimuthal variations, 
at several radial locations, of flapwise and chordwise bending moment, lift, drag, and pitching moment airloads, angle of attack, 
downwash, root torsional moments, and blade tip elastic torsion deflection; radial distributions of oscillatory bending moments and 
harmonics of the lift airloading; root vibratory shears; and trimmed rotor forces, power, flapping angles, control angles, and distorted 
wake geometry. The complete results are compiled in Ref. 1. 

MATHEMATICAL MODELS 

Pertinent features of the numerical solution methods, structural dynamics, airfoil aerodynamics, and downwash modeling of the 
various rotor loads prediction methods are summarized in Tables 3 and 4. This information does not necessarily describe the complete 
capabilities of each program; it refers only to program features applied to the present calculations. A more detailed discussion of 
various methods may be found in Refs. 2 to 6. Each method falls into one of four basic categories, depending on the form of the 
equations of motion (modal or finite element) and the method of solution (harmonic or numerical integration). 

Numerical Solution Methods 

The numerical integration solution method integrates the equations of motion in time from an arbitrary, assumed initial 
condition until a suitably converged solution is achieved. A variety of initial conditions, convergence criteria, convergence rates, 
integration methods, step sizes, etc., is used. The harmonic response solution methods generally calculate structural response using 
airload harmonics based on previously assumed blade motion and structural response. Iterations between the response and airloads 
continue until a converged solution is achieved. 

Structural Dynamics 

Modal equations are obtained from the partial differential equations of a rotating beam, and the generalized coordinates of the 
normal modes are the system degrees of freedom. Finite element equations are based on the force and moment equilibrium of a 
lumped mass or elemental segment of the blade and each element requires several degrees of freedom. In either the modal or the 
finite element representation, the coupling between degrees of freedom is important. This is difficult to summarize and the 
information in Table 3 is far from definitive. Substantial differences do exist, mainly in regard to coupling of the chordwise or 
torsional motion. Different methods are also used to calculate blade bending moments, although complete information is not included 
in Table 3. 

Aerodynamics 

Airfoil aerodynamics and induced downwash are determined by a variety of methods. Static airfoil properties are usually based 
on experimental data, using a table look-up procedure, with unsteady wake effects approximated by Theodorsen theory below stall. 
Unsteady aerodynamics in the stall regime, i.e., dynamic stall, is treated by several semiempirical methods (Refs. 7-1 1) in Table 4. The 

285 



effects of yawed flow on lift and drag are included for some methods. The rotor loads programs used for Cases A 1 , C 1 , and C2 include 
a variety of methods for calculating nonuniform induced downwash (Refs. 12-15) (Table 4). Most are based on a lifting line theory 
with a finite element (potential vortex filaments) representation of the trailed and shed vorticity. Induced velocities are calculated 
from the Biot-Savart law. 

The unsteady potential flow of a finite-bladed rotor in forward flight actually encompasses both unsteady airfoil aerodynamics 
and wake-induced downwash. However, the near and far shed wake contributions are often treated separately. The near shed wake can 
be approximated either by using Theodorsen theory for the airfoil aerodynamic coefficients or by considering an approximate near 
shed wake sheet segment as part of the wake system. The ONERA method uses a doublet acceleration potential theory that implies a 
continuous vorticity sheet, with both shed and trailed vorticity emanating from each blade. Other refinements of several of the 
methods include viscous tip vortex core representations, lifting surface theory for close blade/vortex passage, and wake distortion 
effects due to mutual self-induction of the vortex filaments. 

THE QUESTION OF VALIDITY 

An important question must be addressed when one attempts to compare numerous results from sophisticated computer 
programs that require many input parameters. To what extent are the results affected by errors in programming input data, errors in 
selecting program options, differences in calculating intermediate steps (such as natural modes and frequencies), or approximations 
(possibly unspecified) in the program that do not meet the standard problem specification? Of course, it is impossible to verify 
completely that all calculations have met the specified conditions. And this must always be kept in mind as a possible explanation for 
differences in the results. However, it is useful to examine the static airfoil data and the blade natural frequencies to check the 
consistency of certain basic ingredients in the calculations. 

Static Airfoil Data 

The static, two-dimensional, NACA0012 lift, drag, and pitching moment coefficients used for the different methods are 
compared in Fig. 1 for a Mach number of 0.5. The differences, especially at high angles of attack, indicate that there is not a universal 
consensus on static data of the NACA 0012 airfoil. Data for M = 0.2 and 0.9 (not shown) exhibit smaller and greater differences, 
respectively, than the M = 0.5 data. For Cases A1-A6 atn = 0.33, the maximum retreating blade angle of attack is about 13° to 14° at 
M = 0.23; therefore, the differences in static airfoil data are not considered to substantially affect the comparison of the results. 

Blade Natural Frequencies 

The blade mass and stiffness properties were chosen to provide blade frequencies typical of existing helicopter rotors. Predictions 
by the different methods are compared in Fig. 2 for the lower frequencies of the blade rotating in vacuo. For this condition, the pitch 
angle at 0.75R is zero, the blade static deflections are virtually zero, and coupling terms are very small. Furthermore, in many cases, 
coupled frequencies* were not calculated. Fully coupled (flapwise, chordwise, torsion) results are denoted by a single asterisk, partly 
coupled results (flapwise, chordwise) are denoted by double asterisks, and uncoupled results by triple asterisks. The results are fairly 
close for the fundamental elastic bending frequencies, but a more significant variation, about 10%, exists for the first elastic torsion 
frequency. The frequency differences for bending should not appreciably affect the comparisons of the results, but the torsion 
differences are more troublesome since stall flutter is sensitive to torsion frequency. 

COMPARISON OF RESULTS 

Only a small sample of the results is presented, but this is sufficient to illustrate and support the main conclusions of this study. 
Results are included for trim conditions, oscillatory bending moments, vibratory shears, and azimuthal variations of selected 
parameters. Trim conditions were specified in terms of the rotor lift, propulsive, and side forces. The required control angles and 
resulting blade flapping response and shaft power are also included. In several instances, a complete series of results is not available for 
comparison. This indicates that a calculation was not applicable or that specific output parameters were not available from a particular 
loads program. In certain instances, when results for a specific case were not available, another case was substituted to broaden the 
basis for comparisons. This was done only when departures from the specified case are presumed to have only a minor effect on the 
comparisons. 

Case A3 

The advance ratio of 0.333 and elastic blade motion, dynamic stall, and uniform downwash are included. Considering the 
extreme operating condition, the trim conditions (Fig. 3) are in reasonable agreement, although the specified zero side force was not 

*The coupling included for the natural frequencies is not necessarily the same as the coupling included for calculating forced 
response (see Table 3). 

286 



accurately achieved by all the programs. The radial distribution of oscillatory flapwise and chordwise bending moments is presented in 
Fig. 4. The flapwise moments compare reasonably well, considering the flight condition. 

Chordwise moments show larger differences. First, there is a substantial variation in magnitude along the blade and, second, there 
is a qualitative difference in behavior at the blade root where the moments either increase to a large finite value or vanish. In the first 
case, differences can result from at least three factors: (1) dynamic stall aerodynamics, (2) structural/inertial coupling effects 
(primarily flapwise/ chordwise structural coupling), and (3) the method of calculating bending moments (in the case of modal 
equations). The effects of dynamic stall are considered to be less important for chordwise moments than for flapwise moments (see 
Cases AS and Bl below); however, significant differences in structural coupling are present. For modal methods, bending moments can 
be calculated from internal structural moments of the blade deflection equation (by summing the modal moment distributions for 
each generalized coordinate) or from a spanwise integration of the applied inertial and aerodynamic loads acting on the blade. With a 
limited number of modes, the latter method is more accurate. 

The second difference in the chordwise bending moments is associated with the lead-lag damper. The LCC calculations using 
finite element equations, with and without the lead-lag damper, show that the damper increases the chordwise moments at the root 
compared with the zero moment boundary condition for a pin-end beam without damping. The BHC and SA methods use modal 
equations and exhibit zero bending moment at the blade root (lead-lag damper was not omitted). This could be a result of the way the 
bending moments are calculated. Integrated blade loadings, including the damper moment, would give a root moment, whereas a 
summation of modal moment distributions would not. This is because normal modes determined in vacuo cannot satisfy the pin-end 
boundary condition with a viscous damper (unless complex modes are used) and the moment must be zero at the hinge for each mode. 

Azimuthal variations in bending moments at r = 0.5 (Figs. 5 and 6) show general agreement of the basic waveforms with some 
variations, particularly chordwise. The 1/rev flapwise moment is mainly due to blade twist. The azimuthal variation of the lift 
airloading at r = 0.8 (Fig. 7) begins to show the influence of dynamic stall flutter for the retreating blade and therefore some 
significant differences are evident. 

The clearest evidence of dynamic stall occurs in the elastic tip torsion deflection (Fig. 8). Each method, with the possible 
exceptions of ARC and HH, clearly predicts stall flutter behavior, but the character and amplitude of the motion shows substantial 
differences, probably due largely to differences in the semiempirical unsteady aerodynamic stall modeling. The HH result could be 
interpreted as stall flutter and the low amplitude may be due to the relatively small static pitching moment coefficient at stall. The 
sensitivity of stall flutter behavior to details of the solution method is shown by the two BHC calculations for tip torsion deflection. 
Results for two different azimuthal integration step sizes are shown, A^ = 4° and 10°. The 10° azimuth increment reduces the peak 
torsion response by 5 .5° , or about half. An increment of 6° (not shown) reduced the torsion response by about 2° . Chordwise bending 
moments were only moderately influenced by integration step size, and flapwise bending moments only very slightly. 

The first five sine and cosine harmonics of the vertical and horizontal shears acting on the hub in the rotating coordinate system 
are shown in Fig. 9. The first harmonic components are nearly proportional to the rotor tip path plane tilt and reflect the trim 
propulsive and lateral forces, respectively. Since lateral trim variations are present (Fig. 3), the large variation in the first cosine 
harmonic of vertical shear is not surprising. Generally, the vibratory shears show relatively large variations between the different 
methods, but again this represents the most extreme flight condition. 

Case A5 

This case is identical to Case A3, except that the airfoil coefficients are based on linear theory, excluding stall, unsteady 
aerodynamics, and compressibility as defined above. These results help explain the results of Case A3 by removing the contributions of 
stall. Spanwise distributions of the oscillatory bending moments are shown in Fig. 10. The flapwise moments and their variations are 
both reduced compared to Case A3. The average of the chordwise moments was not substantially changed, but chordwise moment 
variations actually increased. Note that the nonstandard shaft angle for KAC could contribute to reduced chordwise Coriolis 
excitation. The lack of airfoil stall strongly suggests that the large differences in chordwise bending moments may be caused by 
differences in the structural/inertial coupling terms in the equations of motion and the lead-lag damper noted above for Case A3. The 
azimuthal variations of lift and drag airloading at r = 0.8 (Fig. 1 1) show only minor variations. The effects of dynamic stall on the lift 
observed in Case A3 are not present and the minor variations in the drag airload are not sufficient to account for the differences in the 
chordwise bending moment. 

The blade elastic tip torsion deflection (Fig. 12) is substantially reduced without dynamic stall. Not all the results, however, met 
the Case A5 specification of zero pitching moment coefficient (c m = 0). Results for SA and HH retained the linear quasisteady 
contribution of pitch rate damping; BV retained the apparent inertia aerodynamic term. BHC, KAC, and LCC used zero pitching 
moment coefficient. Notwithstanding these inconsistencies, the results illustrate possible evidence of anomalous behavior arising from 
the numerical solution method. BHC and KAC appear to have undamped torsion oscillations while HH shows a numerically 
unconverged solution; the specified value of zero pitch damping may be a contributing factor. Differences in the steady and 1/rev 
torsion imply again that differences exist in structural inertial/ coupling terms in the equations of motion. 

The above conclusions are also supported by the vibratory shears shown in Fig. 13. These shears are not greatly reduced from 
Case A3, indicating that stall is not the only contributor to vibratory shears at high advance ratio. A clear explanation for the 

287 



remaining differences is not available. Possibilities include the struetural/inertial coupling, differences in the numerical solution 
method, or differences in the lateral trim condition. Particularly disturbing are the substantial differences even for lower harmonics of 
the vibratory shears. 

CaseBl 

This case is identical to Case A3 except that the advance ratio is reduced to 0.2 where retreating blade stall becomes much less 
significant. Nonlinear aerodynamics, elastic blade motion, and uniform downwash are included. Only the results for trim and the 
oscillatory blade bending moments are shown (Figs. 14 and 15), and these results substantiate the findings from Cases A3 and''A5. For 
example, the chordwise moments show substantial differences that cannot be attributed to stall aerodynamics because the angle of 
attack of the retreating blade is generally well below stall. 

Cases CI and C2 

For these cases, the advance ratio is reduced to 0.1 and rotor loads are dominated by interactions between the blades and trailing 
tip vortices in the wake. The few results included emphasize the wake-induced velocities and blade lift airload. Cases CI and C2 both 
include nonlinear aerodynamics, elastic blade motion, and nonuniform downwash. Case C2 includes a rigid wake geometry for 
determining downwash velocities and Case CI includes a more general distorted wake. At the low advance ratio, blade stall is virtually 
nonexistent and elastic blade motion is not large; hence the main effects can be attributed to the wake. Several of the methods used 
for Cases CI and C2 did not include the specified airfoil aerodynamics and elastic blade modeling as noted in Tables 3 and 4, but these 
exceptions are not considered to be significant. 

The induced downwash for Case C2, with the rigid wake, is given in Fig. 16. Three methods (BV, MIT, UARL) use a finite 
element potential vortex system and direct application of the Biot-Savart law. The BHC method uses a semiempirically modified 
momentum theory and the OR method uses the doublet acceleration potential theory that implies a continuous sheet of vorticity 
behind each blade. Results from the four potential theory methods are similar, except the BV result is approximately 1/5 the 
magnitude of the others. This is due to an arbitrary reduction in the strength of the wake vorticity, which was found to give better 
correlation with blade loads measurements. All the results show the typical increase in downwash at the rear of the rotor and decrease 
at the front. The blade lift airload (Fig. 17) also shows reasonable agreement, especially with respect to lift impulses caused by close 
blade/vortex interactions. Particularly interesting is the qualitative agreement of the finite vortex method (UARL), the mixed finite 
vortex/sheet vorticity method (MIT), and the sheet vorticity method (OR). The BV result reflects the reduced wake circulation 
strength, but the same blade/vortex interaction's are discernable. 

Results for Case CI with the distorted wake are given in Figs. 18 and 19. The distorted wake is induced into a position close to 
the rotor disc plane and the results become more sensitive to the details of the prediction method. For example, the downwash 
variations for UARL are increased due to the closer proximity of the wake vortices, but high frequency variations were also introduced 
in part by the numerical solution method. The MIT lift airload impulses from the blade/vortex interactions would be considerably 
larger without the use of a lifting surface method that scales down the airloads. 

Summary Results 

A useful summary can be obtained by plotting results for Cases C2, Bl, and A3 as a function of advance ratio. Figure 20 shows 
oscillatory flapwise and chordwise bending, moments (r = 0.5) and blade root torsion moments. The results for CaseAS are also 
included to emphasize the effect of stall in comparison with Case A3 results. The differences in flapwise moment are not large and 
they increase only moderately with speed as retreating blade stall begins. The increased BHC moments at n — 0.1 are attributed to the 
semiempirical correction of downwash for the effect of blade tip vortices. The chordwise moments show large differences for all three 
speeds and also for linear aerodynamics, implying that stall is not the cause of the differences. Root torsion moments (equivalent to 
elastic tip torsion deflection and pitch link loads) are very sensitive to retreating blade stall and hence advance ratio. These results are 
not surprising in view of the number of different semiempirical dynamic stall models currently in use. Since torsion moments typically 
grow rapidly after a certain threshold speed is reached, small differences in this threshold speed could contribute to the large variations 
at n = 0.33. The straight lines joining the data points are not intended to suggest the actual functional variation, but only to indicate 
continuity between the data points of different methods. 

INTERPRETATIONS AND ASSESSMENTS 

It is not uncommon for rotary wing analysts and researchers to assume that rotor blade structural dynamics is well understood 
and that the main obstacle to accurate prediction of rotor loads is an incomplete understanding of rotor aerodynamics, particularly 
dynamic stall. This is certainly true to an extent, but it nevertheless reflects a tendency to focus on one aspect of the problem and 
overlook other areas in need of attention as well. The above results and comparisons show that limitations for rotor loads prediction 
exist in all three main areas: (1) numerical solution methods, (2) structural dynamics, and (3) aerodynamics. A more detailed 
interpretation and assessment of the above results in terms of these three main subjects is given below. 

288 



Numerical Solution Methods 

The results illustrate several examples where accuracy was compromised by numerical techniques and procedures used to solve 
the equations of motion. Considering the complexity of a large number of highly coupled, nonlinear equations, difficulties of this type 
are not surprising. Additional examples of numerical problems were found in subjective reactions expressed by contributors in the 
preliminary stages of calculating their solutions. In several cases, unexpected difficulties were encountered in achieving converged 
solutions, in reaching the desired rotor lift levels or in avoiding apparent blade motion instabilities. In several cases, these difficulties 
were resolved by adjustments to numerical details in the solution method or arbitrary deviations from the standard problem 
specifications. In most instances, problems of this type occurred for the high-speed flight condition. Little discussion of these 
problems appears in the literature, but they appear to be significant. 

Structural Dynamics 

One of the main results of the comparisons was the large differences in chordwise bending moments and vibratory shears even 
when nonlinear aerodynamic phenomena were absent or negligible. This clearly indicates that important differences exist in the 
structural dynamic modeling, probably due, in part, to different assumptions and approximations for structural/inertial coupling and, 
in part, to an inadequate fundamental knowledge of the mechanics of a flexible rotating beam. In the first place, there is insufficient 
information about the approximations used and this contributes to difficulties in drawing meaningful conclusions. It is evident that for 
conditions of minimum coupling, i.e., a beam rotating in vacuo at zero pitch angle, the lower natural bending frequencies show 
relatively good agreement. It also appears that chordwise bending moments are highly sensitive to flapwise/chordwise coupling. For 
example, the methods that calculated uncoupled chordwise moments, BV and KAC, generally showed the lowest results. It is difficult, 
however, to draw more substantive conclusions about other important structural/inertial coupling terms, such as bending/torsion 
coupling. Nevertheless, the results indicate that structural coupling effects are significant, that different approximations and 
assumptions are in common use, and that the validity of these approximations needs to be examined more closely. 

Also, fundamental questions can be raised concerning blade structural dynamics. Certainly, this subject is much better 
understood than nonlinear aerodynamic phenomena. However, the development of nonlinear structural equations for bending and 
torsion of slender beams is still in progress. The widely accepted work of Houbolt and Brooks (Ref. 16) provides only linear equations 
of motion. A recent derivation of nonlinear equations of motion by Dowell and Hodges (Ref. 17) points out some of the difficulties 
involved in more complete treatments. These involve assumptions in neglecting higher order terms, nonlinear strain displacement 
relations, warping of beam cross sections, etc. 

Different methods used to calculate blade bending moments also contributed to differences in the results. These moments must 
be carefully evaluated, particularly when using the modal equations of motion. Results of the modal methods did not show the 
expected increase in chordwise bending moments near the blade root because of the lead-lag damper. 

Aerodynamics 

Although probably not a substantial source of inaccuracy, the degree of variation in the NACA0012 airfoil coefficients was 
unexpected, and certainly should be remedied. The differences in the stall flutter torsion response of the rotor blade were not 
unexpected because the present semiempirical methods for calculating dynamic stall aerodynamics are quite different.** 
Improvements will depend on future research into very basic aerodynamic phenomena. Although it is impossible to interpret the 
present rotor loads predictions in terms of detail aerodynamic phenomena, it is well understood that three-dimensional aerodynamics, 
yawed flow or sweep effects on drag and stall, and blade vortex interactions need much more attention. The rotor wake significantly 
influences rotor loads at low to moderate speeds. Several different methods are currently in use and again many of the assumptions 
and approximations need to be examined more closely. 

CONCLUSIONS AND RECOMMENDATIONS 

The present results have shown that there is certainly much to be gained from direct comparisons of helicopter rotor loads 
prediction methods. A clearer perspective of the state-of-the-art is now available and this alone should promote further improvements 
and advances in (he state-of-the-art. The results also show that, for extreme operating conditions; the different rotor loads methods 
predict results with significant differences for some parameters. The following general conclusions summarize the results with respect 
to the relative quantitative accuracy of these methods: 

(1) Rotor forces, power, control positions, and blade flapping generally showed small to moderate differences. 



**W. Johnson (Ref. 18) presents comparisons of rotor loads predictions where only the dynamic stall methods are varied. The 
calculations are for the high-speed flight condition (Case A) of the rotor configuration developed here. The results obtained comple- 
ment the present comparisons. 

289 



(2) Flapwise bending moments showed small to moderate differences that increase with advance ratio. 

(3) Chordwise bending moments showed moderate to large differences for all advance ratios. This is attributed to structural 
dynamics. ; 

(4) Torsional response showed moderate to large differences, especially at high advance ratio. This is attributed to dynamic stall 
aerodynamics. !s 

(5) Blade root vibratory shears showed large to very large differences. Significant differences occurred even with linear 
aerodynamics and a satisfactory explanation is not available. 

(6) At m = 0.1, the blade lift and downwash calculated by different methods show qualitatively similar results, including 
blade/vortex interactions. Differences were largest for the distorted wake calculations. 

(7) It is concluded that differences in the results arise from all three basic areas: numerical solution methods, structural 
dynamics, and aerodynamics. 

Based on the discussions and conclusions above, the following recommendations are made for possible ways to improve rotor 
loads prediction methods: 

(1) Standard comparisons should be continued. Sensitivity to additional configuration parameters should be determined and 
comparisons for trend behavior should be made. 

(2) Additional investigation should be undertaken in greater detail to resolve questions raised by the present results. At the 
present time, it is not possible to sort out and assess all the assumptions and semiempirical factors in the problem by examining the 
resultant rotor loads. This applies to comparisons between different methods and to correlations with experimental data as well. 
Computer experiments should be used to study specific isolated aspects of the solution methods and structural dynamics. The 
accuracy of approximations and validity of specific assumptions should be determined. 

(3) Fundamental experimental research should be continued to investigate dynamic stall, blade/vortex interactions, and • 
three-dimensional flow effects. Special experiments should be designed to isolate and simulate specific aspects of the total rotor 
aerodynamic problem. This will eventually aid in the development and validation of analytical modeling for individual aspects of the 
total aerodynamic problem. 

(4) Ultimately [after some progress is made with items (2) and (3) above] , it will be useful to check rotor loads prediction 
methods with experimental data from wind-tunnel tests of a large-scale rotor. For maximum benefit, these tests should be performed 
on a rotor specially designed and fabricated to eliminate all nonessential complexities, and thus enhance the chances for meaningful 
correlations. Also, a new design could avoid some of the operating stress limitations of existing flightworthy rotor systems and permit 
testing at more extreme operating conditions. Sufficient instrumentation would be essential to measure all important parameters 
including steady and vibratory hub forces, blade airloads and stresses, and wake-induced velocities. 

REFERENCES 

1. Ormiston, R. A., "Compilation of Results of Several Loads Prediction Methods for a Hypothetical Helicopter Rotor," NASA 
report to be published. 

2. Anon., Specialists Meeting on Helicopter Rotor Loads Prediction Methods, Milan, Italy, March 1973, AGARD Conference 
Proceedings No. 122. 

3. Johnson, W., "The Effect of Dynamic Stall on the Response and Airloading of Helicopter Rotor Blades," Journal of the 
American Helicopter Society, Vol. 1 4, No. 2, April 1 969. 

4. SADSAM User's Manual, MSR 1 0, The MacNeal-Schwendler Corp., Los Angeles, Calif., Dec. 1 970. 

5. Costes, J. J., "Calcul des Forces Adrodynamiques Instationnaires sur les Pales d'un Rotor d'Hdlicoptere," La Recherche 
Ae'rospatiale, No. 1972-2 (March-April). (See also AGARD Report No. 595, April 1972.) 

6. Scully, M. D., "Computation of Helicopter Rotor Wake Geometry and Its Influence on Rotor Harmonic Airloads," M.I.T. 
Aeroelastic and Structures Research Laboratory Report (to be published). 

7. Ham, N. D. and Garelick, M. S., "Dynamic Stall Considerations in Helicopter Rotors," Journal of the American Helicopter 
Society, Vol. 13, No. 2, April 1968. 

290 



8. Carta, F. O., Casellini, L. M., Arcidiacono, P. J., and Elman, H. L., "Analytical Study of Helicopter Rotor Stall Flutter," Paper 
presented at the 26th Annual National Forum of the American Helicopter Society, Washington, D. C, 1970. 

9. Tarzanin, F. J., Jr., "Prediction of Control Loads Due to Blade Stall," Journal of the American Helicopter Society, Vol. 17, 
No. 2, April 1972. 

10. Harris, F. D., Tarzanin, F. J., Jr., and Fisher, R. K., Jr., "Rotor High Speed Performance, Theory vs. Test," Journal of the 
American Helicopter Society, Vol. 1 5, No. 3, July 1 970. 

11. Ericsson, L. E. and Reding, J. P., "Dynamic Stall of Helicopter Blades," Journal of the American Helicopter Society, Vol. 17, 
No. 1, Jan. 1972. 

12. Miller, R. H., "Rotor Blade Harmonic Airloadings," AIAA Journal, Vol. 2, No. 7, July 1964. 

13. Johnson, W., "A Lifting Surface Solution for Vortex Induced Airloads," AIAA Journal, Vol. 9, No. 4, April 197 1 . 

14. Landgrebe, A. J., "An Analytical Method for Predicting Rotor Wake Geometry," Journal of the American Helicopter Society, 
Vol. 14, No. 4, Oct. 1969. 

15. Davenport, F. J., "A Method for Computation of the Induced Velocity Field of a Rotor in Forward Flight, Suitable for 
Application to Tandem Rotor Configurations," Journal of the American Helicopter Society, Vol. 9, No. 3 , July 1 964. 

16. Houbolt, J. C. and Brooks, G. W., "Differential Equations of Motion for Combined Flapwise Bending, Chordwise Bending, and 
Torsion of Twisted Nonuniform Rotor Blades," NACA Rept. 1346, Oct. 1958. 

17. Dowell, E. H. and Hodges, D. H., "Nonlinear Equations of Motion for the Elastic Bending and Torsion of Twisted Nonuniform 
Rotor Blades," NASA Technical Note to be published. 

18. Johnson, W., "Comparison of Three Methods for Calculation of Helicopter Rotor Blade Loading and Stresses Due to Stall," 
NASA report to be published. 

APPENDIX 



Rotor Configuration 










Radius 




25.0 ft 


Blade mass/unit length 


0.25 slugs/ft 


Chord 




1.83 ft 


Polar mass moment of inertia/unit 
length 


0.03slug-ft 2 /ft 


Number of Blades 




3 


Flapwise mass moment of inertia/unit 




Solidity 




0.070 


length 


0.001 slug-ft 2 /ft 


Twist, Linear 




-10°(fromr=0to 1.0) 


Chordwise mass moment of inertial/unit 
length 


0.029 slug-ff 2 /ft 


Lock number (a = 5.73 rad) 




7.49 


Polar area moment of inertia 


0.003 ft 4 


Airfoil 




NACA 00 12 


Flapwise bending stiffness 


30X10* lb-in. 2 


Blade root cutout 




3.75 ft (r= 0.15) 


Chordwise bending stiffness 


400X10 6 lb-in. 2 


Flap hinge offset 




1.0 ft (r = 0.04) 


Torsional rigidity 


30X1 s lb-in. 2 


Lead-lag hinge offset 




1.0 ft (r = 0.04) 


Feathering axis location 


0.25c 


Control system kinematic pitch-flap 








and pitch-lag coupling 




0.0 


Elastic axis (shear center) 


0.25c 


Lead-lag damper (ideal viscous 


damping 




Center of mass 


0.25c 


about the lead-lag hinge) 




3,0X10* ft-lb-sec 


Tension axis 


0.25c 



291 



Operating Conditions 

Tip-Speed 

Speed of sound 

Density 

Shaft angle of attack 

Rotor lift coefficient, C-^Ja 

Lift 

Nomenclature 
c 

c d 

c l 
C(k) 

C L 

c m 
Cp 

M 

r 

a 

h 

M 
a 

4> 

«s 

= a - a, s cos $ - b, s sin \jj 

0o.7 5 = ^o _ -^is cos ^ - B is sin ^ 

Coordinate Systems and Sign Conventions 
Blade torsion deflection 

Rotor forces 

Lift and drag airload 
Flapwise/chordwise bending moment 



750 ft/sec 




Drag (parasite area) 




25 ft 2 




11 17 ft/sec 




Side force 




0.0 




2.378X10 -3 


slug/ft 3 




Case A 


CaseB 


CaseC 


0.0° 




Forward velocity, ft/sec 


250 


150 


75 


0.0897 




Advance ratio 


0.333 


0.20 


0.10 


16,500 lb 




Advancing tip Mach 
number 


0.895 


0.806 


0.739 



Vibratory hub shear 



blade chord 

airfoil section drag coefficient 

airfoil section lift coefficient 

Theodorsen unsteady aerodynamic function 

rotor lift coefficient 

airfoil section moment coefficient 

rotor power coefficient 

Mach number 

diraensionless blade radius, at center of rotor, 1 .0 at tip 

airfoil section angle of attack 

dimensionless induced downwash velocity, positive down 

advance ratio 

rotor solidity 

blade azimuth (when lead-lag angle equals zero) 

shaft angle of attack, pitch-up positive 

blade flapping angle with respect to shaft axis system, positive up 

blade pitch angle, shaft axis system, positive leading edge up 

positive leading edge up 

shaft axis system, positive thrust force up, propulsion force forward, side force 
toward right 

shaft axis coordinates, positive up and toward blade trailing edge 

airfoil principal axis coordinate system, positive for tension in blade lower surface 
and leading-edge, respectively 

forces acting on hub in rotating shaft axis coordinate system; vertical force positive 
up, horizontal force positive toward blade, trailing edge; harmonics of positive 
Fourier series 



292 



Table 1 .- Special Cases 



Case 


M 


Structure 


Aerodynamics 


Downwash 


Al 


' 


Elastic 


Dynamic stall 


Rigid wake 


A3 

AS 


• 0.33 


Elastic 
Elastic 


Dynamic stall 
Linear aerodynamics 


Uniform 
Uniform 


A6 


j 


Rigid 


Linear aerodynamics 


Uniform 


Bl 


0.2 


Elastic 


Dynamic stall 


Uniform 


CI 


1 


Elastic 


Dynamic stall 


Distorted wake 


C2 


Jo. 


Elastic 


Dynamic stall 


Rigid wake 



Table 2.- Contributors 



Contributor 


Code 


Program Name 


Ames Research Center* 


ARC 





Bell Helicopter Company 


BHC 


C81 


Boeing Vertol Company 


BV 


C60 


Hughes Helicopters 


HH 


SADSAM 


Kaman Aerospace Corporation 


KAC 


6F 


Lockheed California Company 


LCC 


3110 


Massachusetts Institute of Technology*t 


MIT 





Office National d'Etudes et de Recherehes 
Aerospatiales* 


OR 





Sikorsky Aircraft 


SA 


Normal Modes 


United Aircraft Research Laboratories 


UARL 


Normal Modes 



♦Specialized program. 

"(■Calculations performed under NASC Airframe Division Contract N00019-73-C-0378. 



293 



w 
n 
2 



CO 

U 

s 
< 

z 
> 


Structural and 
Inertial coupling 
of Modes or 
Finite Elements 


c 

s 

2 

c 
o 
2 ai 

8-i 

^ s 




Fully coupled normal modes. 
Bending moments deter- 
mined from modal moment 
distributions. 


Coupled flap/pitch. Uncou- 
pled chordwise. Approximate 
torsion moment due to 
chordwise bending. Structural 
effects of twist and pitch of 
principal axes neglected. 


13 

s, 

3 

o 
u 

> 
_ 

u. 


Fully coupled. Bending 
moments determined from 
uncoupled equations at 
zero pitch. 


Flapwise and chordwise 
bending fully coupled. Tor- 
sion structural response 
from flap and chord bending. 
Torsion dynamics neglected. 


Uncoupled. Approximate 
polynomial bending mode 
shape. 


C 
> 

3 

o 

3 

-a 

i* 

5 .S 
o tf 

5.1 


Uncoupled normal modes 
obtained for zero pitch and 
twist. Structural coupling 
terms included as forcing 
functions in modal 
equations. 


- 1 
t~ a o u. 








js a 






■o 








o 








"8 £© 1 © 






*** 

o 5 




CM 


00 


-J 
< 

D 








aa o 




































D 




"8 






















CC 




tj 


C 




= J3 «, 






s 3 






.s " 




CO 




75 

c a, J3 

il| 


T3 n 


c 
o 


« « .2 '5 c 
s: J 3 "H -2 
_ _ °- o £ 
5 3 n .s o 


o 

CM 


J3 


a, "o w 
I8IS 

2 2 ™ o 


CM 


5 S 
- '1 

T3 (0 


ll 

Ll j 

T3 TJ 


= 3 1 1 -8 






O o> 


h- 






















Sou 


CC <- 


r- 


CC CC cm t- <r- 






CC CC j- i- 




CC ,- 


(T CC 


CC DC CO CM i- 




1 ° 


















>■ 

■O in 


























"1 








§1 

o- .*_ 
UJ o 


I 




•o 
o 


£ uj 


'E a) 
LL UJ 


■o 
o 


= E 
C a> 

11 LU 


o 


1 
























> 










_a 
■d 2 


Si « 




■I •» 


•Be 

-a S * 






a S. S s 
■o -2- a fi 


o 
■o ai 








J2 

i 

CC 


1 1 2 1 1 

« 3 OJ 0> ♦= 

" "O « 'S £ 

s &* is 

is a a &£ 




"E « "5i 

fS 1 s 

1 c s = 

2 g S c 

£ S « 5 


E « £ 

Us 

« -D (0 
5 c Q. 

(- a. a 






e s a • 

3 r jits 

S « o -s 

™ 3 2 "O 
2 =; oi re 

2 e £ * £ 
£ 5 ^ > .= 
1- 5. a » i: 


3 "D 

S oj 

c .E « 
■2 2 £ 

2 o | 
2 o <3 




3 6 


2 c . « 
-^ 3 ■= o c 

a I S " a 


(0 




in 


o 


CO 


o 
ib 




m 


tn 


o 




cc £ 


<i:{" 
























. 










Ci 














to ^ 

o> js 
2 * 


-= 2 E 

| | | | | 


s 




"5 8> « 
> c 5 


CM 


i 

3 


s 


CO 
CO 


o 

*; c 


CO 


•a 

i! 




i fe 


< tS Q < X 






.» cc :*: 




Z 






x u. a 




S LU 


c 
o 


» d ? = J 






— 




















.B z S o = 






o ■* 






















o 






O 






o 


LO 


CO 


10 


CC 


at 
£ 


&1 £ = ^ 
« <3 o u. X 






*" < 




m 














to 






















UJ 

o 

o 

CC 

a. 

z 
o 

H 




s s s 










r_ 














» 'i 11 

CC Q. < E 










































_ 


§ , 






















S • . .s 

R s " 1 "• 

2 s g . ■= 

i£*£3 


1^ J= ■= 




O 

09 

o 

Z 

< 

CO 
O 

£ 

UJ 

z 




- 2 3 « 
S ait E 

I-g | S 

o S Etc 


JO 

§ a 




S = 

g E 


CD 
O 
(0 
« C 

I' 8 

— u> 


o 
o 

CO 
CO u 


31 

o 

CO ». 
CO S 

co t _ 


o _ 
to c 

S3 


S IS 

S S |.s 

— CM — 5 


" 5 ■* 
o — 3 
J.S * = 




£ c 


E „ 
o £ 

* > 




OJ 

§ -S s 


c 

■ll 

6 6 




1= 
o 

E "f 

is 

2 » 
ra N 

E S 


c 

o « 
■a o E 
S c H 


Is 


tjl 


?e2 






♦= c -s cc £ 
■= o J2 ,_ o 
£ o co o o 


I! 

a s 
cc a 




| » o « 
*i 8 w S 

r « s o 
•b s a o 
1 5 S i 


-a a. 

3-1 1 

11 1 
cc ^ a 


o 

1! 
: X 

■)< 
>5 


5|| 
« g — 
uj £ | 


2 S 
cc o 


a 

!C C 

*- .2 

3 2 


Hi 

Ol C rf 

£ 8z 

























































S . § 
g> ra 5 o 

« '= B O « 

£ 2 -o oc S 






SI 

c 


01 


■= « ° 2 

> O *- n 

= e y ^ S 

5 £ 1. *■ c 
o x 8 B o 


CM O 

£ +i *~ 

+1 O *: 

2 2- • 


5> a 

(0 o 

" c c 


o 

S u 
a a 

V ^S 

•^5 




"3^5 8 5 i 

_ „ N ,P i2 C 

So y .£< Z 






Q X S i_ O 
O U DQ O U. 


z 




z 


o 

Z 


O g §! « 2 


x 22 
haw 


ZEES 


ll 


o 
Z 


S<-2< -i ct <a s 
<3 < < < <J £ 




•£ S£ 


o v 




_ *= 




_ c 
ro O 


at 






O (j, 


— c 






"2 w .2 ■*= 
2 E ^ © 
f »3s 


M 

E a 




X ID 


18 

E a 


c w 


1*8 

£ S O 


C 2 

H 


Ol is. 

E S 1 


E a 


1 1 






S W UJ o 


X CC 




1 1 


X cc 


II 


X CC 


z £ 


n ai 
X CC 


is 






















< 
cc 


CC 
< 
























UI 








■g 


cc 




o 

X 


> 


X 


< 


a 


t- 


z 










< 




DQ 


CO 




^ 


_j 


s 


o 


CO 



294 



8 
S 

II 

II 



61 

> D 



3 S 5 O » 

H u. 5 > uj 



O 3 P 



CO ; 



3 2 o a 
Z h- > «? 



li? 1 

I I 1 S E 
= f SB 






ifi 

"Sfi 

1 1 



So: £ 

2- © ss v *~ 
H ©' < "« O* 






•g s 






.2 E 

•|"6 
» -S 

i* 

<5°to 

J v 
- s| 

ill 

|V I 

til 

! - « 



O <i» "5 " B ¥■ 

=* 5 m 0) .£ 

Si §1 s 






1 » ™ :s 



.a ° s » s 3 S 

holSi id 



t 'n -S "P - 






s.& 



6 § 



to £ 
.a 5 



t S 

> •= 

o § 

S a 

I- « 



os — 

« '5 w S S 

s a- V * 
2 ' s & - S 

S3jV» 

5 o S ° 1" 



ISli 



n % a 

. fiM 

£ £ 2 
Sis 

IeS 
£ S g 

ia s E 

si; 

.i§ I 

ISl 
111 



If 



t 



> 

I Ji 

i,t| 

III 

_j -fi 'C 



1 

ill 
it! 
ill 



I r 3 
5^1 



III 

3^1 



I iilSi 

3 * £ 5 € a .1 



111 

ill 



'£ 8 I 
3 *£ 



1 
11 



D ITS 



■g .3 .3. 



1 -§l 






z 

>- 
a 
o 



X B 



■» 
Sg 



It 

If 

CO -Q 






9 






2 

s 



I ? 1 ■i 

i s tc ec 



I- K 



li 



s"»i f 

n » 5 1 s 

s's 
t3 






■■» 



■ in 



III 



o 5 

- » la : 

3 S S a a £ 



s 



6 i 




ft - 

1 ia-8 

s * = S 



•si 5 




295 



C, .8 




6.4 r 



- r-^ZJz, 



-.04 
-.08 

c m-. 12 

-. 16 
-.20 
-.24 
-.28 



r 




s\ 






-. 




\HH 


BV "- SA V x 

KAC^. \ 


- 


BHC%, \ 
\ 


i i i 


i i i 



10 15 20 25 30 

a, cleg 



Fig. 1- NACA 0012 airfoil section lift, 
drag, and pitching moment coeffi- 
cients at M = 0.5 used for different 
methods. 




"ARC*** BHC* It* HH* KAC*** LCC** SA*** HIT* 
* COUPLED 
** PARTLY COUPLED 
*** UNCOUPLED 



Fig. 2.- Rotor blade rotating natural frequencies in vacuo, 

#0.7 5 = u • 



FORCES, I0 3 lb; POWER 

20 r 

THRUST 



CONTROLS, deg 



flapping, deg 



10 



Cp/o-XIO 3 



PROPULSIVE 
FORCE 




ARC*BHC* BV HH LCC SA 




ARC BHC* BV HH LCC SA 




I I I I I 

ARC* BHC BV HH LCC SA 



Fig. 3.— Trim parameters, Case A3. 



296 



CHORDWISE MOMENTS, 1/2 PTP, 10 3 in.- lb 



FLAPWISE MOMENTS, 1/2 PTP, I0 3 in.-lb 





\ \ 



.8 .9 1.0 



Fig. 4,— Oscillatory blade bending moments, Case A3. 



20 

15- 

10 

5 



-5 

-10 

-15 

-20 



15 
10(- 

5 



-5 

-10 



/ \ 
/ \ 

/ \ BHC(AI) 

\ 



i — i A 1- 



\.r 



BV 



A 

1 , V,\ , !\, 

I \ i \ 

hi \ / v, 

W \ / 

L W 



• 


/'\ 




/ \ HH 




\ 


„, 


/ 


' 


\ / "■v--^.'' 




\J 








40 

30 

20 

10 



-10 

-20 

30 

20 

10 



-10 

-20, 



30 

20 

10 



-10 

-20 



"/ v 

■ 1 
1 

1 , 


\ BHC(AI) 


1 

l 


1 i ■ 



BV 


i 


\ 







' \ 


HH 








i \ 






.^ 


/ 


'> 




\ r 








v' 


» / 













/ \ 



,v,/- >-^-\ 



■a A » / 



* — ^ — 1— ' 



W 



90 180 270 360 90 180 270 360 

t.deg f.deg 



Fig. 5.— Flapwise bending moment (in.-lb) vs. azimuth; 
r= 0.5, Case A3. 



Fig. 6 — Chordwise bending moment (in.-lb) vs. azimuth; 
r= 0.5, Case A3. 



297 



60 r 

40 

20 





ARC(AI) 



\ /■ 



I20p 

100 

80 

60 

40 

20 



-20 

-40 

-60 

80 

60 

40 

20 



BV 



I I 



\ / * 



" 1 



\ 

\ 

\ f 



LCC 



~N,x ' 



- 




HH 




/\ 














,. 


■ 




y" — "*"--. 


/ 


■ 








■ 









/ ' 



SA 



\ 



\ / v \ i ' 



90 180 270 360 90 180 270 360 

if-, deg iji, deg 



BHC(AI) 




90 180 270 360 90 180 270 360 



Fig. 7- Lift airload (lb/in.) vs. azimuth; r = 0.8, Case A3. 



Fig. 8.— Hade tip elastic torsion (deg) vs. azimuth, Case A3. 



HUB VERTICAL SHEARS, lb 



1200 
S00 
100 



-1200 
-1600 
-2000 




300 
200 



12345 12 3 45 

HARMONIC 



-100 
-200 
-300 
-400 



HUB HORIZONTAL SHEARS, lb 



BHC *CASE A I 




2345 12345 

HARMONIC 



Fig. 9.- Harmonics of hub vibratory shears, Case A3. 



298 



FLAPWISE MOMENTS, 1/2 PTP, I0 3 in.- lb 




.1 .2 .3 .4 .5 .6 .7 .8 .9 1.0 



CHORDWISE MOMENTS, 1/2 PTP, I0 3 in.-lb 



as 












30 


,'SA 


\ 




\ / 


/" \\ 


25 


1 


'HH ~\ 


20 


- \ 

/ 
/ 
/ 


— BV -^ --v 




X\x 


15 


LCC (NO LAG X \\ 
.^ ~. DAMPERS x \ \ 




Z^ BHC~^</\ \\\ 


10 


' 'Z^- - 




5 


7/ 

// 


KAC ^ O^ X \^,\ 




JS 


<a s = -6.42°) ^.. \\\ 




11111 T *->SJ 



J .2 .3 .4 .5 .6 .7 .8 .9 1.0 



I5r 
10 
5 




BV 






0R(A6) 

/ 
/ 



LCC 



n^ y, 



SA 



90 180 270 360 90 180 270 360 



* ,deg 



4, deg 



Fig. 1 1(b).— Drag airload (lb/in.) vs. azimuth; r = 0.8, Case A5. 



KAC 



H 1 1 1 



BHC 




Fig. 10.— Oscillatory blade bending moments, Case A5. 



80 
60 
40 
20 


60 

40 

20 





0R(A6) 



y \ 



BV 



KAC 



■— x HH ^ 

\ / 





90 180 270 360 90 180 270 360 

+, deg f, deg 



Fig, 12.- Hade tip elastic torsion (deg) vs. azimuth, Case A5. 



60 
40 
20 



\ 



LCC 



SA 



\ /~^~s 



90 180 270 360 90 180 270 360 

i)<, deg ^, deg 



Fig. 1 1(a).— Lift airload (lb/in.) vs. azimuth; r = 0.8, Case A5. 



299 



HUB VERTICAL SHEARS, lb 



-1200 



-KAC 
~L.CC 
-SA 



Wr %J 



NOTE: 1/5 SCALE 
FOR 11=1 






12345 12345 

HARMONIC 

HUB HORIZONTAL SHEARS, lb 



FLAPWISE MOMENTS, 1/2 PTP, I0 3 in. -lb 





LCC 
SA 



NOTE: 1/5 SCALE 
FOR n-l 



h 



2345 12345 

HARMONIC 



Fig, 13.— Harmonics of hub vibratory shears, Case A5. 



FORCES, I0 3 lb; POWER CONTROLS, deg 

10 r 



flapping, deg 





J — 1 — I I I — I 
ARCBHCBVHHKACLCCSA ARCBHCBVHHKACLCCSA 



oKfe 



■ i i i i i 



CHOROWISE MOMENTS, 1/2 PTP, I0 3 in.- lb 




Fig. 15.- Oscillatory blade bending moments, Case Bl. 



ARCBHCBVHHKACLCCSA 



Fig. 14.- Trim parameters, Case B 1 . 



300 




90 180 270 360 
f, cleg 



90 180 270 360 
V, deg 




MIT 



UARL 



90 180 270 360 90 180 270 360 

<p, deg +, deg 



Fig. 16.— Induced downwash, Xj , vs. azimuth; 
r = 0.95, Case C2, rigid wake. 



Fig. 17.— Lift airload (lb/in.) vs. azimuth; 
r = 0.95, Case C2, rigid wake. 




90 180 270 
f, deg 



90 180 270 
X deg 



360 



80 


r 
















70 


■ 


MIT 








i 

n 




UARL 


60 


,*» 










II i"> 






50 






A 




' ,1, 


! i i 
; i i 


I 




40 
30 


\ i 


'"'--'v'' 




/ 


/I 


1 t ' 
1 j j 

I l ' 


i 
1 

\ / 
V 


-''' 1l 


20 


■ v\j 








■ i 


! I 
! l 






10 










. ! 


1/ 







-,0 ( 












, !l 








'J ' 


) 90 


180 


270 


360 ( 


3 


90 




180 270 360 






<l>, deg 










+ 


deg 



Fig. 1 8.— Induced downwash, Xj , vs. azimuth; 
r = 0.95, Case CI, distorted wake. 



Fig. 19.— Lift airload (lb/in.) vs. azimuth; 
r = 0.95, Case CI , distorted wake. 



301 



OSCILLATORY FLAPW1SE MOMENT 
1/2 ptp, I0 3 in -!b, r = 0.5 



HH N 

LCCs 

BV- 

-A SA-CH I 

KACW 

j BHCj 

CASE** CASE 




** BHC, Al 



.2 .3 .33 

ADVANCE RATIO, fj. 



.33 



OSCILLATORY CHORDWISE MOMENT 
1/2 PTP, I0 3 in - lb, r = 0.5 

SA^o 




& **BHC, Al ©** ® 



.2 .3 .33 

ADVANCE RATIO, /i 



33 



OSCILLATORY ROOT TORSIONAL MOMENT 
1/2 PTP, I0 3 in.-lb 




.2 .3 .33 .33 

ADVANCE RATIO, (J. 



Fig. 20.- Summary of blade oscillatory moments vs. advance 
ratio. 



302 



DISCUSSION 

Panel 1 ■:■ Prediction of Rotor and Control System Loads 

Moderator Arcidiacono: First of all, I would like to extend my personal congratulations to Bob Ormiston for the very fine job that he 
did in coordinating the efforts that went into this paper. To my knowledge, this is the first serious attempt at a comparative study of 
the various rotor aeroelastic and loads analyses available to the industry. We all know that any analysis at any point in time is the 
product of a set of assumptions, and I am sure that most of us involved in developing these analyses have wondered at one time or 
another just how we were doing relative to the other fellow, or whether his analysis would correlate better with our particular test 
experience. I'm equally sure that in trying to establish our relative positions, we have been frustrated — I know I have been at times — 
in trying to assess the available literature with its inevitable limited description of assumptions forming the basis for any particular 
analysis. Or, perhaps — I have been frustrated by the fact that, and I have been guilty of this myself - that the list of assumptions 
referred to in describing an analysis just refers the reader to a list of references that have to be tracked down. I'm also sure that you 
have probably heard the phrases that the analysis was fully coupled and that variable inflow is included, and that unsteady aerody- 
namics were included, without really being sure of or having the time to dig into the details that were necessary to understand the 
procedures being employed. I think that the mere existence of these frustrations means that technology transfer has, to a certain 
degree, been inhibited and this in itself, to my mind, is a perfectly adequate justification for Bob Ormiston 's paper, because now 
without any words or fol de rol, these are the results, and one can certainly at least calibrate his own analysis relative to other analyses. 
We have seen the results of the paper now and comments are certainly in order. We have on our panel people who, at least recently, 
have been responsible for translating F ■= ma into a useful rotor analysis. These have been introduced by Troy, and I have asked that 
each panel member speak for 5 minutes, and if time permits, address generally the following subjects: Correlation of their analysis 
with tests, identification of any clearly unusual trends noted in their own particular results, and perhaps make suggestions for how 
further checking of analyses of these types could be done. After the prepared comments, we will turn the floor open for discussion 
and I'm sure there will be a lot of discussion. Now, we are going to take the panel members in alphabetical order, and Bill Anderson 
will talk first. 

Prepared Comments 

William D. Anderson 

Research-Development Engineer 

Lockheed California Co. 

First of all, I'd like to again extend appreciation to Bob for the fine job he did here. I'd also like to extend recognition' to John 
Gaidelis back home for his work as a key person in developing the Lockheed rotor loads program, and also really in performing the 
analyses that were reported here. The Lockheed rotor loads program used in this particular study is a finite element analysis which is 
capable of up to about 75 spanwise stations. We generally use on the order of 40 or 50. In this particular analysis, as was indicated, we 
used 23 stations due to the uniformity of the blade characteristics. The analysis is fundamentally a steady maneuver and steady level 
flight type of analysis and it is incapable of doing a transient maneuver, but we feel that transient maneuvers are equally critical in the 
design of rotary wing vehicles. We do have another program which most of you are familiar with in name anyway, and that is Rexor, 
which is much more expensive to run and was not applied because this study was primarily associated with steady flight conditions. 
The Rexor program is a modal analysis; it does represent the total coupled rotor and airframe. I would also like to say there is one 
thing that it does include which was brought up earlier in one of the discussions. It does include the reorientation of the structural 
principal axes which means that it is fully capable of including all the periodic (structural) effects as well. Both programs include stall 
aerodynamics basically as Tarzanin and Harris have developed. The Rexor program is solved by a numerical integration method and 
the program that was used here is basically a harmonic response type of analysis which converges in 3 to 4 rotor revolutions. At this 
time I would like to show some correlation that we have had on some AH-56 data using the program that was actually used in this 
study. Could I have Slide 2, please. 

Slide 2 shows chordwise bending and flap bending at approximately 60% rotor span, and chord bending and flap bending at 
approximately 6% of rotor span versus equivalent air speed with test data out to 210 knots, which is a little over 220 knots true air 
speed. The dots and the solid lines represent the cyclic loads. The dots are test data; the solid lines are analysis. The open circles and 
dashed lines represent steady loads, again test and analysis. You can see fairly good comparison in both the cyclic and the steady loads 
as a function of forward speed. There is some conservatism in the predicted cyclic loads as we go out in speed for the flapping 
moments at the 6% span. 

Slide 3 shows blade torsion moments at approximately 45% span, as a function of speed, again the cyclic and steady moments are 
shown with fairly good correlation. Shown also are the steady and cyclic pitch link loads as a function of speed, and the steady 

collective servo force which is a control load measured in a stationary coordinate system. This of course is sensitive to c m of the 

x o 
blade, blade tabbing, etc. Slide 3 shows finally the cyclic collective control loads versus speed. I think that these results show fairly 
good correlation of the program with Lockheed experience. 



303 



OVERALL CYCLIC LOADS, e • .TEST: CL840/AMCS 

ANALYSIS 

STEADY LOADS S ©TEST: CL840/AMCS 
ANALYSIS 



, ORAG 
i 100 

: 50 
i 



■ 50 

100 
' FWD 



i 



co re 



2 z 



10 

10 

™& 

200 

100 





I ft 



100 
FWD 



5 2 



50 



50 

DN 



-,— -Fi • 1=)— — - =&?-=. 




-©— X- -o_ 



120 140 160 180 

EQUIVALENT AIRSPEED- KNOTS 



OVERALL CYCLIC LOADS • • TEST: CL840/AMCS 

— — ANALYSIS 

STEADY LOADS O O TEST: CL840/AMCS 

___ ANALYSIS 



7 L.E. UP 


















O 20 

-<n « 

m • — 1 

S5 z'° 

Sg"o o 
5" T 

" 10 
I.E. DOWN 




















, 


jg 




*« 




__S_JL_i— 














raartarsss- 


■■ ~ " "*"' 


















TEN. 


2.0 

9 1.5 

O ) 

i 3i.o 














• 


• 














(l^^^ 


z § 
u 

£ o 

0.5 




















• 


© 


J?_ 


© 


® 


m 


















COMP. 


















TEN. 


3 




























—^■^' 


COLUC 
SERVO F 

-1000 
o 


'L__» !! — : J 




* . 





""s" 




r**"?" — ? t ? T 




COMP. 
















• 



120 140 160 18 

EQUIVALENT AIRSPEED - KNOTS 



Slide 2. 



Slide 3. 



I would like to say that, relative to the results of the prepared paper, one question that comes to mind is how can we have so 
many different definitions of the basic NACA 0012 airfoil that has been around for several years? I would also say that the treatment 
of the lag damper seems to require a little more scrutiny as evidenced by the root bending moments for chordwise bending. I'm not 
too familiar with how some of the other programs are solved, particularly the modal ones, and I would like to point out that in a 
modal solution you have to be extremely careful in how the loads are integrated. If one uses the second derivative of the deflection to 
get internal loads it leads to fairly large discrepancies. And in a modal solution it does take a spanwise integration of each load element 
times a moment arm to come up with fairly, realistic flap wise and chordwise moment distributions. 

I would also like to add that with regard to a fallout of rotor loads, vibration that is, it would seem that until a more consistent 
type of rotor loads prediction method is developed that the prediction of vibration will leave a little bit to be desired, I believe. And 
finally, I feel that this effort supports the need for the development of a considerably more comprehensive and universal rotorcraft 
simulation model which is capable of predicting rotor loads and I think it should even be comprehensive enough to be used in the 
prediction of airframe vibration. 

Peter J. Arcidiacono 

Head of Rotor Systems Design and Development 

Sikorsky Aircraft 

I would like to show you some of our correlation, but before getting into that, Slide 4 just shows the relative position of the 
Sikorsky results, compared to the other analyses on the flapwise, chordwise, and the torsional responses. The Sikorsky results are the 
solid symbols and the bars indicate the range of the maximum flapwise moment predicted for the B and A cases. The Sikorsky 
flapwise results tend to be in the mid to low range on flapwise, in the high range on chordwise, and in the mid to low range on the 
torsion. 

In looking at those results I don't think I was surprised by them, generally speaking, except for the result on chordwise, and the 
reason for that was that we were on the high end and our experience has been that we have tended to underestimate chordwise 



304 



RELATIVE POSITION OF SA ANALYTICAL RESULTS 



MAX. 

VIBRATORY 

FLAPWISEOQOOin-tbl 



MAX. 

VIBRATORY 

CHORDWISE (1000 in-lb) 




VIBRATORY 
TORSION DEFLECTION |deg) 


® 




- 


@ 


- 


• 


r 


r® • @ 





ANALYSIS YIELDS REASONABLE TIME HISTORY CORRELATION BUT 
UNDERPREDICTS AMPLITUDE OF CHORDWISE STRESSES 

CH-S3A 

137 KTS 50% RADIUS 





s/ 




1 




f 


*r\ 






// 


/ /* 






// 
// 
/ / 







- / / 


L. / / 


Xte*^^" 1 " 


*^\ 


M " 




/ / 




'■• / 










• / 












■4000 

-aooo 






\w/test 











120 AZIMUTH 240 



fr 






50% RADIUS 




3000 


Tssy^© / 

y ^ANALYSIS 


" 



ADVANCE RATIO 



VELOCITY KTS 



Slide 4. 



Slides. 



stresses. I have illustrated that point here on Slide 5 where, for the CH-53A helicopter, we have the time history of chordwise stress at 
50% radius. Down here we have the peak vibratory stress as a function of radius, test and analysis, and finally over here we have the 
vibratory stress as a function of velocity, test and analysis. As you can see, we have been under-predicting chordwise stresses. 



Slide 6 is flapwise stresses, and these have b«en pretty good 
as far as our experience is concerned, particularly when variable 
inflow and unsteady aerodynamics have been in&aded, and this 
shows the peak-to-peak response versus radius and versus veloc- 
ity. The time history correlations — you saw some of them pre- 
sented by Ray Carlson the other day — are equally good. 



CORRELATION OF FLAPWISE STRESSES HAS BEEN GOOD 
WHEN VARIABLE INFLOW AND UNSTEADY AERO INCLUDED 



In Slide 4 we were on the low end as far as torsion response 
was concerned and that tends to be consistent with our experi- 
ence in correlating with our test information. Slide 7 shows the 
vibratory control load as a function of speed, in this case for the 
CH-53A, for two different Cj/a's, test and analysis. The slide 
also shows the correlation for two different aircraft, differing 
primarily in the amount of blade twist. I think this illustrates the 
point that most of the comparisons that you have seen in Bob's 
paper are comparisons on an absolute basis, and as we all know, 
an analysis can still be extremely useful if it can predict qualita- 
tive trends, and this we feel is the case here. 

Slide 8 summarizes my observations and recommendations. 
I think the exercise was useful and interesting. I was surprised at 





- 


i 


i i 

TEST / /* 
/ / 


- 


4000 






J / 
0/ 
/ 


Vb - 

n 
\\ 

v- 






r 


^/analysis 






' i 


I I 


< 



, ANALYSIS 



my 



/o 



VELOCITY. KTS 



Slide 6. 



ANALYSIS IS LOW IN AN ABSOLUTE SENSE BUT PREDICTS CORRECT 
TRENDS OF CONTROL LOADS 




-J .'...-. L. 



- 


I I I 

CH54B / C"- 53 * 
I / / 


- 


. 


_ 




/ i A 






/ '// 






/ /// 






As^s 






i i i 





200 40 
VELOCITY KTS 



Slide 7. 



OBSERVATIONS AND RECOMMENDATIONS 

• THE EXCERCISE WAS INTERESTING AND USEFUL 

• BASIC AERODYNAMICS VARY WIDELY 

• IF TRUE, ESTABLISH STANDARD TEST FACILITY 

• RESULTS DIFFER EVEN WHEN COMMON AERODYNAMICS ARE USED 

< RUN STD CASES FOR ZERO TWIST IN HOVER AND FWD FLIGHT 

• VARIABLE INFLOW DISTRIBUTIONS DIFFER GREATLY 

• RUN STD CASE WITH PRESCRIBED WAKE GEOMETRY AND CIRCULATIONS 

• UNSTEADY AERODYNAMICS LEAST SOUND PART OF AERO TECHNOLOGY 

• CORRELATE METHODS WITH FULL SCALE, 2-D, FLEXURED. OSCILLATING 
AIRFOIL TEST RESULTS 

• AGREEMENT OF ANALYSES ON ABSOLUTE BASIS NOT GOOD. AGREEMENT IN 
PREDICTING TRENDS UNDEFINED 

• RUN STD CASES WITH DESIGN PARAMETER CHANGES 

• PROVIDE SYSTEMATIC MODEL AND FULL SCALE ROTOR DATA BANK 

Slide 8. 



305 



the 0012 airfoil steady-state data variations, and if that's really true, even after the passage of 10 or 15 years, I think it is ample 
justification to establish some sort of a standard test facility. However, even when the aerodynamics were pretty much the same, the 
results differed and I think this seems to suggest the need for running cases with zero twist in hover and in forward flight. I was also 
somewhat surprised about the variable inflow part of the problem. I had thought there was more agreement on basic technology 
for variable inflow than perhaps there is. Certainly Cornell Aero Labs showed over 10 years ago that the general trends of inflow could 
be predicted, and yet even at this late date people either aren't availing themselves of that technology or they don't believe it. 
Assuming that technology is in hand, I would think some sort of standard case using prescribed wake geometry and prescribed 
circulations would help isolate questions in that area. I think unsteady aerodynamics is the least sound part of the aerodynamic 
technology and I agree with Bob Ormiston on doing such tests as full-scale oscillating flexured airfoils, the airfoils being well 
instrumented. And finally the point I made a moment ago: on an absolute basis the agreement of the analyses in certain areas is not 
terrific. I think we ought to define how well these analyses predict qualitative trends and I would suggest several standard cases in 
which systematic parameter variations are made. 

Richard L. Bennett 

Assistant Group Engineer, Aeromechanics 

Bell Helicopter Company 

C-8 1 is a global program that we use at Bell Helicopter that was conceived and brought about to be a useful design tool. I would 
like to point out that this study that we have just gone through only exercised about 20% of the total program. This study didn't say 
anything about the maneuvering flight capabilities, the stability routines, or the second rotor analysis. The use of C-8 1 is widespread, 
and inhouse it is run on a completely open shop basis, so anyone who wants to can run the program. Consequently, I don't know all of 
the uses. I do know the uses when they have some difficulty, so one week I made a list of the cases that were brought in for diagnosis. 
There was the Cobra at 1 .9 g's; the Model 301 stability study, there were some questions on the fifth scale articulated rotor in a wind 
tunnel, another question concerning a C-8 1 run on an elevator sizing study on the Huey, and finally there were some cases being run 
for the "Drees Special." At Bell Helicopter the "Drees Special" is the windmill. 

I think in this area there is a real need for meaningful correlation studies, and I think that there are four necessary requirements 
to accomplish this task. The first is that we must have an accurate mathematical model, sound in theory and well programmed. 
Second, there must be an accurate source of input data that is readily available. Many of these programs, C-81 included, sometimes call 
for input data that are just a wee bit fuzzy. Third, there has to be a source of test data that can be certified. And fourth, there has to 
be compatibility of the above three items. 

One of the questions that comes up is, why the proliferation of so many different rotor analyses? I think part of it has come 
about because someone writes 1/3. They go out to the flight test, they get their results, and they try to compare the measured and the 
calculated, and naturally they don't compare. So what is the obvious conclusion? Let's start over again and write lj3 in a different 
coordinate system. Maybe that will help. 

Right now at Bell Helicopter we are in the middle of a correlation study using the results obtained from a 1/5 scale 4-bladed 
articulated rotor that was tested by Sikorsky Aircraft. We set out with very noble intentions on this contract. Sikorsky did a 
tremendous job in gathering the data, the physical properties of the blade — mass, stiffness — and they conducted a two-dimensional 
wind tunnel test to get the airfoil properties, and then they went into the tunnel. In our correlation effort it suddenly became very 
apparent that there are aerodynamic considerations for the model rotor that are not included in C-81 , because we could get nowhere 
near the Cj/ff measured, using the 2-D airfoil data. 

Also, we are in the process of trying to come up with a definition of: What is correlation? Bydefinition, all correlation is good, 
it's fair, or it's excellent. At Bell, when someone asks me about correlation, I usually tell them that correlation, like beauty, is in the 
eye of the beholder. So we must come up with a good definition, a workable definition of correlation. I would propose as a 
continuation of this study started by Bob Ormiston to continue in the same general approach that we started on the Sikorsky 
correlation program, and I would propose the fabrication and test of a full-scale 0012 rotor system with known physical properties. 
Quite often the question comes up - what is the torsional rigidity of a UH-1D main rotor blade? This is a very complicated question. I 
can assure that the results calculated by C-81 will be no better than those input data. So let's go through the process on a 0012 blade. I 
would next think that we should take that rotor system into a vacuum chamber test similar to what NASA Langley is doing. We 
should then take that airfoil into a two-dimensional tunnel for tests, because in regard to what Pete said, there's a wide variety of eg 
and C(j. The last point I want to make is I feel that in the helicopter business - in the loads prediction - there is a field that hasn't 
been brought to bear, and that is statistics. The statistical design of wind tunnel tests, and the statistical treatment of the measured 
data. I feel that if we can continue our correlation efforts, meaningful correlations, we can establish the integrity of these global type 
programs, and in the final result we will come up with a better and a cheaper helicopter. 



306 



Wayne Johnson 

Research Scientist 

USAAMRDL, Ames Directorate 

The results Dr. Ormiston has shown us have included many different assumptions and it's my own conclusion that the multitude of 
different results from current programs may be traced to the fact that we are dealing with phenomena that in many cases are on the 
edge of the current state-of-the-art, in structures and dynamics, and particularly in aerodynamics, and I'd like to concentrate my 
comments on a couple of the aerodynamic elements. Even a calculation with everything fixed but the representation of airfoil lift, 
drag, and moment coefficients at stall levels — even that comparison shows significant differences using the various dynamic stall 
models available. You find that the prediction of the overall performance parameters at high thrust is about the same, but there are 
still differences, from 30-50% in the peak-to-peak torsion and bending moments. And when you get down to the detailed 
characteristics of the aerodynamic lift and moment variation, you find that those differ very greatly, so it is quite evident that 
considerable fundamental research remains to be done in this area, and I think the state-of-the-art is in a similar position with vortex 
induced loads. It is interesting that both of these problems involve viscous three-dimensional aerodynamics which I think is probably 
the basic reason that they are so complicated and difficult to solve. 

I believe that some additional calculations of the type shown by Dr. Ormiston might be useful to isolate the specific differences 
as much as possible. In other words, concentrating on just the differences in the structural or the aerodynamic models. It has to be 
recognized, however, as I am sure all the members on this panel do now, that these calculations involve a considerable amount of 
work. It would also be very helpful to have some more detailed descriptions of the capabilities and limitations and assumptions of the 
existing methods. That alone would go a long way toward improving the assessment of the existing programs. 

One thing that is required in the area of experimental rotor airloads work, I think, is a greater attempt to measure more elements 
in this complicated interactive system. For example, besides just the blade airloads and blade motion, a simultaneous measurement 
also of the wake tip vortex geometry and induced velocity and the blade stresses is needed as well. With more information about the 
intermediate steps, it would allow a more direct assessment of the existing methods and also facilitate the development of better 
methods. It is clear, of course, that conducting such a test would be quite a job. 

Andrew Z. Lemnios 

Chief Research Engineer 

Kaman Aerospace Corporation 

In light of Bob Ormiston's comparison of all the analyses to date, I would suggest that perhaps we should modify Newton's 
second law to F = Kma where K is a function of the individual investigator. Along those lines, incidentally, I would like to 
apologize to Bob for not getting our results to him in time to be included in his presentation but they will be included in the written 
version of his paper. I would just very briefly like to explain the aeroelastic loads program that we use at Kaman, and how we couple 
that into our blade bending moments program. It is primarily a harmonic analysis program. 



Slide 9 shows primarily what we call our rotor blade design 
loop, and as you can see, we get our inputs from our various 
departments, our weights group and our stress group to give us 
our weight and stiffness distributions. From our weights 
department we get a breakdown of the blade weight as a function 
of many rotor stations and, obviously, our computer program can't 
handle that so we have to re-configure the whole blade and that 
takes a few days or even a few weeks perhaps, especially for the 
nonlinear type of mass distributions we are talking about. I'd like 
to get into a little bit more of that later on. Also, the same thing 
from our stress group and then finally after we have all of our 
physical characteristics we go through and get our blade statics 
and dynamics and go into our aeroelasticity and trim program 
which is what we call our 6F program, six modes of freedom. 
This is then used to evaluate blade stability from a linear 
standpoint and then we go into a nonlinear iteration procedure 
which is basically a harmonic analysis to calculate the airloads 
over a trimmed condition, and we force rotor trim to user 
specified hub forces and we just allow the control inputs to vary 
until we do get those forces at the hub. We then output all of the 
rotor performance characteristics. The airloads are then 
harmonically analyzed from the aeroelasticity program and are 
input into what we call our blade bending program, which is 
again basically a harmonic analysis, and these are used as forcing 
functions to generate the flapwise responses and edgewise 



ROTOR BLADE DESIGN LOOP 
ANALYTICAL TOOL 




PHYSICAL CHARACTERISTICS 



I STATIC DROOP, FREQUENCIES, 
MODE SHAPES, CENTRIFUGAL 
FORCE 



LINEAR TRANSIENT RESPONSE 

(NONLINEAR STEADY RESPONSES, 
LOADS, ANGLES OF ATTACK 

I ROTOR PERFORMANCE. 
HUB FORCES 



FLATWISE BENDING MOMENTS, 
SLOPES, DEFLECTIONS, 
VERTICAL ROOT SHEARS 

{EDGEWISE BENDING MOMENTS. 
SLOPES, DEFLECTIONS. 
HORIZONTAL ROOT SHEARS 



Slide 9. 



307 



COUPLED 
BLADE MOTIONS 

*FLAPPING 
*FEATHERING 
*LAGGING 
*SERVO FLAP 
*FLAPWISE BENDING 
*TORSION 



COUPLED 
INERTIA TERMS 

*CORIOLIS 

ACCELERATIONS 
*HIGHER ORDER 

NONLINEAR TERMS 



REAL 
AIRFOIL DATA 

*ANGLE OF ATTACK 
*FLAP ANGLE 
*MACH NUMBER 

Slide 10. 



responses of the entire rotor system. Slide 10 shows basically what is involved in the aeroelasticity program . We have the six modes — 
four rigid body modes — used for the stability analysis. Also, when we get into the nonlinear equations we incorporate all of the higher 

order nonlinear terms. For airfoil data we have tables of eg, c^, 
and c m as a function of angle of attack and Maeh number. We 
do not have the unsteady airfoil data that were generated either 
by Sikorsky or Vertol, we just have an approximation using 
Theodorsen derivatives which are truly applicable only in linear 
incompressible flow. 

I would like to echo Bill Anderson's comments regarding 
the lag damper. We have found that the lag damper characteris- 
tics are some of the least known characteristics that not only we 
but other manufacturers have, or don't have. For example, it has 
been assumed in most instances that when you talk about a 
damping rate for a lag damper you have a constant damping rate, 
and this is applicable over the entire range of frequencies. Well, 
this is quite valid and it's very good when you are talking about 
the rigid body lag mode. However, if you are going to apply a lag 
damper to a system, unfortunately it also works in the first edge- 
wise mode and the second edgewise mode, and your lag damper 
has a very nasty dropoff characteristic in damping rate as you go 
up in frequency, and a very equally nasty buildup in liquid spring 
rate as you go up in frequency. So your higher frequency edge- 
wise modes could be seeing effectively a cantilever root condition rather than a pin-ended damped root condition. This is something I 
think that needs to be further investigated and also something that needs further definition as far as the lag dampers themselves are 
concerned. Pop off valves and nonlinear characteristics of the lag damper are things that we all recognize physically, but yet we don't 
model very accurately in a mathematical sense. 

I would like to ditto again Bill and Pete's comments regarding steady and unsteady airfoil data. I highly recommend that we get 
some government agency back in the business again of generating some very reliable and useful airfoil data, both steady and unsteady. 

Richard H. MacNeal 

President 

The MacNeal-Schwendler Corporation 

I'm basically a stand-in for Jim Neff who did the actual work, so it perhaps is a little less self-serving if I add my congratulations 
to Bob Ormiston for what I consider to be a landmark achievement. 

Pete Arcidiacono sent around a letter to all of us prospective panelists in which he solicited some written comments and I didn't 
get around to sending mine in, but I think the questions he asked are very good, so I will repeat the questions, and I will give my 
particular answers. 

First question: What general type of correlation have you experienced in the prediction of flatwise, chordwise, and torsional 
moments at conditions near maximum steady flight speed? I'm a little hesitant to answer that one in view of the lack of correlation we 
have seen. Until I came here today I used to think that we got around 25% agreement between the peak values as measured in flight 
test and as calculated by SADSAM. In some cases I'm sure it's true, perhaps in others it is not. 

Second question: What do you feel are the most important potentially critical assumptions (if any) of your program in each of 
the following areas? Aerodynamics: It obviously has to be in connection with the static airfoil characteristics, I believe the moment 
coefficient is probably the most important one and the one that is least well known. Also, in connection with the method of dynamic 
stall which is used, which quite apparently, if you look at the results in detail, show that quite a lot of differences must exist in this 
area. Dynamics: In our program SADSAM — it is a fully coupled method — the most important assumption that we make is that the 
mechanical coupling effects are linearized about an average, deformed blade position, with one exception, which is that the Coriolis 
effects are treated in a nonlinear fashion. Method of solution: We use the SADSAM program which I might say is available at public 
data centers in case anyone wishes to use it. It is the only method of the group that was shown that at the same time uses a transient 
method of analysis in the time domain coupled with a non-modal solution. There are certain advantages to this I believe. In the first 
case, the transient method will show any potential instabilities whereas the harmonic method will not necessarily do that, and the 
direct method as opposed to modal avoids the problem of the lag hinge damper — not showing sufficiently high stresses. 

Third question: Which of the following would you recommend for evaluation of the various programs? (a) Further analytic 
computations using standard cases: Yes, I believe we should as a first instance try to resolve the discrepancies between the various 



308 



calculations. This maybe means peeking at the answers, which would be fundamentally experimental, (b) Experiments: I believe we 
ought to have wind tunnel tests on a simplified rotor where we know the facts about it very well. In the area of fundamental 
technology, we should do a great deal more to try to get common aerodynamic data. 

Question 4: What do you think has led to the current situation which is characterized by a multitude of programs yielding quite 
different results? Answer in 25 words or less. I can answer it in six words. First three words: Immaturity of the technology. Second 
three words: Lack of communication. 

Five: Do you feel that we or the government should be working toward a single rotor loads analysis following the general 
precedents set by NASTRAN? Now, I have a privileged position in answering this question because I worked on NASTRAN. I have a 
one-word answer to this question: No. 

Six: Is there any other major point that you would wish to make? The major point I'd like to make: Please carry on. 

Frank J. Tarzanin, Jr. 
Chief, Rotor Loads Unit 
Boeing Vertol Company 

Due to time limitations, I would like to discuss the program correlation for most of the time, and then just quickly go through 
the other issues. First, I would like to address the flap bending moment problem. Here we have, Slide 11, alternating flap bending 
moment data versus radial position for the H-34 helicopter. This data was obtained in the Ames full-scale tunnel and the reason we 
used the wind tunnel data was that we had a wide range of airspeeds from 100 knots to 175 knots, and hopefully we elminate some of 
the turbulence effects that are experienced in flight tests. The solid line represents the predicted flap bending moment envelope, and 
the four dots represent the test results. As you can see, the theory generally overpredicts in every case and the highest over-prediction 
occurs at the high airspeed, generally in the middle of the blade. This over-prediction is not a major concern, since the results are 
conservative when compared with wind tunnel data, and generally fall well within the bounds of the flight data scatter. In other words, 
were you to actually get a flight data envelope there would be quite a bit of scatter in it and the analysis would fall within that scatter. 



H-34 ROTOR 
FULL SCALE WIND TUNNEL TEST 



150 KNOTS 




£14 r 

£|2 

-4 

< 

Q- 8 

CO 

x 6 
o 

Z 

^ * 

I- 

u_ 2 



J 6000 



2 4000 



2000 • 



.2 .4 .6 .8 1.0 
RADIAL POSITION ( r /R) 



H-34 
FULL SCALE WIND TUNNEL TEST 

V=I75 KTS. 



1 



ijzS_ca£k£^_T0 



I I ANALYSIS 
V7A TEST 



iff 



rfl fba ,-w, 1L 



n-7, ru 



3 4 5 6 7 8 9 

HARMONIC FREQUENCY PER REV 



Slide 11. 



Slide 12. 



Slide 12 compares the lift and the flap bending moment harmonic components for the H-34 rotor. This is at the airspeed 
condition of 175 knots and the radial position is at 65% span, which is where we had the large over-prediction at the 175 knot 
condition. The analysis is represented by the open bars and the test results by the hash bars. As you can see, the analysis overpredicts 
the 1 per rev and this explains that the flap bending moment envelope over^)rediction is primarily due to the 1 per rev. Now, the 1 per 
rev is over-predicted in the middle of the blade and not at the root and the tip. I don't know why. If you look at the second, third, and 



309 



fourth harmonic components, there is really very good correlation. You can tell that the correlation tends to go downhill quite a bit as 
you get out to the higher harmonics, and it becomes clear from looking at this that you could expect to predict vibration for rotors 
with a small number of blades, but as you get out to a higher number of rotor blades the expected correlation I think would go down. 

Slide 1 3 shows the 3 per rev vertical hub loads versus airspeed and the phasing of these hub loads versus airspeed for the CH-46 
helicopter. Now, this data was obtained by directly measuring the loads in the hub drive system, the main shaft. The main shaft was 
instrumented with gauges and then cross-calibrated to get out all the coupling effects, so this is directly measured data. The analysis is 
the dashed line and the test, the solid line. As you can see, there is really very good correlation, both in the magnitude and the phase, 
and this is expected if you look at the H-34 results and notice that there was really good correlation — when you look at the 3 per rev 
airloads and the 3 per rev flap bending moments. And you would expect if you went to a five or seven blade rotor the correlation 
wouldn't be as good. 

Next I would like to discuss the pitch link load correlation. Now, all the previous pitch link load theory /test comparisons that we 
have shown, were compared with either model wind tunnel data or what we called smooth flight data. Now that was when we 
carefully went through the flight data and checked the flight cards and found flights where the data was very repeatable and the air 
was smooth, very little turbulence. In these conditions we found that we either predicted the loads or we were a little high, so this 
time we tried to just take scatter bands, just a number of flights at a given condition and compare the loads with the scatter of the 
flight test data. 

In Slide 14, we have the CH-47C helicopter at 46,000 lbs. gross weight, 4 in. aft eg, and the Cj/a of the aft rotor is 0.102. The 
solid lines represent the range of the measured pitch link loads for five flights. These are supposedly all the same conditions. As you 



3/REV VERTICAL SHAFT LOAD 

CH-46A FORWARD ROTOR MEASURED SHAFT LOADS 
GW = 19,520 LBS, CG 7.7IN FWD, = 264 RPM 



3,000 



<n 1,500 




a 2,000 



£ 1,0 




r 








CH-47C 




/ 




46,000 LBS 
245 RPM 
3,000 FT 

4" AFT 
C T /ir = .l02 






--T 

A^ T 




A 


a/ 





-TEST SCATTER 



-THEORY 



60 SO 

TRUE AIRSPEED - 

Slide 13. 



40 80 120 

AIRSPEED KTS. 



.05 .1 .15 .20 .25 .3 .35 
ADVANCE RATIO 



Slide 14. 



can see, the theory now predicts just the bottom of the scatter band, so clearly there are turbulence effects and so on that are just 
raising the loads. In Slide 1 5, we have the same thing for 50,000 lbs. gross weight, again this is for three flights. Again, a very large 
scatter band and the analysis is just predicting the lower end of this band. 

Last, and you can see why it's last, is the chord bending moment in Slide 16. Typically, we have not been able to predict the 
chord bending moment. Here we have chord bending moment versus airspeed. This is the maximum bending moment, and the analysis 
is predicting only about 40% to 60% of that load. Now, we have done a number of investigations and discovered that the equivalent 
viscous damping at the root end plays a very important role in determining what the maximum loads are. As you can see in the lower 
figure, we have been able to bring this predicted chord bending moment up to the test data by raising the equivalent viscous damping 
from 23,000 to 60,000. Now, really this is not an appropriate way to represent the damping. It is a much more complicated 
phenomena. There is a pre-load, there is a frequency effect, and an equivalent viscous damping doesn't seem to be adequate. Other 
things that may affect it are the actual chordwise stiffness. If you try to calculate chordwise stiffness, you have to consider if the 



310 



3,000 



O2P00 



£ 1,000 - 



CH-47C 

50,000 LBS 

245 RPM 

3,000 FT 

MID CG 

C T /cr» .112 




TEST SCATTER 



40 80 120 

AIRSPEED KTS. 



.05 .10 .15 .20 .25 .30 .35 
ADVANCE RATIO 



Slide 15. 



CH-46A 

GW»I9,520 LBS CG- 7.7 IN FWD 



LAG MOMENT 

VISCOUS DAMPING 



PRELOAD | 



- EQUIVALENT VISCOUS 
DAMPING 




20 40 60 80 100 120 140 
AIRSPEED~KTS. 



V = I35 KTS. 



= 10,000 




ANALYSIS 
C j • 23,470 
IN-LBS/RAD/SEC 



90 180 270 360 

BLADE AZIMUTH ~ DEG. 



Slide 16. 



trailing edge is buckling, or how effective are the skins, and so on. Another thing is the coupling. The C-60 program does not couple 
chord bending. It is coupled in flap/pitch, and chord bending is done at the end, uncoupled, and this may play a role, and finally the 
drive system. Charlie Fredrickson has shown that the drive system could play a very important role in determining the coupled 
modes of the chord bending. So this summarizes the correlation. 

I'd like to make one little comment on the downwash. On the downwash plots it was shown that the Boeing Vertol analysis was 
predicting about 1/10 of the induced velocity at the 80% radial station. We use an analysis very similar to Sikorsky's and what 
happened was that when we originally did the calculations, we had the same order of magnitude results for downwash that they had. 
But we had some flight data where we flew the CH-47C, which is a tandem, and varied the pitch attitude until we got the downwash 
from the forward rotor to run right into the aft rotor. When we tried to correlate with these loads we found that the theory was 
greatly over-predicting the flap bending moment. We went back, scratched our heads, and tried to figure out why. Well, we started 
iterating the downwash on itself. In other words, we let the downwash prediction affect the lift distribution and that would then give a 
new downwash distribution, which we fed back again on the lift, and we continued doing this iteration and we discovered that the tip 
vortex strength went down to somewhere between 1/6 and 1/10 of its original value, depending on the flight condition. So when we 
put that back into the analysis - at that worst condition - we found out that the load was now correlated. We then ran the normal 
separated conditions where the vortex was not running right back into the next blade and we found the correlation was still O.K., so 
we continued using this method of iterating the downwash on itself. 

One final thing — I think we should continue with this work. 



Richard P. White, Jr. 

Executive Vice President and Director of Engineering 

Rochester Applied Science Associates, Inc. 

As you probably noticed, I don't have any results up there. We at RASA have been working for about 6 years in developing 
various computer programs for rotor loads prediction, and free wakes, for NASA. I just couldn't find a way to finance this study, but I 
thought it would be very interesting if we could have because I'm not as discouraged as some of the other panelists. In the work that 
we have been doing we haven't really had the opportunity to have our own type of hardware to try to correlate data with. In the 
correlation we have obtained in the development and checkout of these programs, we have had to use what is published, i.e., what is 
available without all the background behind it. In doing that we have found that you have to look in very great detail at certain things 
you might not expect. I think that many times in the industry, probably when they are working with their own rotor system they can 
find a way that they Can turn a knob to obtain very good correlation with their rotor system. I think this is why I am not so 
discouraged. Ydu might have gotten the feeling that from the results that you saw, that rotor loads prediction analysis ain't worth a 
"Billy-be-damned." I don't think that's true. I think within each company, the way they have developed their programs for their rotor 



311 



systems, and the knowledge they have of their rotor systems, I think, and some of them have shown this, that the use of their program 
for their rotor systems does a far better job than it would do, necessarily, for somebody else's rotor system. In some respects that's not 
too-surprising, because a lot of the input is a function of art and not science. 

A couple of comments that I do want to make. What did surprise me in the results was that since (how much it might be laid to 
the difference in airfoil characteristics I don't know) the first three harmonics of loading are generated just by the trim control settings 
and q, I was surprised to see the large variation in the predicted results of the various theories and I can't lay it to the analysis being 
developed for one theory or another. It's too fundamental and too basic. A number of years ago, using published loads data taken for 
the H-34, UH-1 A, and the XH-51, 1 worked out a non-dimensionalizing parameter that had advance ratio and loads and the loading 
functions of the load that the total rotor system was carrying, and I found a very interesting thing. Here we had three different types 
of rotor systems, three different types of characteristics that were rather dramatically different. The first three harmonics of the sine 
and cosine functions laid right on top of each other. So there is not a big difference in these lower harmonics of the rotor and I'm a 
little surprised that we can't do a better job of prediction. A lot of the work and association that I have had with rotor loads has been 
the inclusion of wake analyses, free wake effects, and I think from what you saw today, the wake effect is an important effect. The 
free wake as compared to the prescribed wake is an important effect, and it's kind of surprising to me as well that on the two 
prescribed wake calculations that were done, the induced velocity distributions were so different. Maybe we're having trouble 
computing what the strength of the tip vortex is. I think that the reason for the work that has been done by a number of investigators 
starting way back with Ray Piziali and Frank DuWaldt, was based on the premise that you ought to be able to predict the lower 
harmonics of rotor loading very well, and I think maybe 10 years ago we were doing a better job maybe than we are indicating now. 
So what did we do? We said, "Gee, what you need is the wake effects in there to get the higher harmonic loading functions," and 
Frank and Ray showed that this could be done and then the industry picked it up and they worked it in their own way, which they 
should have done, and maybe because we have complicated the program so damn much to get the higher harmonics in, we have 
screwed up all the lower harmonics. I don't believe it. If the majority of the industry is computing the loads for their own rotor 
system, I expect almost excellent agreement and I'd believe the agreement in the lower harmonics. I do think there is more work that 
needs to be done, and I think Bob ought to take the responsibility for it and I think he ought to get himself a fairly refined rotor 
system, build the blades, incorporate the structural features and things that are needed in it, put instrumentation all over the place, 
and run some very controlled experiments to investigate the various areas that were proposed in the calculations that were conducted, 
and then he should spend the rest of his life analyzing that data so we would have some good consistent data. He opened up the 
Pandora's box, now he's got to get all the pieces in it [so] that we all have something to work with. He needs the induced velocity 
distribution, the loads, the responses, etc. I think it would be very bad if we did what we did in the past with the H-34, the UH-1 A, the 
XH-51, all the data was collected, it was all tabulated and put in thick reports and everybody got a copy and it went up on the shelf 
and that's the last time anybody really looked at it, except if they are looking for one particular point. I think in many respects that 
was a waste of money. All that data was there but never really analyzed. I would hate to see it repeated again. 



312 



SURVEY OF PANELISTS 

Panel 1 : Prediction of Rotor and Control System Loads 

Moderator P. J. Arcidiacono posed five questions to the members of Panel 1 before the meeting. A summary of the responses was 
to be presented at the conclusion of the panel discussion, however time was not available for this. The responses are pertinent to the 
attendee's questions and the panelists' discussion and warrant inclusion in these Proceedings. Responses to the questions were optional 
and not all members responded. The following summary of their responses was prepared by the moderator. 

1. What general type of correlation have you experienced in the prediction of flatwise, chordwise and torsional moments at 
conditions near maximum steady flight speed? 

- Flatwise moments (3 replies: Boeing, Kaman, Sikorsky): Predictions quoted to be within ±10%. 

- Edgewise moments (2 replies: Boeing, Sikorsky): Predictions low by 20 to 40%. 

- Torsion moment (4 replies: Boeing, Sikorsky, Kaman, RASA): Results inconsistent, indicating this is the most difficult to 
predict. 

2. What do you feel are the most important potentially critical assumptions (if any) of your program in each of the following areas? 

- Aerodynamics (5 replies): Unsteady aerodynamics were unanimously chosen as being of concern. 

- Dynamics (5 replies): Most answers indicated a concern as to whether all of the dynamic coupling terms had been included. 

- Method of solution (This refers to the techniques used to solve whatever dynamic and aerodynamic equations have been 
deemed worthy of solution): Not much concern expressed here but two replies indicated that the harmonic method may 
result in some limitations. 

3. Which of the following would you recommend for further evaluation of the various programs? 

a) Further analytical computations using standard cases: Five of seven responding endorsed further work here. (This time a 
funded effort, however.) 

b) Experiments 

1) Rotor: General agreement on need for systematic detailed data on large scale rotors tested in the 40 X 80 wind tunnel. 
Only one response also recommended smaller scale model tests as well. 

2) Fundamental technology: Steady and unsteady airfoil tests by NASA were the only specific tests recommended. In 
general, industry members appeared to favor rotor tests at this time rather than fundamental technology work. 

4. What do you think has led to the current situation which is characterized by a multitude of programs yielding (in some areas) 
quite different results? (25 words or less) 

This has resulted from the basic evolutionary nature of the scientific method which calls for hypothesis, correlation with 
test, revised hypothesis, test, etc. The process has tended to go on relatively independently in each company. The resulting 
analyses are molded by the data available to each group. To the extent that the data correspond to different configurations 
which may emphasize different aspects of the same overall problem, different groups will inevitably evolve somewhat 
different methods. This situation has been aggravated by the lack of systematic test data (and funds in some cases) against 
which to thoroughly correlate the analysis. Perhaps when all things are considered, the comparisons are really fairly close. 

5. Do you feel that we or the Government should be working toward a single rotor loads analysis following the general precedents 
set by NASTRAN? 

Seven replies received: Three no, two yes and two possibly. In general, there seemed to be some desire for such a program 
but a great deal of skeptism as to whether it was practical or even possible. There was concern as to whether any program 
could be general enough to anticipate all future configurations. For example, if such a program had been developed 10 years 
ago, would it have included all the hingeless rotor design variables since found to be important? One of the no responses 
stated that the money would be better spent supporting systematic rotor parametric tests. Also, of particular interest was 
the response of Dick MacNeal, who, of course, has been directly involved with NASTRAN. He was definitely opposed. 



313 



QUESTIONS AND ANSWERS 

Panel 1 : Prediction of Rotor and Control System Loads 

Bob Wood, Lockheed California Co.: I would like to again congratulate Bob for the study and also make several comments. In 
particular, I would like to just raise the question that has frequently been brought up, if we had a standardized method. I think if we 
had a standardized method, I think we would still face this F = Kma point that was brought up by the panel to some extent. There 
are, within each rotor system, variables that would have to be introduced by the manufacturers that would not be included in the 
model. Let me give several examples: Leading edge balance bars. Some of the manufacturers, for example, integrate these intergrally 
into the spar, which means that the CF load is picked up integrally along the blade. Others let all of that CF load be transferred 
right to the tip, which means then that you have a moment. That probably would not be incorporated in a universal model. Secondly, 
the lag damper, of course, also varies from manufacturer to manufacturer, and generally may have a relief valve and is highly nonlinear- 
Another aspect that we at Lockheed got involved in was in implementing the Bell C-81 global program. We found that in using it, of 
course it was developed for a hard swashplate control system, so we had to take one whole section of that program and remove it to 
put in the gyro control and the flapping feedback which are characteristic of the Lockheed rotor system. So I think even if we had a 
standardized method it would be tailored as it was picked up by the manufacturers. But I am curious with respect to Dick MacNeal's 
comment, having been one of the primary participants in the NASTRAN thing, as to what his reason is for saying no to having a 
standardized method since we are actually saying here that we would like to have standardized input but nonistandardized methods. 
I'd like to hear your comment, Dick, if I could. 

MacNeal: I see I didn't get away with my one-word answer. I think that there is great virtue in diversity, particularly when there is a 
great deal of doubt as to the physics of the problem, the methods of analysis, etc. If we settle on one particular approach, we will all 
use it and we will all go over the cliff together like lemmings going into the sea. 

Frank Harris, Boeing Vertol: Well, I have been over to the wind tunnel for the last 4 years and now I'm here and I think this problem 
sort of looks like a charcoal broiled steak, it's very well done on the outside and pretty damn bloody in the inside. I can't help but tell 
you that as the proceeding went on here I got madder and madder and madder, and I'll tell you why. I'm here at a dynamics meeting 
and yet at the conclusion I hear that everything is going to work fine if we get the dynamic stall. There's Ericsson to blame, Reding to 
blame, Tarzanin, Carta, "None," and Harris, and I refuse to sit here and accept the blame for what is obviously a problem. Then you 
complained that the wake isn't correct, so you are going to work on Piziali and DuWaldt and Scully at your next meeting probably. I 
think this is an excellent table that you have summarized here, but I don't think there is a guy in the room here, or two guys in the 
room that could write the equations from Lagrange, or work a list of equations of ma, and I personally wouldn't go one step further 
until I could see somebody write down ma. You've got a good chance to do it, you don't have any air particles to contend with, it's 
all structure. So maybe there ought to be 5 guys that sit down and if those 5 guys could write and then program successfully and 
numerically evaluate just ma, I would be willing to consider the problem further. 

Frank DuWaldt, Calspan Corp.: I have to observe that when industry does its internal comparisons it seems to do much better, and I 
would like to suggest that this is because they are tuned to their configurations by economics. Therefore my comment is that what we 
have seen today is the range of disposable ignorance and this is the part that you use to make your configuration look good to 
everybody else inside. 

Bob Wagner, Hughes Helicopters: I think I must take some exception to Frank Harris' comments. But the thing that I saw and felt 
this morning was a kind of structural connection between what the analysts are getting and what happens in the real world. Maybe the 
connection is only metaphysical and I think it must be, but those of us who at times have had the problem of taking helicopters off 
the production line and getting them to a proper degree of smoothness to be delivered, these machines are built presumably to the 
same specifications, the same drawings, and everything else. Yet, there is as much difference in the behavior of those machines in the 
real world as is indicated by the differences in the predictions of supposed identical components here. That would result then, in a 
proposal to perform an analysis, not on identical components, but on components which differ within manufacturing tolerances, and I 
would not care which analysis is used there. But some of the vibration characteristics that we experience I am sure are related to the 
differences in the components themselves and not to the differences of the methods of analysis. 

Ray Piziali, Vizex, Inc.: Some of the comments have been made by the panel members on how we should make our correlations. I 
think if you look at what has been done over the years, rather gross comparisons are made. If we all compare the lift values that were 
predicted up here, they compare very well; if you look at the time history of the lift at one radial station, they don't look quite as 
good, and if you look at the radial distribution of a given harmonic of the lift, the comparisons become worse and worse. Now for the 
application for which you are going to use your program, your predictive technique, some things are more important than others so we 
have no need to look at some of the detail. However, if you are trying to improve your analysis so that you can use it in many 
different applications, I think we must look at these details, look at the radial distribution of the harmonics of the load, the 
differences in the generalized force in each degree of freedom, etc. 

Ken Amer, Hughes Helicopters: I would like to follow up on Dick MacNeal's comment in answering the question of why he thought 
that there was such a wide disparity among the various programs, and three of his words were "lack of communication." Right now, as 
a result of lack of communication, in essence we are each the judge of the accuracy of our own program, and human beings being that 
way we generally convince ourselves that our programs are pretty good. I would like to suggest that under government sponsorship a 
critical cross-evaluation of programs. In other words, pick pairs of companies and pay each of the two companies to evaluate the other 
company's program. Personally, I would rather be criticized and embarrassed by someone else finding an error in my program rather 
than, on the day my helicopter flies, being embarrassed by finding out I failed to predict the blade loads correctly. 

314 



SESSION V 

APPLICATION OF DYNAMICS TECHNOLOGY TO HELICOPTER DESIGN 

Panel 2: Control of 1 /Rev Vibration 



William F. Wilson, Moderator 
Dale F. Benton 
Raymond G. Carlson 
R. J. van der Harten 
Robert Jones 
James J. O'Leary 



Panel Members 

Senior Dynamics Engineer, Bell Helicopter Company 
Superintendent of Helicopter Maintenance, Tenneco, Inc. 
Supervisor, Rotor Dynamics Section, Sikorsky Aircraft 
Deputy Managing Director, KLM Noordzee Helikopters B. V. 
Principal Research Engineer, Kaman Aerospace Corporation 
Chief of Dynamics, Boeing Vertol Company 



315 



THE USER'S PROBLEM 

Operational Viewpoint on Control of OneJ'er-Rev-Vibrations 

R. J. van der Harten 

Deputy Managing Director 

KLM Noordzee Helikopters B. V. 

Our company operates 3 Sikorsky S-6 IN helicopters on a 24 hr, fully IFR certified, service to oilrigs, and for harbor pilots, to ships 
on the North Sea. This virtually unrestricted day and night service requires the highest technical reliability of the airframe, its 
components, instruments and electronics, which can only be achieved by maintaining constant low vibration levels. Also it became 
clear, when IFR operations became more extensive, in particular for the harbor pilot service, that the tolerance of the crew for 
vibrations decreased with the increase in workload. 

Together with the National Laboratory (NLR) we measured the vibration levels, which we had been able to achieve in two of our 
aircraft on the basis of pilots acceptance, one being acceptable in one-per-rev (3.4 Hz), but barely acceptable in 5 per-rev (17 Hz) 
levels, and the other being acceptable in 5-per-rev, but barely acceptable in one-per-rev levels. It was found that: 

— pilots accepted only 3.4 Hz acceleration levels at or below 0.005 g laterally and vertically, and 1 7 Hz acceleration levels at or 
below 0.07 g laterally and 0.02 g vertically. Higher 17 Hz levels are mainly objectionable, because they cause blurring of 
instruments and are very fatiguing. 

— an increase in 3.4 Hz acceleration levels will increase 1 7 Hz levels. 

With the present S-6 IN rotor technology, the desired one-per-rev vibration levels could be maintained with a tracking flag 
method, which I developed as an analysis aid for rotor problems, when in 1955 I had the honor to be Jan Drees experimental test pilot 
and "ad hoc" vibration engineer for the development of the "Kolibri" ramjet helicopter. Basically this method consists of: 

1 . Tracking the rotorblades at low and high rotor rpm at low pitch, to obtain a reading of changes in track of the blades in 
relation to each other with rpm changes, thus with airspeed. Trailing edge adjustments were then made to obtain constant 
track between low and high rpm, being careful to obtain the minimum stick shake (for reversible controls), caused by 
unequal aerodynamic moment coefficients of the blades. 

2. Tracking at low and high collective pitch at high rotor rpm, to obtain a reading of a change in track indicating a chordwise 
balance problem, requiring a change in blade inplane position, relative to each other for two bladed rotors, or indicating an 
out of phase condition for articulated rotors, which can only be corrected with blade track changes. This is called the phase 
(or lag) tracking method, indicating that tracking is carried out for best phase instead of track. This is a major point. This 
discovery became the basis for attaining as low as practical vibration levels possible with the present state-of-the-art. 

The method requires very skilled pilots and engineers, and vibrographs if the pilot cannot feel and analyze vibration changes correctly, 
and can be rather time consuming. 

When I was responsible as chief test pilot for a U. S. Army overhaul program of their helicopters, the method was also used 
successfully on Sikorsky helicopters. As Sikorsky balances their blades dynamically to close tolerances on the rotor tower for equal 
mass and aerodynamic moments, vertical one-per-rev vibrations were only a problem if blades had been repaired badly, or damaged in 
handling at the trailing edge. Preventing this last problem with good workshop procedures, only lateral one-per-rev's had to be tackled. 
A lateral one-per-rev occurs mainly when unequal lift/drag properties of the blades, or damper characteristics, are involved, causing out 
of phase conditions. If the differences are not too extreme, changes in track of the blades, corresponding to the indications at low and 
high pitch, and carefully observing the change in vibration level at each adjustment will bring the blades in phase. The correct position 
in phase will be indicated when vibration levels are below perceptible and/or the blade will maintain correct track, relative to the other 
blades at low and high collective pitch (e.g., on the S-6 IN, a 1/4 in. down adjustment in track will decrease blade lag 1/4 in. at the tip). 
Of course, when using the flag method, it is necessary to make certain that dampers and rotorhead are in good condition, and blades 
are matched with less than 1000 flight hours difference. When correctly tracked for best phase in this manner, low and high blades 
may be up to 1 inch apart in track. The aircraft will then be smooth throughout its flight envelope and only needs a retrack if a blade 
is changed. If you have more than 1 inch out of track you may have trouble with vortices striking the other blades, giving a 5-per-rev 
vibration. 

When the aircraft and it's components became older, however, it was more and more difficult to obtain the best results, because 
the reason for the "out of phase" condition could not be made visible. Thus it became too much of a trial and error method. Also 
when using the flag method on the S-61N, it was only possible to obtain a reasonable high collective pitch reading when the wind was 
calm, due to the turbulent and heavy downwash, which at the length of the tracking flagpole to be used, could lead to blade-tip 
damage. 



316 



Thus we started an investigation of tracking equipment, which would allow the average pilot and mechanic to analyze and solve 
out of phase problems, which meant that the blade position in phase and track should be made visible even at the highest possible 
collective pitch position. The best solution seemed to be a stroboscopic system. Tests were carried out with the assistance of the NLR 
which proved that with this system, out of phase conditions could be made visible even at high collective pitch and wind conditions up 
to 20 knots. It was however very difficult to analyze the phase pattern, due to the large distance of the blade tips to the cockpit and, 
even with a concentrated light beam, it was only usable at night. Also the 5 strikers, required for the magnetic pickup to trigger the 
stroboscope, caused false and erratic phase readings. 

Around that time the Naval Air Service, which also became interested in our trials, was approached by the German firm Drello, 
that, under contract with the German Army, was developing a TV monitored stroboscopic tracking system. This system seemed to 
present the solutions for most problems because: 

— it could be used in daylight. 

— it used electronic phase control to keep the reflectors on the monitor, when changing rpm, thus blade phase position. 

— it could present selected blade pairs for comparison. 

— it presented the pattern of the blades on a monitor, which could be placed in the cockpit, thus enabling the pilot and 
mechanic to observe the changes of blade pattern with different rpm and pitch conditions. 

— it used a photocell to trigger the stroboscope, and thus prevented the inaccuracy found with strikers. This also saves the time 
for installing the strikers and magnetic pickup. 

Last but not least the system was offered at half the price of the Chicago Aerial Industries tracker as used by Sikorsky, which is more 
accurate. Thus we developed the system with Drello to its present stage, and acquired the first production type tracker in the 
beginning of last year. 

The first actual analysis and track adjustment with the prototype equipment was carried out in 1972 under supervision of Mr. 
Andy Lapaty, Supervisor of the aerodynamic test section of Sikorsky Aircraft. I must mention that at that time Sikorsky did not 
believe that track changes that small could cause proportional phase changes. The aircraft concerned was measured for one-per-rev 
acceleration levels before and after phase tracking. At 1 00 knots and 1 20 knots the former lateral acceleration levels of respectively 
0.008 g and 0.012 g were found to have decreased to 0.004 g or less. Measurements of the second aircraft which had been phase 
tracked before revealed even lower figures, namely 0.000 g at 100 knots and 0.0025 g at 1 20 knots. This last aircraft had already 
flown 300 hours after the last tracking. In fact, one aircraft flew 600 flight hours without changing its one-per-rev levels, and could 
have carried on without retrack, except for the fact that a blade had to be changed. Sikorsky has checked the phase tracking method 
on the rotor tower and on production aircraft confirming these very satisfactory results. 

As adjustments for best phase are made for the low to high pitch pattern indications, sometimes a slight increase in lateral 
one-per-rev levels, when on the ground has to be accepted. In the case of the first aircraft this increased from 0.01 1 to 0.012 g. It may 
be noted that vibration levels at low pitch on the ground are not an indication of the actual vibration level in flight. After a phase track 
therefore a hover is always carried out to ascertain that the one-per-rev is within limits. In-flight trackings have never been found to be 
necessary when using the phase tracking method, nor accelerometer checks, as best phase always corresponds with best lateral 
vibration levels. The Drello blade tracker is only used, after the blades are tracked first in the normal way with the flag, when the 
one-per-rev levels are found to be unacceptable by the maintenance pilot or by other pilots on normal flights. Installation, analysis of a 
problem and readjustment takes from 1 to 2 hours presently, depending on the type of trouble. 

Since the introduction in May 1972, the equipment has been used 12 times. Practically all major one-per-rev troubles with our 
aircraft have been found to be caused by damper discrepancies. It should be mentioned that these discrepancies also can cause an 
increase in one-per-rev levels with increased coning and flapping (high altitude, gross weight and turbulence), which are particularly 
disturbing to pilots when flying in IMC. 

In one case one damper, giving an erratic pattern in phase was found to be scored, one damper to be leaking badly internally and 
two leaking to a lesser degree, all on one aircraft. This aircraft even developed a vertical one-per-rev, probably due to the badly leaking 
damper which, without the Drello tracker could not easily have been located, even if it was noticed that this one-per-rev only appeared 
at high pitch at airspeeds over 100 knots, thus had to be caused by a damper. The equipment also makes it possible to predict incipient 
damper troubles indicated by lagging, leading or erratic behavior within still acceptable limits which are approximately a 1/4 inch out 
of phase condition to both sides of normal phase. This has been proven on one aircraft, where approximately 200 flight hours after 
this condition was noticed, we had to change the damper in question because of vibration complaints, which solved the problem. It 
also recently solved a vertical 5-per-rev problem in 2 aircraft, when scored dampers were located and changed. It is therefore a first 
rate accident prevention tool and has saved us quite a lot of time in analyzing a problem, assessing its danger to flight safety 
immediately, and correcting the problem at a suitable time or directly in between operational flights. 



317 



We still have a lot to learn, and this will take time as we have to wait for a problem to occur and then, by analyzing and 
correcting it, step by step, develop methods in such a way that reproducible results can be obtained to solve specific problems, so that 
instructions can be made for the mechanics and maintenance pilots to recognize and solve the different problems as presented on the 
monitor, even without the use of accelerometer indications. 

General Equipment Specification 

The Drello Blade tracker, model SKB 483, is designed to track helicopter blades. It consists essentially of a flashlight-stroboscope 
of high intensity, and a television unit. It permits accurate observation of the individual rotor blade tips for high/low and leading/lag- 
ging errors. Observation of up to 6 rotor blades is possible, and satisfactory results can be obtained with tracking in bright daylight. 

Scotchlite foil targets are attached to the blade tips, Slide 17. These targets will reflect incident light in the direction of the light 
source. This means that when the stroboscope flashes are directed on the blade tips from a distance of 10 to 20 m, Slide 18, they will 
be reflected in the direction of the lamp. Owing to the high light intensity of the flashes, the targets will reflect sufficient light even 
under bright sky conditions to have the built-in television camera in the reflector produce contrasted pictures on the monitor for easy 
evaluation. 





OPTERS B.V- 



c an i a 



Slide 17. 
The televised observations of the targets has certain advantages over the direct observation method, such as: 

— Accurate evaluation of the full scale images of the targets by applying a graticule to the monitor screen, Slide 1 9. 

— Observation of the target images from any desirable position, as for instance from the helicopter cockpit, Slide 20. 

— The tracking operation can be observed by several persons at the same time. 



318 







Slide 18. 




Slide 19. 



The flashes are triggered by a pulse light-barrier placed on 
the ground underneath the rotor and directed vertically against 
the sky. The light-barrier having a small aperture angle, the 
flashes are triggered with great accuracy as soon as the leading 
edge of the blade passes over the light-barrier. The actual trigger- 
ing pulse to the stroboscope can be delayed electronically with a 
potential meter thus correcting horizontal movements of the 
blades with changes of rotor rpm and pitch enabling the pilot to 
keep the blade reflectors on the monitor, while vertical move- 
ments of the blade reflectors are corrected by cyclic pitch. 

To determine lead/lag errors of the blade tips, the pulse 
light-barrier has to be offset by one blade section. 

When operating the equipment with switch "TRIGGER- 
ING" in position 3, no reference pulse transmitter is needed. 
Selector switch "BLADE NO." is set to the number of blades on 
the rotor for cyclic counting of the blades. With this method it is 
not possible to choose the blade which is to start the cycle. It is, 
therefore, necessary to use differently shaped targets for identifi- 
cation of the blades. By means of switches "BLADES- 
PRESELECTION" any desirable pairing of the blades is possible. 



Button "BLADE— SKIPPING" permits to offset the pairing by one section, e.g., if blades 1-2 had been preselected, the pairing 
will then apply to blades 2-3, 3-4, etc. 



319 




Concluding Remarks 

For 2 days, I have been listening to your excellent lectures 
on research to improve vibration control. I am happy that at last 
the manufacturers are becoming concerned with vibrations. I 
hope that from my story, you can see that we also do our best to 
improve on your somewhat wobbly product. Give us a chance by 
listening to and acting upon our complaints and don't be too 
skeptical about them. We have to make a living with your 
product, which may ultimately be your living also. I strongly feel 
that the present state-of-the-art could provide us with blades and' 
dampers which remain accurate and troublefree until they reach 
a reasonable T.B.O., which would prevent unnecessary 
one-per-rev problems — keeping our aircraft on the ground when 
they should fly. I know the industry recognizes the problems, 
but to my mind are weary to perfect existing systems which we 
have to work with and in which we have invested hundreds of 
thousands or even millions of dollars. 



Slide 20. 



THE USER'S PROBLEM 

Dale F. Benton 

Superintendent of Helicopter Maintenance 

Tenneco, Inc. 

After listening to all these fine presentations I would like to make a few statements which concern the people who put the work 
you gentlemen are doing into practical use. I have heard a lot of statements about the term assumption. Well, in our field it has to be a 
fact, we can't assume something is going to work. We have to make it work because the people we carry do not like it if it does not 
ride like a 707. And we are faced with a lot of problems. We are able to correct many of them, for others, there is nothing we can do. 
We just have to live with them. 

One of our most common problems is the 1/rev vibration which is usually left to the operator to take care of, to keep his ship 
free of this vibration. This can be done in several ways, but we have to rely upon a refined method to assure the lowest level of 
vibration. I have called upon the Chadwick-Helmuth Company, Inc. for support with their electronic tracking and balancing devices, 
that are working very well for us at this time. I would say it is giving us improvements of up to three times the life of some parts. A 
very important part of the increased life is the tail rotor and drive shaft bearings. By using this machine we have up to three times less 
parts usage as well as manhours to keep the systems functioning. This is true for all rotating systems, the engines, and other parts, just 
due to 1/rev vibration getting out of hand. We have had to go to the manufacturers with these problems because we are very limited in 
the field. We needed a backup and Chadwick-Helmuth supports us in this area. 

I feel that there are any numer of people here who are doing fine work but have probably never ridden in a helicopter out in the 
field when it's in service. I think this might be helpful, I feel that there is a gray area between your work and the area that is connected 
with the operation of the aircraft. I talk to some people and they say, "Yes this is my area, but that's not my problem." You talk to 
the other people and they say, "This is not my area, it's their problem." So when it comes into service we are left holding the bag. I 
think this panel is a good idea, but I think you should do a little more of what we operators do in the field. We have a real close 
relationship, even with our competitors, and we listen to each other's problems and try to work them out, because after all we are all 
in the same business and we have to make it go. I feel that you are doing a great job but possibly we are trying to advance too far 
without continuing to work out our past problems. We say, "This is good enough, this is all we can do." Well, I don't really believe 
this. I think there is enough knowledge and technical equipment available for us to go ahead and develop a much smoother system. 
Thank you. 



320 



TECHNICAL ASPECTS OF 1/REV VIBRATION 

William F. Wilson 
Senior Dynamics Engineer 
Bell Helicopter Company 

There are five topics that I want to talk about. Before I get into these topics, I have four slides of a test that we ran at the Army 
Aeronautical Depot Maintenance Center where we actually ran through a series of four rotors and made some step-by-step adjustments 
to see the sensitivities of various parameter adjustments that can be made on a rotor: sweeping, balancing, tabbing and rolling the 
blades. 

Slides 21 and 22 show the effect of blade balance on a UH-1D helicopter at 80 knots. Slide 21 for the spanwise balance shows 
how much 2 ounces of blade weight will produce in terms of pylon motion. Now this is vertical motion measured at the pylon at 1 per 
rev. Similarly, Slide 22 shows the chordwise sweep adjustment and the effect of adjusting the drag brace. An interesting thing that we 
observed in that test concerns the correspondence between tab and pitch link adjustments. Slides 23 and 24 show the vibration level 
measured at the pilot's seat at 80 knots airspeed on both aircraft. Each step in Slide 23 represents 1/2° tab angle change. You will 
notice that 1/2° for a little tab out on the tip of the rotor causing a change of .04 g's. Keep that in mind when we talk about levels in a 
minute. Slide 24 shows that for 4 minutes of blade pitch change the vectors are in virtually the same direction. This indicated to me 
that the difference between tab and roll changes in a blade were pretty small. I mean they virtually give the same effect. I don't have 
the plot at 120 knots, but there is a slight air speed effect difference with the tab. It does increase with air speed slightly, but it's not a 
significant change. 



PYLON MOTION CAUSED BY 
VARYING SPANWISE BALANCE 



PYLON MOTION CAUSED BY 
VARYING CHORDWISE SWEEP 




0.001 0.002 

INCHES 



R.H. VERTICAL 
TRANSMISSION 
MOTION AT ONE/REV 

UH-ID 80 KTS 

EACH STEP REPRESENTS 
A 2 OZ. WEIGHT 
ADDED TO BLADE BOLT 




R.H. VERTICAL 
TRANSMISSION 
MOTION AT ONE/REV 

UH-ID 80 KTS 

EACH STEP REPRESENTS 
SHORTENING THE DRAG 
BRACE BY .0023 INCH 



Slide 21. 



Slide 22. 



PILOT SEAT VIBRATION CAUSED 
BY VARYING TAB SETTING 



PILOT SEAT VIBRATION CAUSED 
BY VARYING PITCH LINK LENGTH 




PILOT SEAT 
ACCELEROMETER 

UH-ID 80 KTS 

EACH STEP REPRESENTS 
A 1/2 DEGREE TAB 
ANGLE CHANGE 




PILOT SEAT 
ACCELEROMETER 

UH-ID 80 KTS 

EACH STEP REPRESENTS 
FOUR MINUTES BLADE 
PITCH CHANGE 
(.009 INCH) 



Slide 23. 



Slide 24. 



321 



DISCUSSION 

Panel 2: Control of 1 /Rev Vibration 

Moderator Wilson: The first topic is what levels should we be shooting for? I have a couple of slides, Slide 25 is from a 1964 NASA 
report* on bioastronautics and it shows perception levels in the 3-5 Hz range of about .0025 to .0030 g's. Now this is the kind of 
number that Mr. van der Harten has been talking about. What can you feel? It doesn't check with what my experience in helicopters 
has been. It's surprising that you can feel down that low, but it's possible that for IFR we may have to go there. You can see that 
unpleasantness starts around .03 g's but the perception level is an order of magnitiude lower than where it becomes unpleasant. So the 
question becomes: What is the goal? 

100 

5 
4 

3 

2 



" 








































































































f 


r~ ' 


■H 










vt 


LUNTARY TOLERANCE 
UNINJURED 










/ 


-t- 

A 


\ 


| 










\ 






L/^o 


r N 


^T 














z 




1 


"I 


I 


^* 


V l 




/ 


S MINOR INJUR* 

/depending ON TIME 

1 1 1 














1 






\\ 


1 1 




— Short time tolera 
*-*Qne minute tale 


nee 
















S2 




' \ 


klw 




ance 
















i 


?R 


\ 






II 










-A 

v 






V\ 






w 


-T 


iret 




nute to 










- i 


A 


/,_, 


^ 


$5 


f^ 


s \\V 










u/ 








J 




\ 


f 


i l 




T 






^NXS 3 


S3 


\\ 


\V>\ 


sS\\ 
















i 


— "i ' 'r 






> V "1 


SS 




^X> 


Xss 1 - 


^ 


^ 


1 


N-- 








/ 






U 


/^ 


^ 


^ 




(^^ 


Cx^ 


•^ 


^ 


^ 










/ 






'1 


/ 








P 


P 


/ 






at ion. tolerance 
nilitcry aircraft 

g exposure) 


/ 




3LEf 




*\ 




•: lor r 

:• (Ion 


; 








f •:>•:■■ 
'/W 

f :■:: 






• • <j 


• snv-;-, 


w. 














/ 










■'■■■■• 






^ ■'•■■'•••■•^i 
































.UN 


'LEASA 


1TNESS 






Threshold for perception 


1 
































•p 




















A 


,m 


\ 


1m 












Z 
















V 












J 














— ::. 
















\ 


^4*. 


* 














/in 

/ '" 
















Thraiholcf for percept! or 
at fingertip (tongentiot 
vibration) 












r •:: 




PTI 

I, 


ry : 


















,-UJJL 






' llll 










-Ull. 








MM 



2 3 4 5 10 20 30 40 50 

— •» FREQUENCY - eycln/ne 

*■ -Jfj-fMC -1 ) 



100 200 300 500 1000 



Slide 25. 

Consider just the 3-1/2 to 6 Hz range and what various people have said over the years about what the level should be. Slide 26 is 
self-explanatory. Our experience is that peoplewon't start adjusting rotors until vibration gets above .03 g, that's about where they will 
begin to think, "Well, we've got to do a rotor adjustment" and they can generally reduce it to that range. But, I have not really seen 
the case where they can really feel down below .01 g, because the other frequencies bothered them to the point that they don't do 
that. I would like to ask the panel members for their reaction. What should the desired goal be for a vibration level at main rotor 1 per 
rev and should the levels vary with rpm? 

Carlson: Mr. van der Harten has done an excellent job of either describing us or exposing us, depending upon your overall attitude, I 
guess. We are able to achieve and s ought to aim at levels on the order of .005 g and that experience shows that levels on the order of 
.01 g are felt by the individual in the aircraft. These are levels that are achievable as Mr. van der Harten has indicated and I would 



♦Bioastronautics Data Book, Edited by Paul Webb, NASA SP 3006, 1964. 



322 



© 

© 
© 
© 



PERCEPTABLE CABIN VIBRATION 



.08- 
.07- 
.OH 
.05- 

.04- 
.03- 

.02- 
.01- 
.00 



Goldman 




- Betheseda 1948 




Gorrill and Snyder 




- Boeing 1957 


... Q> 


Parks and Snyder 


, 3 


- Boeing 1961 


gl 


Chaney 


ra <c 


- Boeing 1964 


CD CD 


Bell Experience 


8 .E 


- 1974 


<C In 


Data From KLM 




Sikorsky Experiment 




- 1974 





©.05 
.035 

.025 



© 
.003 



01 



.01 



© 

.005 



Slide 26. 



imagine then that this is our aim. In order to reach them, it is a 
complicated process and all of us have to work hard at it. Now, 
Fm only able to speak of the levels of our aircraft which have 
1 per rev that is rather low. They are under the levels of normal 
body resonance and I am unable to answer with regard how this 
enters into the other aircraft which have higher rotor speeds. I'll 
pass.it on to others at this stage. 

Jones: The UH-2 has automatic tracking which we will be get- 
ting into later. From an actuation standpoint the automatic 
tracker actuates at .014 g's so therefore in theory we are keeping 
our helicopter within .014 g's of 1 per rev. It has hysteresis in the 
system so that when it gets down to about .007 g the actuator 
cuts off and then does not start again until the level reaches 
.014 g's. Now as far as frequency goes, I think Bell and Kaman 
are approximately in the same area. Bell runs at about 5.4 Hz, at 
least in their UH-1, which again is the human factors problem. 
Our UH-2 is approaching that. I think we are about 4.3 or 4.4 Hz 
from a 1 per rev standpoint, so we are at least approaching this. 
So I assume there would have to be some form of rpm variation 

on acceptable levels. And to really illustrate this, again in our development of our UH-2 helicopter, we had a vertical bounce on the 
ground, which we worked very hard to get rid of, but I couldn't even tell you how we did, it just seemed to disappear, but we did get 
rid of it. That vertical bounce was very near the human factors problem area, the test pilots could hardly write on their knee pads, so 
levels can be quite important. 

O'Leary : I would just like to make a few comments in a couple of areas. One is that the newer specs — as we relate what's happened 
to the vibration spec requirements at 1 per rev — have obviously driven us all down towards attaining better levels, as we come from 
the .15 g's down to the .05 g's of the AR-56 and the later generation aircraft specs. We have done pilot seat shake testing under 
controlled conditions and reported this at various symposiums. This has shown that levels like .02 g and .03 g are acceptable for what 
our experienced pilots consider long term operation — for an excellent oi outstanding qualitative assessment of the aircraft. It is 
assumed that it would not affect them operationally or fatigue them. I think the important thing that we found from our 
own flight experience on some of our later helicopters is that as the n per rev levels are being driven down to meet .05 g's at b per 
rev or n per rev, the awareness of the 1 per rev is greater than it was before. They are complaining more about other frequencies now 
that we are getting rid of the ones that used to be up in the .3-.4 g range. In fact our testing showed that when we did the testing by 
individual frequencies and attained levels that would receive an outstanding rating for long period operation and then took the same 
attained g levels at each frequency — taking the same g level that gave an individual rating of outstanding and adding just 1 per rev, 
4 per rev, and 8 per rev together, the rating became completely unacceptable. So in the presence of other frequencies, the require- 
ments for 1 per rev are even lower than perhaps we thought in the past. Our current goal was around the .03 g range and it has been 
demonstrated I think on one of our tandems and on the BO- 105 that those kind of levels are attainable. 

Wilson: You can see that we have some disagreement among ourselves as the load calculators do, as to what the flight levels are. Now 
experience has shown to me in the flight tests that we have done that the .01 to .03 range is attainable on most rotors and we are 
somewhat in agreement with Vertol on that line. 

Our next topic for discussion, I want to keep very brief. The question is: What is being done by industry to insure that these 
goals are being met? We are doing two things at Bell. We are looking into the possibility of automatic devices. We have had a couple of 
programs in the last two years. We are also doing work on the blades themselves and trying to make better blades using what we call 
closed cavity bonding tools that will give better control of contour and twist in our rotors. Recent tests have shown that at least on a 
first-try basis for the first 10 or 15 blades we have made, we haven't needed a tab. We've just been rolling them into track and going, 
and it has been very successful and satisfying to date. Would you all care to make any comments on this? 

O'Leary: I would just amplify the point about the closed mold process for controlling the airfoil contour. We really feel that that has 
been demonstrated on aircraft like the BO-105 and our tilt rotor Model 222. There is another area that we also are concentrating on 
and that is the control of the fuselage frequencies around 1 per rev. On our large tandems, generally from an air resonance standpoint, 
we have lower limits on where these frequencies can be placed and that generally assures that you are not going to have a frequency 
for the lowest mode, especially for very large helicopters like the heavy lift, that will be coincident with 1 per rev at some operating 
condition. So there are structural changes that Dick Gabel will talk about later actually being installed on this aircraft to keep the 
modes closest to 1 per rev above 1 .6 to 1 .7 per rev. On our other helicopters we are consciously aware of keeping the amplification of 
these modes whether they be above 1 per rev or below 1 per rev to some nominal value like 2 and we are so conscious of it that, for 
our current aircraft, we plan to shake test the aircraft to assure in fact that the fuselage modes are far from 1 per rev so that any 
residual loading won't be amplified by the structure itself. 

Jones: The only comment that I really have is that, although right at the moment I don't get in with our service group or our 
manufacturing groups, we do keep our blades in very close tolerance, and as far as the stated question: "Are we doing something about 



323 



it?" Well, I sure feel that Kaman has at least done something about it in the fact that we do have automatic tracking in flight on all of 
our UH-2 helicopters, so that track is maintained in flight. 

Carlson: I'm only able to add here, I guess, that at Sikorsky we work hard at the initial balance and tracking of the blades in that each 
blade is balanced with weights at the tip and chordwise we have tabs on our blades which are used. We have a master set of blades to 
track each of them against and then ultimately each blade is tracked. These all give us what are in general reasonably smooth levels. I 
would add a small item, I suppose it ought to be brought out that it isn't only the blades that enter into this, it's the hub as well in 
that we have to have our hubs lined up and well made also. 

Jones: May I just interject something here, I just want to beat Sikorsky out on one point here. We don't have a master blade, we have 
a master master blade. 

Carlson: We also have that. 

Jones: Oh, do you! I got a big kick out of it. I didn't realize we had a master, master blade, and that is the blade that we whirl to get a 
master blade which we then use to whirl blades to meet tolerances. I assume then when the master blade wears out, we whirl another 
one to get another master blade, to the master master blade. 

Wilson: Concerning that subject, we will talk a little more about whirl testing in just a minute. I would like to go to the next slide and 
if we can I would like to keep the comments very brief on this subject. Really I would like to find out whether we are in favor of or 
not in favor of electronic tracking and balancing devices for correcting main rotor and tail rotor balance. The question is: What about 
electronic tracking and balancing devices? Can we save dollars for the operator? 

Benton: We are definitely sure that electronic devices are good. Some of these are really expensive but they are valuable to the 
operators in dollars saved. We used to take a 1 per rev out of the ship until we would get into 2 per rev that we were unable to do 
anything about and then they failed. So many of us thought we were inducing the 2 per rev so we would leave the 1 per rev to cover 
up something that you folks had stopped short of and left in. So with this electronic machine we found out that we were actually 
hurting ourselves by covering up the 2 per rev. 

van der Harten: Well, our good friends at Sikorsky build a wonderful aircraft. Their S-6 1 is considered as a kind of DC-3. We'll let that 
alone, I think, but we don't have that problem which you have, Mr. Benton, not as much I should say, but we save a lot of money by 
controlling vibration. The only thing you need is time to do it, so you need a quick analyzing method and that's what we have been 
trying with the Drello method. Also, you run into problems if you are not able to observe carefully what the phase is doing and what 
the track is doing. And I always hear track, track, track - I don't care about track. As long as my blades run the same paths through 
angles of attack and through changes of airspeed, I couldn't care less. The only thing I worry about: If I go too far out of track — let's 
say 2 inches, that's about the ultimate limit, the maximum limit we have flown — you will get into the same trouble as Mr. Benton is 
talking about, you will get into trouble with the n per rev. Apparently you get vortices hitting the other blades. That is a theory, my 
good friend Andy Lapaty would say, "You are theorizing again, prove it to me." Well, that's our problem, to prove these things, but 
we find out things and I think it's up to the industry to prove what is really happening and we hope that you will do it and give us the 
equipment to quickly get through it. But in talking about dollar savings, 1 per rev causing 5 per rev, or 5 per rev being caused by taking 
out a 1 per rev, in our case — when we were able to prevent that, after a lot of experimenting — we found a big saving, in the other 
expensive indicators and VGI's, unscheduled removals. You are talking about $ 1 ,000 repair every time. If you save that for five years, 
and it was more than that I can tell you, you have saved $5,000. If the equipment cost $10,000, it takes you two years to get out of it, 
and the main thing is that you don't have your aircraft on the ground all the time because the pilots come back and complain about 
them. 

Benton: If I may I would just like to add one thing. Talking about dollars saved, with the electronic devices we can usually do a ship 
in four hours say, but before we had them, sometimes it would take 3 or maybe even 4 days to work out a problem, so that's a big 
savings in time, and the machine off of the job, and that is many dollars. 

Wilson: To go on to the next topic: What about whirl testing blades against a master? Does it pay off since airspeed cannot be 
simulated? Here is something that the operators don't get into, some manufacturers do. We haven't whirled a blade at Bell I guess in 
10 years on anything but a helicopter. The other manufacturers here all whirl every blade that they build. They talk about whirling 
against the master and bending their tabs and getting them set. I would like to see if they really think it's justifiable since airspeed is 
not present on the whirl stand. Maybe we don't understand what you are doing. 

O'Leary : We do whirl our blades and we do adjust the tip weight to take care of chordwise moment, and we also adjust tabs to take 
care of rpm variations, and also track the blades within pretty close tolerances. I think the general philosophy and the general 
occurrence has been less time is spent on inflight tracking if we do that. That doesn't mean that we never do inflight tracking, but less 
time is spent on that and less flights are required for that. We really think we are reducing the number of variables for the inflight 
tracking problem when we take into account these things such as balance, chordwise, spanwise, and the effect of rpm, and therefore 
we will reduce the number of inflight tracks that we have to make. 



324 



Jones: Again, I'm just going to have to say that I agree with Jim O'Leary. I just want to point out, I think Kaman possibly has more 
of a tracking problem than other conventionally controlled rotors. We use a servo-flap; this means a very soft torsional system for our 
control. It means I think that our tolerances have to be much better; therefore, we do rely upon whirling and really matching blades. 
We also have one other item we have to match and that is not just the blade but we have to match our servo-flaps and so therefore we 
really feel we have to do this. From the standpoint of costing I have no answer for that aspect except I think that if we did not do that 
we'd really have tracking problems. And again, we really do rely upon the automatic tracker with the inflight system . 

Carlson: Well, I would only want to point out one item here. I suppose it holds that the more blades that there are on a rotor the 
harder it is to track it wholly on the aircraft. It's almost impractical if there are a large number of blades as we have, and our answer is 
that all of our blades are tracked against the master. 

Wilson: I think it would be to their interest to run some programs to see — we have run programs — to see where various factors have 
entered into the cost of helicopters and I'm certain that whirl testing does enter into the cost of blades. Just like balancing or anything 
else enters into it, and with the electronic tracking that we are doing now maybe that's a possibility that they might be thinking about. 

O'Leary: I think a comment that I would like to add to that is that — with the new generation of blades where we really feel that the 
airfoil contour control can be much better maintained with better production processes such as closed molds — in fact this change 
with airspeed will be less of an effect and has been less of an effect on aircraft that do this. And therefore by tower tracking and 
getting all the variables out, the one remaining variable which is this characteristic of the airfoil, differences will be minimized. 

Wilson: We have one more topic. The question came up with Kaman's rotor and with ours also, we've been experimenting with 
automatic trim devices. We've looked at electric pitch links. We made one system that would automatically bring the rotor in track in 
flight. You could bend the tab up or down by 4 degrees and still track the rotor by adjusting the pitch length. We had both manual 
and automatic control so that we could feed back an accelerometer signal into it. We decided after we got to that point that the 
question was really: Does the operator really want automatic trim devices? Does he need them? 

van der Harten: Well, flying Sikorsky helicopters, I would say we wouldn't need that because if you set your blade on the ship and if 
the blades stay like they are - I mean do not crack up somewhere and have to be taken off, or if the dampers do not go out of whack 
which is the most important problem, internal leakage, valves and things like that — then you don't have to change track, the aircraft 
will remain smooth. If the blades would hold out without having to change them for 2000 hours I'm pretty sure I wouldn't track the 
aircraft for 2000 hours. If the dampers wouldn't give any trouble. Actually it is different on Bell helicopters and aircraft that do not 
have pre-tracked rotor blades. I think that what we have found that the pre-tracked rotor blade will not change its characteristics for 
quite a number of hours and will remain interchangeable. The main problem is that the dampers go out of whack. 

Benton: I really don't see the need for automatic pitch change links, etc. It's an added expense, and to the people in the field and for 
many one-aircraft-operators, it's just an added complication and expense to them. I will go back to one thing and agree with Bill 
Wilson, if they could do away with the tabs and hold the blades to a little closer tolerance. The tabs are continually getting bent, a 
man throws a tie-down strap on, he bends the blade, they run it into the wall, or they will bend it during installation. So we have 
built-in trouble to start with, but as far as the automatic tabsare concerned - no I don't think we need them. 

Jones: I sort of agree with the operators, an automatic system could be expensive from the standpoint of maintenance and so forth, 
but how about some form of indicator so you would maybe know at least which blade is a problem, which one is high and which one 
is low, not necessarily an automatic tracker but at least this would tell you where the blade is in flight and when you came down you 
could possibly do something about it. 

Benton: The pilot usually tells you that when they get out a little bit. 

Jones: Well, do they know which one it is really? 

Benton: No, but that's where the electronic device comes in. We can throw it on in about 30 minutes and we can tell immediately 
what has happened. Then for a one man-hour job, we could maybe replace this automatic device we would put on for many dollars. 

van der Harten: Well, I disagree a little bit, but that's mainly because of flying larger helicopters and thinking about airline use of 
helicopters. There it would be nice from the flight safety standpoint to have an indication that something has gone wrong in the rotor 
head. The articulated rotor is the smoothest rotor system you can fly, especially if your blades are pre-tracked as well as they are at 
Sikorsky. Your problem is when something goes wrong in the rotor head itself, dampers again. I am fed up with dampers after I have 
found so many with our Drello tracking system. If you have a system in your aircraft that we will show you that the blades are going 
wrong in phase rather than in track which could be shown by photocells or something like that in the fuselage, and the aircraft was big 
enough to take the expense, I would say do it, but that's for bigger aircraft mind you. 

Wilson: Well, of course the size of the device may be pretty small. Even on a 206 which is a 3,000 lb machine, we might be able to 
afford a pound or two for keeping the rotor in track if we can afford so many other pounds for so many other things as we have often 
done in the past. 

325 



QUESTIONS AND ANSWERS 

Panel 2: Control of 1 /Rev Vibration 

Bill Paul, Sikorsky Aircraft: I have a comment on Mr. Wilson's comments. First of all, I think when your vibration level gets as low as 
it is getting to on the Noda-Magic system , I think you will find that you might require more careful control of your track, particularly 
in the commercial environment. That's my observation, but it may not be correct. I think that 1 per rev is a very subjective thing, it 
depends on what you are used to. If you are used to — as Mr. van der Harten is, as well as he knows — a very low vibration level, then 
.005 g is not acceptable. If you are at .01 g to begin with, you might have thought .005 g was fantastic. The second point is, Mr. 
O'Leary's comment, I don't think we are going to improve 1 per rev. I may be wrong, but my observation is that with cambered 
airfoils, where we are now talking about 9,000 in-lb versus 900 in-lb for symmetric airfoils, that even though you are exercising better 
control, the control almost has to be precise and I don't mean just precise coming out of the barn, but precise after a few thousand 
hours of flying, maybe a few hundred hours anyway. So one of the things that I would suspect is going to happen with cambered 
airfoils is that — and we are all using them, by the way — is that you are going to see more out of track in forward flight, and the 
problems may not occur on the ground or on a whirl stand. The same thing holds true with higher twist blades and I think that will 
show up so I think that we are coming into an era where we might find more of the out of track taking place, *- I mean out of track in 
this sense, high speed, and not so much the 1 per rev concern. Finally when you talk about composites, and we are all using those as 
well, you damn well know that control of the weight is very difficult even with molding techniques that we have, so I think the 
problem is just beginning in a different era. 

O'Leary: I think from the weight standpoint the tolerances on the balancing chordwise and spanwise that we will use with these 
blades is even more stringent than we had in the past which recognizes that fact. 

Dick MacNeal, MacNeal-Schwendler Corp.: I believe Mr. O'Leary that you were the only one who mentioned the problem of 1 per rev 
structural resonance. I would like to ask whether you would care to comment on the problems posed by extremely heavy wing stores 
such as long range fuel tanks. 

O'Leary: I guess my only comment is that if they can change the characteristics of a structure to move modes closer to 1 per rev, then 
obviously you would have to consider means to avoid that. 

Dick Gabel, Boeing Vertol: Speaking about the tracking again versus the vibration level, we have also had a lot of experience where in 
special programs to improve track on test vehicles, we have taken great pains to match pitching moments and pitch link loads and tabs 
and a whole raft of things. We've gotten the vibration level down at the expense of getting blades out of track, which is often possible, 
because if there are discrepancies from blade to blade you can do better by putting them out of track than in track to get the g's 
down, but when we do so the pilots turn out to be more concerned about that disturbing shadow out there than they are about the 
1 per rev, so its really a balance between the aesthetics of it and the g's of it. 

Jones: May I just make a point about that. Again, when I talk about our automatic tracking, admittedly we are tracking to g level 
and not track. However, I've never heard that comment from our pilots. I just sort of want to make the statement that we haven't run 
into that one specifically. 

van der Harten: Well I don't think the pilots are bothered about blades out of track. We are talking about flying real IFR, I mean 
going up with no visibility, letting down with no visibility, and at the last moment seeing where you are going to land, take off, 
vice-versa again, turbulence, everything. They don't care about blades off track, we have flown with blades out of track, no problem 
there. They do care about vibration, because vibration disturbs the high work that they have when flying under those conditions. So if 
you can do it by bringing your blades out of track, and maybe for the bigger ones by getting automatic control in there, to reach the 
higher speeds without 1 per rev vibration, then do it. I'm getting scared a little bit that in the future technology you are calculating a 
higher 1 per rev load, which I heard from the first question. I think this should be prevented at all cost if you want helicopters to be 
acceptable to passengers, since they are used to vibration levels of 707's and 747's. The second thing is the pilots will be bothered very 
much by it and we still don't know the effects if you are on longer flights and you have to cope with vibration levels that are shaking 
you up. It is all at the wrong frequency, 1 per rev, 5 per rev, it is all between 2 and 20 Hz and, as is proven in many reports that does 
something to your body system, and how much we don't know. 



326 



SESSION V 

APPLICATION OF DYNAMICS TECHNOLOGY TO HELICOPTER DESIGN 

Panel 3: Integrating Dynamic Analysis and Helicopter Design 

Panel Members 
William F. Paul, Moderator Chief Engineer, Sikorsky Aircraft 

Rodney W. Balke Assistant Group Engineer, Dynamics, Bell Helicopter Company 

Richard Gabel Manager, Rotor Technology Staff, Boeing Vertol Company 

J. Ford Johnston Senior R&D Engineer, Lockheed California Company 

James R. Neff Chief of Dynamic Analysis, Hughes Helicopters 



327 



DISCUSSION 

Panel 3: Integrating Dynamic Analysis and Helicopter Design 

Moderator Paul: We are not going to spend too much time on formal discussions this morning because I think this has to be art 
interactive panel in discussion with you. Therefore, what we are going to do is to spend 5 minutes, each of us, summarizing ouj 
position as to where we stand with integrating dynamic analysis into the design and where we think we should be going, and then open 
the discussion to the floor because I think that's where the most fruitful discussion should take place. So, I would like to proceed 
immediately with Mr. Rod Balke and have him present his position. 

Prepared Comments 

Rodney W. Balke 

Assistant Group Engineer, Dynamics 

Bell Helicopter Company 

The design process that we are all immersed in can be broken down into three basic phases: the concept formulation, the iterative 
design loop, and the final design and documentation (Slide 27). I have chosen, as a format, to take you through a three-month 
response to a military RFP. The concept formulation may already be complete at the receipt of the RFP. The final design and 
documentation we know from experience will take 1 to 1-1/2 months. This leaves approximately 1-1/2 months for the iterative design 
loop. For the dynamicist, as you know, that's where it's at. Now we are going to talk a little bit about - what it's like, where it's at - 
to make a few points, and hope to stimulate some discussion. 

To set the stage, Slide 28, we have requirements for rotor frequency placement, loads, stability, and requirements for vibration 
control, which for Bell is generally a rotor isolation system. For the fuselage we have frequency placement and vibration levels, for the 
drive system we have frequency placement and stability. We have major programs in each of these areas. As you have heard, we use 
C-81 for rotor work. We have NASTRAN available, although detail modeling must be done. We do start in many cases, particularly in 
the vibration control area, with simple 1 or 2 degree-of-freedom systems to keep that gut feel. 

TO SET THE STAGE 

ROTOR 
DESIGN PROCESS ~ f"!^ ENCY P «-ACEMENT MYKLESTAD 

(3 MO. RFP RESPONSE) " SStlTY DYN5 

CONCEPT FORMULATION UIHJ . Tinw rmlTDn , 

(USUALLY COMPLETE AT RECEIPT OF RFP) ^"r^OR SKKton SYSTEM NASTRAN 

"355X fl S /2 G M07° P FU - S FR A EQUENCY PLACEMENT NASTRAN 

(AHHHUA. 1-1// IWU.J VIBRATION LEVELS 

F,NA SAPPnnv /D ?TMO i ; AT,0N """'FREQUENCY PLACEMENT NASTRAN 

(APPROX. 1-1/Z MO.) STABILITY HYBRID 

Slide 27. Slide 28. 

The first pass, Slide 29, the rotor frequencies, begins with extrapolated parameters and assumed hub impedance. The loads begin 
with level flight and maneuvers using the existing program version. The vibration control begins with the development of a NASTRAN 
model of the isolation system on a rigid fuselage, the point being to keep it simple. The fuselage vibration, again a simple NASTRAN 
stick model. We start program modifications at this point or new programs as required. Slide 30, the tempo picks up a bit; the rotor 
frequencies, the first real parameters become available, but we are still using assumed hub impedance. We find that maneuvers are 
taking too much time, we go to a fallback position and switch to a hybrid version of C-81. For vibration control we obtain hub 
impedances from the NASTRAN model. This goes back to what Jim Staley was pointing out, the need for integrating the hub 
impedance from the fuselage analysis into the rotor system, combining the rotor analysis with a program such as NASTRAN. For 
fuselage vibration, we develop a 2-D model of the fuselage and start a 3-D model. The tempo continues and we've got to hurry to meet 
the schedule here, Slide 3 1 . We get a parameter update, we now have some hub impendances, we complete the loads and forward them 
to rotor stress. We obtain rotor excitation forces to apply to the fuselage. We combine the vibration control and the fuselage vibration 
into one 3-D model; we must now freeze program modifications. We then come up to the 4th pass, Slide 32. The fatigue group finds 
some minor stress problems, the rotor is beefed up in this area. For vibration control we apply the rotor forces and get the first look at 
the vibration picture, and suddenly we are struck with the design freeze and no further changes are made. Now the points being made 
are that we should keep the model simple consistent with the design status, which provides fallback positions. We should support the 

328 



analysis with model tests and correlation, particularly for checking out new analyses. We should not support that desire to make 
another pass, because it has to be documented; and we do need an integrated rotor/ fuselage analysis. 



FIRST PASS 



SECOND PASS 



ROTOR 



FREQ - EXTRAPOLATED PARAMS. AND ASSUMED 
' HUB IMPEDANCE 

LOADS - LVL FLT AND MANEUVER/EXISTING 
PROGRAM 



VIBRATION CONTROL 

- DEVELOP NASTRAN MODEL ON RIGID 

FUSELAGE 

FUSELAGE VIBRATION 

- DEVELOP SIMPLE NASTRAN STICK 

MODEL 



PROGRAMS 



START MODIFICATIONS 

Slide 29. 



ROTOR 
FREQ 



- FIRST PARAMS AVAILABLE/ASSUMED 
HUB IMPEDANCE 
LOADS -MANEUVERS TAKING TOO MUCH TIME, 
SWITCH TO HYBRID VERSION OF C81 

VIBRATION CONTROL 

OBTAIN HUB IMPEDANCES FROM NASTRAN MODEL 

FUSELAGE VIBRATION 

DEVELOP 2-D MODEL OF FUSELAGE 
START 3-D MODEL 



Slide 30. 



THIRD PASS 

ROTOR 

FREQS - PARAM. UPDATE + HUB IMPEDANCE 
LOADS -COMPLETED LOADS FORWARDED TO 
ROTOR STRESS 
-OBTAIN ROTOR EXCITATION FORCES 

VIBRATION CONTROL AND FUSELAGE VIBRATION 
-COMBINE ISOLATION SYSTEM AND 
FUSELAGE INTO 3-D MODEL 



PROGRAMS 



FREEZE MODIFICATIONS 

Slide 31, 



FOURTH PASS 

ROTOR 

FATIGUE GROUP FINDS SOME STRESS PROBLEMS 
ROTOR IS BEEFED UP IN THIS AREA. 
NO FURTHER CHANGES MADE. 

VIBRATION CONTROL AND FUSELAGE VIBRATION 
APPLY ROTOR FORCES AND OBTAIN FIRST LOOK 
AT VIBRATION PICTURE. 

DESIGN FREEZE 



Slide 32. 



Richard Gabel 

Manager, Rotor Technology Staff 

Boeing Vertol Company 

I'll limit my remarks to one portion of the design integration process, Slide 33, the rotor loads to airframe analysis, and it 
parallels certainly what Rod Balke has just shown. The upper part of Slide 34 is a flow diagram. You apply criteria on frequencies and 
stresses and compute frequencies and go around that loop until you modify the blade for those requirements. Then, concentrating on 
the vibratory loads going into the fuselage, the lower set of plots show some vibratory hub loads. We try to have as a backup to all 
those computer programs — which as you saw this morning are "very unreliable," other methods to make sure that we are conservative 
in the prediction of the vibratory loads. And one that we have been finding quite useful is models where we scale the hub loads from 
dynamic models. As a backup to even that we have also used the supply of information from previous aircraft and scaled that to the 
new aircraft, where that's possible. We look at all those and take the highest and work from there. 

On Slide 35, looking at the airframe, in the work on the HLH we first established the maneuver loads and vibratory loads for the 
airframe. We do a preliminary shear moment diagram in the old fashioned way, work with design in sizing the original stringers and 
frame and then as that begins to solidify we get more detailed in the model that's used and start to use NASTRAN as shown on the 
next slides. There's a whole series of them, it's so complicated now that we break up the fuselage into pieces even in this preliminary 
design stage. There's the forward pylon, Slide 36, and the forward cabin, Slide 37, and the next one, Slide 38, is the aft cabin and 
finally, Slide 39, the rear pylon on the HLH. Of course that's because it's such a huge aircraft. The way this impacts design is that 
nowadays we size stringers and frames to the NASTRAN prediction and get that solid line of EI, Slide 35 - grouping it on the EI as a 
simplification. We specify the strength and dynamics requirements, and then we judge the results with the very same model. The use of 
one model has been a very big improvement. We used to have multi-programs and have legislated those out of existence pretty much 
and stuck with NASTRAN for one basic model for both stress and dynamics use. Now it is established that the vibration requirement 



329 



■ ROTOR BLADE, HUB AMD CONTROL SYSTEM VIBRATORY LOADS PREDICTION/ 



BLADE DESIGN CRITERIA 

- NATURAL FREQUENCIES 

- ALLOWABLE STRESSES 

- STIFFNESS 

- CONTROL SYSTEM LOADS 

- WEIGHT 

BLADE PRELIMINARY 
DESIGN 



INTERACT IPS OF AIRFRAME DYNAMIC ANALYSIS AND DESIGN 



COMPUTE BLADE NATURAL 
MODES AND FREQUENCIES 
COMPOTE BLADE, HUB, 
AND CONTROL SYSTEM 
VIBRATORY LOADS 



MODIFY BLADE MASS 
AND STIFFNESS, 
CONTROL STIFFNESS, 
ETC. 



ARE CRITERIA 

SATISFIED? 



- WEIGHT 

- STRESS 

- STIFFNESS 

- NATURAL FREQUENCIES 

- VIBRATION 



VIBRATORY 
BUB LOADS 



PRELIMINARY DESIGN 



# PRELIMINARY ANALYSIS 



T t 



DETAILED DESIGN 



HLH PREDICTED VERSUS DYNAMICALLY SCALED FORWARD ROTOR 4/REV 
VERTICAL AND FIXED SYSTEM ROLL-PITCH MOMENT 



——DYNAMICALLY SCALED FROM MODEL 347, C T /o = .081, 
V TIP " 750 FPS 
O COMPUTER PROGRAM PREDICTED 



4/REV ROLL-PITCH 
VIBRATORY MOMENTS 



STIFFNESS MATRIX 



STATIC 

EQUIVALENT 

LOADS 



DYNAMIC ANALYSIS 

- MODES AND 
FREQUENCIES 

- FORCED RESPONSE 



STRESS ANALYSIS 



CRITERIA 
SATISFIED? 



C IDENTIFY STRUCTURAL 
MEMBERS FOR MODI- 
FICATION 

9 MODIFY DESIGN TO 
REDUCE STRESSES, 
TO REDUCE DYNAMIC 
RESPONSE 





4/REV VERTICAL 
VIBRATORY FORCES 






LBS 




4,000 


/ 


2,000 


r — ^y 








60 100 140 180 
AIRSPEED - KNOTS 




100 140 
AIRSPEED - KNOTS 



INCORPORATE OPTIMUM 
STRUCTURAL CHANGES 
IN DETAIL DESIGN 



Slide 34. 



EXAMPLE - HLH AIRFRAME DESIGN 



AIRFRAME CRITERIA - REQUIRED STIFFNESS 



. STRENGTH REQUIREMENTS 
WEIGHT STRUCTURE 



Slide 33. 



VIBRATION CRITERIA! AVOID NATURAL 
FREQUENCIES NEAR 1/REV AND 4/REV AT 
MINIMUM WEIGHT 



is different and we are faced with working out the differences 
between the two of them. And here we have been able to stiffen 
the fuselage in the area at mid station around 600 so as to beef 
up the stringers to satisfy the vibration requirement where it's 
obviously greater than the strength requirement, and then at the 
frame 200 to 400 region where the vibration says it should be 
softer than the strength, we compromise by considering cutouts 
in the airframe to accomplish the softness without compromising 
the strength. Another advantage of these NASTRAN type pro- 
grams is that you get very detailed stress information on each 
member, and hopefully the airframe will be lighter by virtue of 
the analysis having gotten so involved. 



2. 



■■■VERTICAL 
1. 




_>- 



wim 




200 400 600 
AIRFRAME STATION - INCHES 



1000 



Slide 35. 



330 



H L H FWD CABIN 



HLH FWD PYLON 




GRIDS: 225 

ELEMENTS: 602 

MASS POINTS: 28 

LOAD CASES: 100 




Slide 36. 



Slide 37. 



H L H AFT CABIN 




HLH AFT PYLON 




GRIDS: 422 

ELEMENTS: 1198 
MASS POINTS: 33 
LOAD CASES: 100 



Slide 38. 



Slide 39. 



331 



/ Ford Johnston 

Senior R&D Engineer 

Lockheed California Company 

I 

A year or so ago I was able to participate in a very successful program, the AH-56 AMCS, and what I observed there led me to 

some thoughts on the management and the attitudes of the dynamicist in such a program. I found this program was one in which the 

dynamicist was very closely in the loop and participated in the preliminary design, and carried on through. And rather than go into the 

details of how this is done, I thought I would like to talk about some of the principles that were evident to me in this process. 



CLOSING THE LOOP AT LOCKHEED 

WE HAVE TO: 

BE AN ACTIVE ELEMENT IN THE LOOP 

HELP ESTABLISH THE DESIGN CRITERIA 

BE RIGHT, NOT CONSERVATIVE 

DEVELOP A TRACK RECORD FOR "RIGHTNESS" 

BE MEANINGFUL - COMMUNICATE 

WE USED AN OMBUDSMAN 

Slide 40. 

A REASONABLY RIGHT ANALYSIS 



> 

u 

z 

o 













• 


FULLUNST . ' . £ = 4 - CONF | G No- , 
C"(l<, m, h)=I.O/ ° fl =0-02 

TEST DATA REDUCED TO 
COMMON BASE CONFIGURATION 




•* 

^^1 




ARROW SHOWS 
AMOUNT OF t 
CORRECTION X 










-^, 


L 











260 280 300 

ROTOR SPEED RPM 



320 



^CORRECTECT — JO ROLLER 
BEARINGS 




Slide 41. 



Slide 40. I found that the dynamicist has to be a very active 
element in the loop, to help establish the design criteria, and in 
order to be this and to do this, he has to be right; he has to be as 
nearly right as he can and not give himself margins of 
conservatism. And he has to have developed a track record for 
being right so that management can pretty well depend on his 
inputs, and then of course he has to change from the 
nomenclature that he uses with his fellows and be able to 
communicate to the design people, the aerodynamicists, etc. 
Having been in that position I put in this last part that we call the 
ombudsman. It's a term to indicate a man who can help you at 
City Hall if your case is right, and since I was in that position as 
staff to the director of engineering, I was able to say, "Yes, that 
makes sense," or, "Let's have another go at that." And I think 
that it's very useful for the dynamics group to be able to work 
against a chief engineer who has a background and experience in 
dynamics. Now, just to illustrate a couple of points, if I could 
have the next slide, please. 

This one I call a reasonably right analysis, Slide 41. It came 
early in the development of the AH-56 and it pointed to the fact 
that we had a potential 7P flutter problem. What we had in the 
beginning was the dashed line to illustrate linear analysis, and just 
recent we completed — as reported in one of our contractor 
reports* — a full wake analysis, and that was pretty good, it 
showed good correlation. Actually the first analysis alerted us to 
the problem, and showed us the ways in which we could fix it, so 
I considered that pretty good. 

Slide 42. That was unfortunately conservative, it is a re-do. 
Put yourself in the position that you are looking at the analytical 
curve in the absence of the test data because the rotor hasn't 
been whirled yet. You look at that analysis and you say, "With a 
little bit of spread that could be unstable within the expected 
overload gross weight range of the vehicle," and so it affects your 



*"Rotor Blade Wake Flutter" by W. D. Anderson and G. A.Watts, 
Lockheed Report LR 26213, Lockheed California Co., Prepared 
for U. S. Army Aviation Systems Command and USAAMRDL, 
Ames Directorate, Contract No. DAAJ01-73-C-0286 
Mod. P00005, Dec. 1973. 

Companion Reports: 

"The Reactionless Inplane Mode of Stiff-Inplane Hingeless 
Rotors" by J. F. Johnston and F. Conner, LR 26214, Dec. 1973. 

"Flight Test and Analytical Data for Dynamics and Loads in a 
Hingeless Rotor" by R. E. Donham and S. V. Cardinale, 
LR 26215, Dec. 1973. 

"Flying Qualities of a Gyro-Controlled Hingeless-Rotor 
Compound Helicopter" by A. J. Potthast and A. W. Kerr, 
LR 26216, Dec. 1973. 



332 



\ risk analysis of this system, and this is a second go around in which a more conservative look was taken at the couplings involved. The 
'■ first analysis goes down closer to the 0.02 range prediction-wise and when the test results came out we breathed a sigh of relief, but 
nonetheless it cost us quite a bit. 

Slide 43 is one that Bill Anderson showed you yesterday and it was an excellent example to me of the communication of taking 
the results not of a root locus plot but of turning that into terms that other people can see, roll rate per inch of stick and inplane loads 
per inch of stick. You can see that too small a blade droop angle would get you into a loads problem and into multiple problems of 
roll response, and too large a droop angle would get you into a possible excessive response, a too lightly damped roll mode. The actual 
design value of 2° droop was taken. 

THIS ANALYSIS WAS UNFORTUNATELY CONSERVATIVE 



0.06 



c T /°- 

0.08 0.10 



0.12 



ROTOR LIFT, 1000'S LB 
15 20 25 



30 



STABLE 



O 

Z 



-0.04 




REXOR TOWER MOMENT, 
a, -I 1/2% I IN -LB 

A ^ MIN. O0 

DMPG. □ |00K 

O200K 
A 300 K 
V 400 K 



Slide 42. 



THIS PRESENTATION GOT THE POINT ACROSS 




FREQUENCY - CPS 
Slide 43. 



333 



James R. Neff 

Chief of Dynamic Analysis 

Hughes Helicopter 

I think the basic problem is a matter of getting inputs from a dynamic analysis into the initial stages of the design at the earliest i 
possible time, and this is a very serious problem and one where we need to have programs that may not be quite as rigorous as/ 
NASTRAN or the global programs, and that we can get some information put together and get some data together in order to affectf 
the initial stage of the design. Providing these inputs usually requires extensive variations of parameters to be made on structures 
whose characteristics are loosely defined and in a state of flux. The airframe at that point is not well defined, stores configurations are 
usually not defined, they certainly are being changed. So we need analysis methods that are inexpensive and require a minimum 
amount of input data preparation in order to meet design schedule deadlines. Sufficient rigor of the analysis must be retained to assure 
accuracy of results, at least consistent with the level of definition of structure existing in the initial design stages. You have to have 
enough accuracy in your analysis to be able to depend on it to some extent. As the design proceeds and becomes solidified, more 
detailed and rigorous methods are justified. The programs we are using at Hughes to do this are SADSAM and NASTRAN. Another 
thing that we use in the early stages is, as Dick Gabel mentioned, quite a bit of scaling from existing test data and from models. May I 
have the first slide, please. 

Slide 44 is supposed to be faintly humorous, I know my boss laughed the first time he saw it, but maybe it's more a crying matter 
than a laughing matter. Anyway that's about the kind of problem that we face in trying to meet the requirements of the design people. 



The Bask Problem 






FINITE ELEMENT PROGRAMS 








SADSAM 


NASTRAN 


PROBLEM SIZE LIMIT 


80 D.O.F. 


PRACTICALLY 








UNLIMITED 


TYPE OF MODEL USED 




LINEAR BEAM 


THREE 






AND TORSION 


DIMENSIONAL; 






ELEMENTS 


FRAMES, PLATES, 






("STICK MODEL") 


SHELLS 


INFORMATION REQUIRED 




GROSS SECTION 


DETAIL ELEMENT 






PROPERTIES, 


PROPERTIES, 






GEOMETRICAL 


GEOMETRY, AND 






LAYOUT, MASS 


MASS DISTRI- 






DISTRIBUTION 


BUTION 


TIME TO OBTAIN AND ASSEMBLE 






INFORMATION 




DAYS TO WEEKS 


WEEKS TO MONTHS 


COSTS TO RUN TYPICAL PROBLEMS 




• NORMAL MODES 




$10 to $80 


$50 -$400 


• DAMPED MODES 




$15 to $40 


_ 


• FREQUENCY RESPONSE 


$6 PLUS $1.00 


$50 - $500 






PER FREQUENCY 




• TRANSIENT RESPONSE 


$20 to $50 





SPECIAL CAPABILITIES 




TRANSIENT RESPONSE 
OF NONLINEAR 
SYSTEMS, ROTATING 
COORDINATE TRANS- 
FORMATION (GROUND 
RESONANCE. ETC.), 
COMPLEX CONSTRAINTS 





TIME FROM DESIGN INCEPTION 

Slide 44. 



Slide 45 is a little example here of the two types of analysis 
that we are using and a comparison to give an indication of what 
they are being used for and why. The SADSAM program is 
limited in size of course and NASTRAN is practically unlimited. 
In the case of SADSAM, linear beam and torsion elements or 
stick models are used. This is a simplified analysis and it's more 

adaptable to getting some early answers out. With NASTRAN we Slide 45. 

set up three-dimensional problems with frames, plates and shells, 

and of course it is rigorous and by the end of the design we need this type of analysis. The information required: Cross-section 
properties, geometrical layout and mass distribution — those are gross section properties — for SADSAM. This is information that you 
can get together fairly rapidly. In NASTRAN, detailed element properties, geometry and mass distribution; it takes quite a bit of time 
to get this ready and it's been my experience at least that we don't have that type of information ready in time to apply it early in the 
design when it's really needed. The cost to run these programs is a significant factor in that you need to make variations of parameters 
and you need to run a lot of cases. If you have a program that costs you a stack of money to run one case and you want to run 
20 cases, you pretty soon run out of budget, and of course we have stringent budgets to meet, I think that's true of everybody. We 
have some special capabilities with SADSAM of course which are not available in NASTRAN. We can do transient response of 
nonlinear systems, we do rotating coordinate transformations for ground resonance, and conceivably we could match fuselage 
impedance to the rotor. We have not actually used it in that sense so far, but we could. We can put in complex constraints which also 
would help solve the rotor fuselage interaction problem. 

Slide 46. These are the problems that we apply the programs to. For main and tail rotors, we get natural frequencies, response to 
flight loads, steady stability, and mechanical stability, using SADSAM both for preliminary design, and final design and substantiation. 
There is quite a difference I think between the type of program you need early in the design and later on for final substantiation. For 
the fuselage and fixed aerodynamic surfaces we use SADSAM to solve for natural frequencies, response to flight loads, unsteady 



334 



TYPES OF PROBLEMS 



PROBLEMS 


SADSAM 


NASTRAN 


MAIN AND TAIL ROTORS 






• NATURAL FREQUENCIES 


A,B 




• RESPONSE TO FLIGHT LOADS 


A,B 




• STABILITY (FLUTTER AND DIVERGENCE) 


A,B 




• MECHANICAL STABILITY (GROUND 
AND AIR RESONANCE) 


A,B 




FUSELAGE AND FIXED AERODYNAMIC 
SURFACES 






• NATURAL FREQUENCIES 


A 


B 


• RESPONSE TO FLIGHT LOADS 


A 


B 


• STABILITY (FLUTTER AND DIVERGENCE) 


A,B 





A = PRELIMINARY DESIGN 

B= FINAL DESIGN AND SUBSTANTIATION 

Slide 46. 

stability — flutter and divergence, in the preliminary design stages. Then we use NASTRAN for the final design and substantiation of 
fuselage natural frequencies and responses to flight loads and to determine vibration levels. 

William F, Paul 

Chief Engineer 

Sikorsky Aircraft 

As moderator, I'm going to wind up the session with a few minutes of discussion and observations. I'm going to address myself 
solely to the rotor designer's challenge and problems because I think he is in a unique position. He has a product which is very difficult 
to change and which is very vital to the helicopter, and yet he is very reliant on everyone else's input to it. I think that's very much 
different from the problems that you have in the airframe or even in the transmission area where you tend to control your own 
destiny. In this area, and having been there, I'm faced with the first problem which is kind of a dichotomy. On the one hand, you can't 
afford to make a mistake, even with respect to picking something as small as the trailing edge strip, whether it be graphite or fiberglass 
to control edgewise frequency, because if you do make that mistake, you probably won't discover it until the aircraft is flying, but 
you've already qualified the rotor blades and you have spent a great deal of money in that qualification program. So even a small a 
problem like that begins to trouble you. On one hand then you need to have the best tools you can get. On the other hand — the 
dichotomy part - the requirement for innovation in rotor design won't wait for documentation and correlated analysis.You've got to 
proceed, even though you don't have that analysis, so that's what the designer is faced with. In my case, and our case at Sikorsky, I 
guess we handle this problem by wanting to get as many methods as possible to exercise in order to examine the decisions that we 
make. Anywhere from the most simplistic to the most sophisticated, and what we are really looking for is the red flag. We make the 
decisions on the basis of the fundamentals, but try to find out where the red flags are, where the problems are and then trace those 
down and make sure they are believable. Now, I would like to show you one example of how in the past we might have made a wrong 
decision but using advanced technology - and I really mean that, advanced technology, particularly with respect to unsteady 
aerodynamics — we were able to make a design decision that we wouldn't have made in the past. 

Slide 47. This is a composite blade which we are developing now, and we had to make a decision on the airfoil, whether it would 
be cambered or symmetrical - that's really a cambered airfoil with a reflex trailing edge to reduce the control moment. The elastic tip 
deflection in degrees is plotted versus blade azimuth angle, and this is an indication of what the blade tip is doing and what the control 
loads are doing. Now, the attractiveness of a symmetrical airfoil is the lower load, and when you look at the performance from an L/D 
point of view, particularly on the advancing side, it didn't look like the symmetrical or cambered airfoil had much difference, so when 
we used the steady aerodynamics and the constant inflow the result would indicate that the decision should be made for the 
symmetric airfoil because of the same performance and the lower control loads. 

Slide 48 is the same case but run with unsteady aerodynamics and variable inflow. It took a week to get this plot but we were 
willing to wait that week. Running time to us is not that important when we are running the first or second cases. Two things 
happened. First of all, we saw a general increase in the steady moment, but more importantly, the symmetrical airfoil exhibited a high 
frequency oscillation; you can call it stall flutter if you want. Now, the question is whether to believe this information, that's what the 
designer faces, and that's where we spent the month to try to understand this result. Was it the input data, was it the information that 
we had, or was it in fact true. We did find out, as a matter of fact, that the symmetrical airfoil in Slide 47 did in fact have stall except 
that torsional tip deflection and the control load didn't show it. 



335 



o 

1- 
o 

LU 

_l 
ti- 
ll! 

o 
a. 

i- 



< 
_j 
uj . 



10 



2 r 



I NOSE UP 



SYMMERTICAL (REFLEX CAMBERED) 
AIRFOIL 




90 180 270 

AZIMUTH ANGLE, deg 

Slide 47. 



360 



.SYMMETRICAL (REFLEX CAMBERED) 
AIRFOIL 




90 180 

AZIMUTH ANGLE, deg 

Slide 48. 



270 



360 



But in Slide 48, what happened was that the stall showed up in the stall flutter because we used unsteady aerodynamics and 
variable inflow. So having looked at this thing and dissected the problem — we spent a month doing it — we finally made the decision 
to go to the cambered airfoil, and accept the larger steady moment but the vibratories would be considerably lower. Now, two 
messages come up here. First of all, we are using advanced technology in design. We've got to use it. If it's available, we will use it, and 
we are. Secondly, I think you have made fantastic strides, regardless of the inaccuracies, the qualitative trends are there, and that's 
what we are looking for. You have made fantastic strides in getting unsteady aerodynamics and variable inflow into our programs. And 
I think that this is a very profound thing because in the past I remember that looking at the difference between steady and variable 
inflow there were very small effects, but it seems like the incorporation of unsteady aerodynamics and its impact on stall has had a 
very profound effect on our decision making process. But the other item and the thing that I would like to ask you to do please is to 
be sure that in your programs you have the ability to dissect the problem. We've got to understand the answer that you get. There's no 
question about it. We will not make a decision based solely on computer output, so if it takes us a month to get that answer, we might 
have to make an arbitrary decision before then. That is a problem and an area where you must concentrate your efforts. 

Now, Slide 49 shows the same thing, the elastic windup 
versus azimuth, but what we are looking at here is a conventional 
low twist blade with a symmetrical airfoil, no camber 
whatsoever, and the same airfoil that we showed before, the high 
twist cambered airfoil. Now, the main point I want to get at here 
is that on the advancing side we are running angles of about 
6° -8°, now maybe this is not right, maybe it's 4° -6° or 6° -8° or 
8° -10°, I don't know, but the key thing is that we are showing 
that with the high twist blade and the cambered airfoil and the 
kind of aspect ratio blades that we make, you are going to get a 
significant elastic windup of the blade, particularly at high speed. 
Now, in doing so we've got to incorporate this sort of 
information into our performance decks, because when you go to 
ask the performance engineer for the computation of 
performance he gives you the information where the line is right 
across here, the rigid blade analysis. Generally speaking, that's 
true since he has his own computer decks. So the point I would 
like to emphasize here is that I think the aeroelastics must get 
into performance and performance engineers must begin to quote 
performance information based on the dynamic information. 




90 180 270 

AZIMUTH ANGLE, deg 

Slide 49. 



360 



Slide 50 is a typical plot of the frequency versus rotor speed, 4 per rev, 5 per rev, 6 per rev, the various lines, and the two 
modes, these are coupled modes as it turns out, flapwise and edgewise. Now, why do I show this simple plot? Because I don't think we 
really know how to measure the edgewise natural frequency of a rotor blade. And I'm still concerned that we don't know how to do 
that in spite of the advances we have made in other areas. And a typical example we had here, there's a dot right here and that dot is 
the rap test or shake test that we had statically. We expected the information flow along this line and what these dots show are the 
actual test data that we had on the whirl stand. It's actually worse than this. I should have redrawn the plot, because the departure 
from that line, it's very significant, the slope differences. The key point is that if you operate at this rpm, what do you believe? Do 
you believe the analysis, or do you believe the test data, or do you extrapolate this information on up? Now what happened was that 
we were not able to determine the edgewise frequency because I think we were exciting the control system to try to find the 



336 



frequency, or we were putting sheets below the rotor to try to provide some sort of excitation, 
and the problem is you don't know what the edgewise force is, so you can't normalize the 
edgewise displacement to the edgewise force to determine what the edgewise frequency really 
is, and you get a great deal of coupling from the flapwise mode, particularly with a high 
degree of twist. So if there is any deficiency in the fundamentals, I would say that we should 
concentrate on defining better methods of determining the rotating edgewise natural 
frequency. 

Now, going to the simpler analysis, Slide 5 1. We have developed a computer aided design 
program for rotor blades. It has the ability to select any different airfoil, all the airfoils are in 
the data bank, and it also allows you to vary the cg-ac offset, compute values of inertia, 
stiffness, etc., and it saves a great deal of time, particularly on the computation of statics. 

We have also incorporated in this, Slide 52, a menu to allow you to select the forward 
flight condition and actually perform a forward flight analysis of loads and dynamics based on 
a simplified Normal Modes program, a couple of flapwise modes, one edgewise mode, one 
torsion mode, constant inflow, but it's a great tool for the designer. Now what he selects here 
is the velocity and the air density and all the parameters that one needs and on the output side, 
is the trim condition and it shows that you do in fact need trim. 



i TEST POINTS 

6fi 5fl 




100 200 

ROTOR SPEED, rpm 

Slide 50. 



300 




"GEE 



■_•..-.. l .-.._ 



Slide 51. 

Then, of course, Slide 53, he is able to predict and see the 
stresses versus span, the blade motions versus azimuth, or any 
element that is available in the batch computer process. The 
advantage here is that the designer is allowed a degree of freedom 
he never had before and that is — at least using simplified 
methods — to iterate the design around to something that is 
optimum for him, but having in there the constraints the dynam- 
icist has to place on him. I remember when we first got this 
program running, we had one of our most senior designers, and 
honestly he was really more of a most senior draftsman — 
because the designer of a rotor blade often is the guy sitting back 
in Aeromechanics — but in this particular case, we got this 
10 year designer in front of the tube, he got out the edgewise 
natural frequency, it turned out to be 3.9 per rev and he knew 
that wasn't acceptable. I asked him, "What do you do about it," 
and he didn't know, but we worked with him and we got the 



."Si' . - . ■ 






-• • HC k.56 eoi 



YT " J ^ffirM ITi 1V i iT ri W nff lir ^ c 1 T T^7r*i"i<iiriWf' *TrT'*^TIlHHiMlr i HW i T i ffiMTBflt J 

Slide 52. 




Slide 53. 



337 



mode shape plotted out on the tube, he was able to apply some masses and stiffness distributions and optimize the frequency, 
optimize the stress, and the cost, and the weight, and the reliability, and the rest of it. And this is what we are going to try to do 
more and more of, encourage our young engineers, and some of the older ones, to becomes designers and not draftsmen, by giving 
them some of these fundamental tools in the CRT. 

I'd like to make one final point on one subject we haven't covered at all, and that's vibration, n per rev vibration. I think the 
biggest challenge we face is the new challenge the government has given us which is ±.05 g's with n per rev and although I know this is 
maybe not the area we should be discussing — or maybe we should be discussing it — this is the area we are going to have to 
concentrate on, I feel. When I looked at the correlation of the various different methods I saw on the first panel this morning, I wasn't 
startled at the differences between the stresses and motions, but I was very much startled by the very significant differences in the 
rotor loads, particularly at frequencies of 4, 5, and 6 per rev. Jan Drees if you are out there you know there has beena lot of criticism 
about two-bladed rotors but the one advantage is that certainly you could predict the loads and you look at the 4, 5, and 6 per rev 
that we have here and you look at what our dynamicists are doing in terms of looking at the proper phasing and the modes and the 
contributions and it looks like a waste of time when you look at the variations that we have here. I think it is a very, very serious 
problem and not one that is going to be completely solved by vibration control devices and I look forward to maybe having some sort 
of a session some day on just that subject. 

Finally, in concluding my remarks, I would say that from a rotor designer's point of view, or any designer's point of view, he 
needs all the analysis he can get. He needs the sophisticated and the simple, and at face value you might say that that's very expensive, 
but when you consider the impact of that decision and the investment that you are making in a helicopter, it's a very, very small 
investment and one that's worthwhile, I think. That concludes our formal discussion, and we are looking forward to some stimulating 
questions and comments from the audience. 

QUESTIONS AND ANSWERS 
Panel 3: Integrating Dynamic Analysis and Helicopter Design 

Bob Wagner, Hughes Helicopters: I would like to make a comment that I think arises mostly from Rod Balke's presentation relative to 
the lack of time, lack of information, and the ulcers that arise during the proposal. And I would propose that there not be any 
dynamics analysis present in a proposal. Since the information on which to base a firm analytical solution is agreed is not available at 
that time, it would appear then that we are producing a whole bunch of fiction to be read by other people who attempt meaningfully 
to assess the same. Rather, I believe, that if we were told in a proposal, "Tell me how you are going to analyze these structures, first 
give me a delineation and an expression how you are going to do it, what methodology, and second, how are you going to handle the 
problems when they arise in flight tests?" Those two questions, I think, if we were to spend 100 pages each in describing those to the 
government or to the customer, rather than the 100 pages of analytical printouts, etc. that we now supply, I think that we would be 
better off. 

Balke: I'd like to comment on that just briefly. Of course, as you know, the requirement comes from the military for analysis in the 
proposal, however, my experience has been that you do need some lead time even with crude parameters to begin to get a feel for 
what ballpark you are in, particularly when you are working with designs that deviate from previous designs. And secondly, I agree 
with you, I think maybe a lot of effort should be diverted into a better follow-through in correlating shake tests with the analysis, 
updating the analysis, to give you a good handle on any future work that needs to be done. 

Glid Doman, Boeing Vertol: My question ties right in with Bob's. I wonder if the panel or the room feels that the customer should 
allow more time during design competitions, or perhaps that we need to induce him to encourage a different philosophy. Would we 
really do better with more time? 

Gabel: I'd like to comment on both of them. I think a lot depends on the energy and perhaps the defensiveness of the contractor on 
certain points during a proposal. In our UTTAS preparation, the big issue of course was the stability of the hingeless rotor, and we put 
out a massive effort during that proposal in which we did a huge analytical job, plus a huge model job which went into the proposal, 
and hopefully we convinced the government evaluators that we had done a good job and the aircraft would be stable. Now, there is 
an example where probably we could almost stop at that point. We had done such a vast job of preparation and modeling on that issue 
in the proposal, that what has developed since then, in the actual design of the aircraft has really been sort of minor parameter 
variations beyond that. So, I don't believe that it's true that proposals have to be goodness and mercy, but rather they can be much 
more factual with our advanced technology of today. 

Doman: That was done despite a large lack of time. 



338 



QUESTIONS AND ANSWERS 

Session I — Rotor System Dynamics 

Chairman: Kurt H. Hohenemser, Washington University 
Co-chairman: James R. Neff, Hughes Helicopters 

Paper No. 1 

fflNGELESS ROTOR FREQUENCY RESPONSE WITH UNSTEADY INFLOW 

D. A. Peters 
USAAMRDL, Ames Directorate 

Kurt Hohenemser, Session Chairman: First of all, for a hingeless rotor, say, with one of the usual flapping frequencies of 1 .05 or 1 .1 
or something like that, you have three modes really to consider: You have the advancing mode, which is somewhere on the order of 
magnitude of frequency of 2.0 and you have the coning mode which is around 1.0 or 1.05 and then you have the regressing mode, 
which is on the order of 0.2 or something like that. Now, would it be correct to say that for the advancing mode, and for the coning 
mode, you should really forget about the inflow? While for the regressing mode, you can use, without too much error, the quasi-steady 
approximation? This is my first question. 

Peters: In general, you are correct about the advancing and regressing modes. From [Fig. 61 one can see that induced flow is not 
important for the advancing mode, while quasi-steady induced flow is adequate for the regressive mode (.2/rev) except at very low 
Cj/a in hover. The coning mode, however, which is near 1 per rev, is in the region where unsteady induced flow could be a significant 
factor in rotor response. 

Hohenemser: Well, my next question is the following: In these programs like Rexor and the C-8 1 , by Lockheed and Bell, actually 
time delays were used for the inflow. Now, did you compare the time delays that were used in the industry with the time delays that 
come out of the potential theory that you used? 

Peters: As far as the Lockheed program, we did not check the time delay they used. However these potential flow time delays, the 
one for thrust was used in an early report [ Ref . 8 ] where they did transient response - put in a step in collective pitch and watched how 
the thrust came up - and that correlated well with that data. The time constant for roll and pitch, the Kj, correlated well with the data 
that Professor Hohenemser had gotten at Washington University, but the potential flow theory is just for zero lift and I'm sure that once 
lift is on the rotor these values may change considerably. 

Hohenemser: So, in other words, you are saying for the higher advance ratio, you have not made a comparison, or rather the lags 
that have been put in into Rexor or C-8 1 , really have nothing to do at all with your potential lag, is that wha f you are saying? 

Peters: That's right. 

Hohenemser: Maybe we could take the opportunity here by somebody who is familiar with Rexor or C-8 1 to say something about the 
lag times that were actually put in there. Maybe somebody else made a comparison with what Mr. Peters had. 

Jan Drees, Bell Helicopter Co.: I can speak only for a development we had in C-81 which was in connection with a study on gust 
response and we felt we had to put in a model to represent the induced flow change due to thrust change when you fly into a gust. We 
took into account two effects, one was kind of looking at the large scale induced flow field assuming a sudden change in lift which was 
done progressively [and would be] felt in the wake and we made calculations for how the induced velocity would change with time, at 
various advance ratios. There you find that if you change the collective pitch or you have a sudden change in thrust, that of course the 
wake will initially not know all this so you work with the induced flow that pertains to the previous lift value. And then gradually it 
will change to the new value and if you think in terms of a collective pitch change, initially you would develop a lot more lift but that 
would die out after a little while. Now, that would lead to very large lift overshoots, which you actually can measure in say a hovering 
rotor. And I think NASA has done some work in this respect. But you have to take into account also the near wake effects if you have 
a sudden change in lift on the blade. You have a starting wake behind the blade which tries to reduce the lift, momentarily, so that for 
very high and sudden inputs, and high frequencies also, you're more into the non-steady blade aerodynamics and that will mitigate the 
lift overshoot. But we found more or less the same conclusion that the effects are most pronounced at the low advance ratios. 

Bob Ormiston, AMRDL, Ames Directorate: I just have one question, Dave: The method that you discussed is based on a fairly simple 
wake model, and there is a lot of work being done on unsteady aerodynamics and downwash theory in other areas of the industry, and 
I wonder if you could contrast these methods with the method you are using in respect to the differences in applications with these 
methods. 

Peters: Yes, most of the prescribed wake and free wake programs for the induced flow, as was mentioned before, have to do with high 
frequency response, and with the shed wake of the blade and the fine details of the induced flow [for] rotor airloads, where this 

339 



model is just concerned with the low frequency response and just the gross effects. But these programs could be used to develop 
the L matrix. Remember the L matrix relates induced flow to the thrust and roll moment. These programs could conceivably be 
used to determine the nine elements in that matrix, to determine a better induced flow model as a function of thrust and advance 
ratio. 

Dick Bennett, Bell Helicopter Co.: Would the techniques that you have used also be applicable to an articulated rotor or is there 
something special that makes it only relevant to a hingeless rotor? 

Peters: Of course the methods would be applicable to an articulated rotor, but of course it's not as important a problem because the 
articulated rotor does not develop as large a moment so therefore the effect of induced flow would in general be smaller. But if you 
have a sizeable offset and do develop moments, then it could be used, it is not restricted to hingeless rotors. 

Paper No. 2 

DYNAMIC STALL MODELING AND CORRELATION WITH EXPERIMENTAL DATA ON AIRFOILS AND ROTORS 

R. G. Carlson and R. H. Blackwell 
Sikorsky Aircraft 

G. L. Comrnerford 
United Aircraft Research Laboratories 

P. H. Mirick 
USAAMRDL, Eustis Directorate 

Ray Piziali, Vizex, Inc.: Could you comment on the apparent, at least it appeared to be an apparent contradiction between your two 
dimensional comparisons and those you obtained on the rotor. If I interpret it right, your comparisons with the two dimensional tests 
indicated less aerodynamic damping from your predicted technique. The amplitude of the response appeared to be larger than the test 
data, whereas it appeared to be just the opposite in the full-scale rotor correlations. 

Carlson: Are you speaking of our tests of the oscillating model? 

Piziali: Yes, the airfoil. 

Carlson: I wish I was able to answer that, to be perfectly honest. I don't really have the answer. It could have to do with the property 
of the airfoil model and of the fact that it isn't actually a two-dimensional model but you know has end effects. 

Piziali: It could be in the three-dimensional effect? 

Carlson: Yes. Possibly that's it. 

Kurt Hohenemser, Session Chairman: May I ask you a question. I didn't quite understand, could you explain once more what is the 
basis of the time unit in your time delay method? I have the impression that you really should use a different unit if you go along the 
blade span. Is it based on the rotational time or on what is it based, that unit of 6? [r*, Fig. 1 ] 

Carlson: It is based on the local inflow over each rotor element, each element of the blade along the radius. In other words, it's the Six 
— where r is the radius to the element — plus the local airspeed over the blade element. 

Hohenemser: I see. So actually in real time, the delays are different along the blade? 

Carlson: They are. Well this is probably a problem that it has, actually there isn't enough interaction between the elements on the 
blade. And each one is stalling on its own at a different instant. 

Hohenemser: Also I noticed you had one graph without any scales on it for damping. Was that left out purposely? 

Carlson: No, it's in the paper, I think. Actually our visual aids were rushed at the end here and it was left out. 

Hohenemser: It might be interesting to show this once more and explain to us what the scales actually are. 

Carlson: O.K., that's [Fig. 3] — The bottom of the trough here curves at a mean incidence angle of approximately 20° . The damping 
begins at drop off in the area of 12°. [Below that] I basically have potential flow, and the damping that's associated with it, and at 
higher angles it drops off. The magnitude of the damping — it's a non-dimensional term and I would have to refer to one of the earlier 
references for description of it. 

340 



John Ward, NASA Langley : Ray, how did you handle unsteady drag in these calculations? 

Carlson: Well, this isn't handled at all, actually. In all of this, all we have is the drag that you pick up from two-dimensional steady 
tests, that's all. Now, this is an area that ought to have more work and it's an area that I suspect will help us out in giving more 
accurate edgewise response that we have had. However, it is an area that isn't handled in here. 

Tarzanin, Boeing Vertol: In your time delay theory, you have a separated flow eg vs. a. You didn't talk about it here, but you do in 
the paper. How is that determined? 

Carlson: It's based in part on looking at the hysteresis loops of blades and in looking at a blade in terms of the lift that it would have 
if the flow was purposely fully separated and the magnitude we had here was on the order of sin 4 a. 

Tarzanin: So you didn't simply put a line through the bottom of the loops. 

Carlson: No. 

Tarzanin: Second question: Has the a, A, B method changed significantly from the previous reported results? 

Carlson: No, we have altered the means of scaling it and I left this out. We are using it on other airfoils, basically scaling the hysteresis 
loops around the two-dimensional characteristics of other airfoils. If there is any improvement in the method, it is in this area. 

Tarzanin: Has it been a Mach number correction? 

Carlson: It is in there, in a way, but — 

Tarzanin: The same method essentially as before, right, you just did a 1 kf 1 — M 2 correction? 

Carlson: Yes, that's all it has in there, right. 

Tarzanin: So in the theory, improvement is due to the downwash, essentially. 

Carlson: It is actually, yes, we have worked harder at modeling it and using as accurate a model of it as we are able to. 

Glid Doman, Boeing Vertol: I'm wondering if in the analysis of the flight test data — [the] interpretation of the flight test data push 
rod loads — you've looked into the possible interrelationship of the loads in the stalling blade to those in the adjacent pushrod, in terms 
of swashplate motion. The hypothesis being that the rotor might not have flutter at all, but you might be seeing the stall load in the 
stalling blade appearing in the adjacent push rod. 

Carlson: We have looked at it and all of our blades are acting comparably at equal azimuth angles. 

Doman: The swashplate is moving? 

Carlson: It is. Really, I haven't any indication that this is the cause of it. I suspect we could look into it more but — well there isn't 
any indication of it that I am aware of. 

Jan Drees, Bell Helicopter Co.: We are always struggling with the question: Do you have to include yawed flow effects, like Frank 
Harris showed that you should many years ago? In your correlation studies, are you using this? 

Carlson: No. It isn't in here. 

Drees: And why not? 

Carlson: Well, that's even harder to answer but we haven't put it in and it is interesting and I don't know how major an item it 
actually is. It's on our list of items to look at as a possible improvement area. 

Drees: Usually the effect is that it reduces the stall effects. It would make the correlation worse. 

Carlson: Well, that's possible. 

Marcel Kretz, Giravions Dorand: I would like to know to what extent the similarity conditions were respected during your tests? As 
far as the Reynolds number is concerned, for instance, and the reduced frequency. 

Carlson: Well, in the tests the Reynolds number isn't accurate, it isn't representative of the actual aircraft. However, the reduced 
frequencies - in the model that is torsionally sprung, the mass and inertia of the airfoil are all accurately scaled. However, it is a small 

341 



model and it is off in the Reynolds number that we had. Our scaling of the unsteady aerodynamics in each instance, is about the 
two-dimensional steady data of the actual model that I was using, and this is also accurate. 

Kretz: Second question, did you run into flutter during your tests? 

Carlson: Yes — well it depends on your meaning of the word. Is a high torsional oscillation at the natural frequency flutter? If it is, 
yes. There is a question here, is it stall flutter or is it a high amplitude response of the blade due to an impressed moment at an instant. 
I have a little of each acting, I think. I have a change of exitation as I enter stall and I also have a negative aerodynamic damping 
entering in to build up this oscillation afterwards. From that angle, I have a little of each and it's a little hard to split off which is which 
here. 

Kretz: Are those oscillations explained by your theory? 

Carlson: Yes, I am able to pick up the amplitude and the overall physical characteristics of this oscillation. From this angle it is 
explained, yes. 

Bill Wilson, Bell Helicopter Co.: In one of your earlier slides, probably the third or fourth slide, you showed in the time delay, that 
you had three different scales going from low frequency to high frequency, would you care to put some numbers on those frequencies, 
please? 

Carlson: Well those are actually made up in my head from looking at typical hysteresis loops. I'm actually hard pressed to put a 
number on any of them. 

Piziali: With regard to the apparent discrepancy between the two-dimensional and three-dimensional comparisons, you mentioned in 
your full-scale comparisons that the computations included the induced velocity variation. Was this induced velocity variation 
compatible with the solution at each instant of time or was it like a fixed inflow distribution, non-uniform ■, but fixed in time? The 
reason for the question is that it's possible, maybe, that since the blade does not stall over its entire span, you may be getting damping 
from the wake over the unstalled portions of the span in your analysis. 

Carlson: Well, it is handled in iteration, actually, and there really isn't the total instantaneous interaction there. I run the blade 
response program for the blade loadings and motion, then I use it to describe a model of the wake. This, in turn, describes the inflow 
over the rotor and then I run the blade response history over again, and if it isn't altered appreciably, then I hold that I have a 
reasonably accurate modeling of the inflow. If it is altered, then I iterate again on the modeling of wake and the inflow. And I have. 

Paper No. 3 

COMPUTER EXPERIMENTS ON PERIODIC SYSTEMS IDENTIFICATION USING ROTOR BLADE TRANSIENT 
FLAPPING-TORSION RESPONSES AT HIGH ADVANCE RATIO 

K. H. Hohenemser and D. A. Prelewicz 
Washington University 

Earl Hall, Systems Control, Inc.: Did you have any convergence problems using the extended Kalman filter? This is one of the major 
criticisms of this approach when it is applied to realistic data. 

Hohenemser: Well, I can tell you something else that is in the paper and that I didn't mention here. We first used a very simple 
equation, namely Mathieu's equation, of a periodic system which had much less periodicity than the system that we used later, and we 
tried to apply a nonlinear estimation where the parameters and the state variables were estimated at the same time, and that converged 
only if you really knew pretty well the values of the parameters. So if we really put relatively large errors in there, and if we were not 
very careful to match the initial covariances with the initial values of the estimates, that actually did diverge. But with this linear 
estimator, we didn't have any problems with divergence at all. Only after smoothing the data with a digital filter. Well the convergence, 
I guess, is always there because as soon as the covariance matrix components go to zero, you just don't get any change of data any 
more. So it converges, obviously, as soon as the P components are all zero. But, of course, it converges to a wrong number. 

Hall: Did you do any investigation on the confidence bounds of the parameter estimates you were getting? In particular, did you 
investigate what the effect of the pre-processing filters was on the confidence bounds of the parameter estimates? 

Hohenemser: No, we didn't do that at all. Of course, in order to do that, we really would have to have gone through a lot of analysis, 
I mean either do it theoretically, we didn't do that, or we could have done it experimentally by just running these things a number of 
times but we didn't do that. Well, we ran it maybe two or three times and they all looked about alike, like the one that we showed 
you, but that of course doesn't mean that we actually got confidence bounds. 

342 



Hall: In the paper you mentioned selecting elements of the information matrix or the interpretation of the information matrix 
here- P, and looking at elements of that and possibly figuring out problems related to identifiability . Have you considered alternately 
using the elements of the sensitivity function themselves, 3H/9a, in order to give an indication of those modes which are and are not 
excited? 

Hohenemser: No, we didn't do that here. Since the system is relatively simple, we have only these two modes and so you have to be 
sure that both of these modes are well excited, obviously, and that was the reason why we had these two gusts, first the gust up and 
then the gust down, in order to make sure that they are both well excited. We first tried it with only one gust and that didn't always 
give us good results, so you have to be sure that you have sufficient excitation in the system, in order to get a good result. That is quite 
clear. 

Professor Bryson, Stanford University: I think it is a very commendable beginning in using this type of technique for a complicated 
problem. I think that Earl Hall has indicated that the real problem is a lot harder and I think you are aware of that, too, that the 
modeling errors creep in and you have to make some estimates, using this type of theory, of some white noise or some colored noise 
that enters into the process equations themselves. And if you then use accelerations as your measurements there's a further complica- 
tion, namely that your measurement noise is correlated with the process noise because the accelerations contain them. This has been 
handled too but it's still a long way, it seems to me, from these beginnings to a system that will work on measured real data. 

Hohenemser: Well, we'll see. We are presently running wind tunnel tests and want to try it out in the real world. But, of course, we 
tremendously exaggerated the measurement noise here. Actually our wind tunnel measurements are much more accurate than what we 
assumed here, so I would hope that the modeling errors that we will obviously have in the wind tunnel tests, will compensate, 
somehow, the tremendously exaggerated measuring errors that we have assumed in our computer studies. 

Paper No. 4 

DYNAMIC ANALYSIS OF MULTI-DEGREE-OF-FREEDOM SYSTEMS USING PHASING MATRICES 

R. L. Bielawa 
United Aircraft Research Laboratories 

A. Z. Lemnios, Kaman: In the model that you investigated with the aft chordwise eg position, did you assume coincidence between 
the aerodynamic center, the elastic axis, and the quarter chord? 

Bielawa: I assumed that the aerodynamic center was at the quarter chord. 

Lemnios: Therefore, you had a detrimental cg-ac offset effectively? 

Bielawa: Yes, I was introducing it purposely because I knew a priori that this would demonstrate the instability very well. I might add 
that the equations that I used are intended only to demonstrate the mechanisms of the matrices and are not intended to break new 
ground in that area. 

Lemnios: Following this, one more question. You therefore had an aerodynamic spring effect basically from the torsion mode being 
fed into your bending or flapping modes. Did you have a corresponding — I don't see how you could have — but did you have a 
corresponding term of an aerodynamic stiffness due to moment in the torsion mode as well? 

Bielawa: No. You do get a centrifugal moment, of course. 

Dick Bennett, Bell Helicopter Co.: I was wondering about the applicability of the method to forward flight and second, have you done 
any work in adapting this method to understanding what you get out from a Floquet type of analysis? 

Bielawa: Well the answer to the second one is simpler, no, I haven't. I would answer the first question by saying the method is not 
geared to any kind of a rotor eigensolution problem. It's geared to a more general class. If you can describe the equations of motion as 
a series of constant coefficient matrices, that is all you really need. 

Hohenemser, Session Chairman: Well I think Dr. Bielawa really deserves congratulations for a very simple and very effective means of 
making the complicated relations between the various dynamic variables visible and I have the feeling that this method will really lead 
to very large progress in really understanding these complex connections that we have in instability problems. Thank you, Dr. Bielawa. 



343 



Paper No. 5 

SOME APPROXIMATIONS TO THE FLAPPING STABILITY OF HELICOPTER ROTORS 

/. C. Biggers 
NASA/Ames Research Center 

Peretz Friedmann, UCLA: I have one comment which I would like to make, I don't really see in what way your conclusions differ 
from the work which is given in the various papers by Professor Hohenemser because if I can recall your conclusions correctly, 
conclusion number one said that fixed coordinates give more insight and all I can say about this conclusion is that insight is something 
which varies from person to person and some people can get insight also using Floquet theory. That's not really that important. And 
the second conclusion is that for low advance ratios you can consider the equations to be constant coefficient equations which is 
essentially mentioned a few times in Professor Hohenemser's papers, and the third conclusion is that at high advance ratios you have to 
go to more complicated methods which is again a comment made in Professor Hohenemser's papers. So what really is different from 
what you show? 

Biggers: O.K. Professor Hohenemser's work mentioned the constant coefficient, kind of as an aside — incidentally: Here it is. But I 
wanted to take a little harder look at it. I found it a little bit surprising that it would work on the articulated type of rotor which 
Professor Hohenemser had not mentioned in his work. I was not terribly surprised that it worked well on the hingeless [rotor] but you 
notice that with the articulated rotor, there is a very large frequency and damping change. Whereas, with the hingeless rotor, I had to 
greatly expand the scales on these plots to show the effects clearly - so the changes in frequency and damping on the hingeless 
machine are not very high. But with the articulated rotor, you get some very significant changes and the constant coefficient 
approximation for the low frequency modes, does a surprisingly good job. 

David Peters, AMRDL, Ames Directorate: I think it's not correct to say the hingeless rotor doesn't show the effects and the 
articulated does. It depends on the Lock number and P. And for any P value you can find a Lock number where these effects are big 
and other Lock numbers where the effect is small. 

Biggers: That's true. 

Peters: The same with the articulated rotor as you found, at a Lock number of 6 no periodic effects, hardly at all, at the low advance 
ratios. But [if] you get exactly the right Lock number then you can get the effects. 

Biggers: Right. This is more clear from some of the other figures that are in the paper. The yn plane, I think, shows it pretty clearly. 

Earl Hall, Systems Control, Inc.: As you know, I've been associated with Floquet analysis for some time. I guess I've had a question 
thrown at me many times. There's been a lot of mathematical work done on the only method available for solving linear differential 
equations with periodically varying coefficients. But I think particularly with respect to the frequency entrainment which is associated 
with half of the basic cyclic variation, there is still in my mind and I think perhaps for many of the people in this room [a question] : 
What experimental verification do we have that such a frequency entrainment exists and how does it manifest itself in an actual 
helicopter? Is there anybody who has some information along those lines? I mean, the thing is there, whatever it is, it's there, but what 
does it look like? 

Maurice Young, University of Delaware: I would simply like to suggest that since the constant coefficient method works so well at the 
very low advance ratios, why not use it once again with a simple perturbation in advance ratio? In effect, you have the low, or 
zero n solution, you perturb it once in the first power of advance ratio, and I would think you would have an excellent approxima- 
tion without the complications of the periodic coefficient differential equations. 

Biggers: O.K. that has been done by Wayne Johnson in one of the references in my paper, and that work is very useful. So that 
technique is also available. The other thing about the equations that I discussed, is whether you use it in the constant coefficient or 
with the periodic coefficient, the non-rotating degrees of freedom do include all blades. 

Young: What benefit, if any, in this type of approximation, could be gained by reducing the differential equations to a standard fornix 
eliminating the middle term, or the damping term, by a proper transformation? 

Biggers: You mean, a state variable form? Oh, O.K. The Hill's equation form. I don't know, I didn't look at that. One thing is that — 
one point I want to make on this thing is that there are undoubtedly cases where the approximation will not work. There are other 
cases where it is a very good approximation. And the point is there are still going to be a lot of cases where all the standard techniques 
that have been developed through the years can be used. But you need to check them using Floquet theory or some similar device. 

Hall: If I might express a theoretical point of view, this question of transforming a periodic system to a constant system is treated 
extensively under the theory of Liapunov transformations. This says, that if you have a periodic system, there does exist a constant 
coefficient system which will represent the response of that. This Liapunov transformation by the way has the rather nice theoretical 

344 



property that it is almost impossible to find. However, it might be interesting to see whether or not approximations such as this can be 
interpreted in terms of Liapunov transformations and if that is the case, perhaps the degree of applicability of such transformations 
could be extended such that they would have more application to the interpretation of flight test results. Just a comment. Thank you. 

Hohenemser, Session Chairman: I might mention something with respect to Dr. Hall's question. If you go back to the Lockheed 
history, they very definitely had that one and a half per rev that comes in from the Floquet theory so it has been observed in 
hardware, it's not that it's just theoretical. 

Pete Arcidiacono, Sikorsky Aircraft: A more practical question: Have you exercised your analysis and made these comparisons for 
four, five, and six bladed rotors and come up with an estimate of what advanced ratio you feel the constant coefficient method would 
work? In a three-bladed rotor you show it's only like 2/10ths, 3/10ths advance ratio. 

Biggers: No, except for the coupling that might come in due to slightly different flow fields due to different numbers of blades, I 
think you can look at the 7-/* planes that are in the written version of the paper and get a fair idea of the places where you can expect 
trouble from periodicity. 

Arcidiacono : Did you look at the higher number of blade problem? 



ers: No, I looked at the 4-bladed and that is in the written version of the paper and as far as the constant coefficient approxima- 
tion is concerned, it looks just like this one [the 3-blade easel except you have another degree of freedom which is blades one andthree 
up and two and four down. You know, differential coiling kind of thing, and for that one, when you do the approximation all the 
periodicity goes away, it decouples from all the others and it looks just like the hover rotors so that was not much use. 

Arcidiacono: Thank you, I would like to compliment you on your paper. I think you've made a very real attempt to communicate the 
results of your work to the audience. 

Hohenemser: I agree with this. Let me say one more word on this. Maybe this should be in the form of a question: Isn't it correct that 
below the branch point in the complex plane, not when its on the real axis, but if you have a branch point in the complex plane, and 
you are, with your parameter, below that branch point, then the constant coefficient approximation seems to be quite good. But once 
you are beyond that branch point where you had actually a single complex root of the Floquet transition matrix, then you have to be 
careful. Sometimes the constant coefficient approximation might be useful but it may also not. Would you agree with this? 

Biggers: That's correct. If it turns out to be a case where it does turn the corner you have to be careful and examine it. 

Frank DuWaldt, Calspan Corp.: Kurt, I'd like to first of all congratulate Jim again on the paper. There was a study in which the 
expansion was made in advance ratio, and in which we tried to interpret some old flutter results that we had which appeared to us to 
support the belief that the integer and half integers are in fact experimentally identifiable in forward flight flutter cases. The work was 
reported at the Air Force Flight Dynamics Laboratory V/STOL Conference in Las Vegas in September of 1969. 

Biggers: Thank you. 

Hall: That was the work of Daughaday and DuWaldt at Cornell [Aeronautical Laboratory] . That was what he [DuWaldt] was referring 
to which was done on a model. 

Paper No. 6 

FLAP-LAG DYNAMICS OF HINGELESS HELICOPTER BLADES AT MODERATE AND HIGH ADVANCE RATIOS 

P. Friedmann and L. J. Silverthom 
University of California, Los Angeles 

David Peters, AMRDL, Ames Directorate: I have one comment and one question. The comment is on the radial flow term. I don't 
know if radial flow is the best term to call it. In every paper I have seen on flap-lag in forward flight, except I think an earlier one of 
your's and Pin Tong's, it's in there and it's not called radial flow. It has to do [with the fact] that the blade is not at ^ but it's at 
\ji + f because of the inplane motion, you must take that into account. The question is that there are two assumptions you made in 
the beginning in order to get what you call the gut feel and the one assumption was that you use the hover value of trim of |3 , and no 
cyclic pitch or no S or /3 C trim. And also the assumption that you used the hover induced flow rather than the forward flight 
induced flow. And from the work we've done, those kind of assumptions will actually give quantitatively inaccurate results, and I was 
wondering how limited you think your conclusions are, based on those two assumptions? 

Friedmann: As you realize, I am very honest in presenting my results, and that is why I showed you how important collective pitch is. 
And as you well know, if you are aware of my report which was funded by AMRDL-Ames when I was at MIT, I started out by putting 
in trim and then after that experience with the trim procedure, I came up with the conclusions that using trim gives you a kind of 
effect which has a tendency to mask the effects associated with: a) periodic coefficients, and b) what is the effect of including the lag 

345 



degree of freedom. In order to be even more honest, I have now a program which does it with the trim procedure and its quite 
significant and its very case dependent. It's difficult to generalize. 

Earl Hall, Systems Control, Inc.: With all due respect Professor Friedmann, many of the results that you presented were also in 
Reference 10, of your document. I would like to ask one question. I have also been working recently in a problem similar to what you 
have described here. It is for others to go into the details of that. But one recurring problem that we have faced, and I would like to 
know if you found a solution to it, is how to determine what frequency you actually have. In other words, Floquet theory has an 
ambiguity to it in frequency. Now there are analytical and experimental techniques for determining these. I'm just curious if you have 
attempted to address this problem in your work, or have done what we've done which is rely on some of the more expensive 
computational procedures to determine the frequency. 

Friedmann: O.k. I'll answer your questions starting with the first part. I am very well aware of your report and I have been aware of it 
for a long time. And, I should also be honest in saying that from your report, I did not get a satisfactory gut feeling and that's why I 
decided to look at this thing, too. And I think you would agree with me that when you worked on this problem, you were more 
control systems oriented than helicopter blade dynamics oriented, and that's why maybe some of the gut feeling was missing. And 
regarding your second question, my recipe which I don't consider to be a good one, or the best one, but it's the one which for me 
works, is that I start my computations at n equal to zero and I usually know what's going on at hover and that's how I physically 
identify the various modes and I simply trace them . And when they split, I still easily can keep track of them . That's the only method 
which as far as I know works. 

Hall: That appears right now to be the only method. You can do it by other methods but they can assume all the storage of a 
360 computer. 

Hohenemser, Session Chairman: Maybe I should mention in this respect, I believe that what he really did has not very much to do with 
periodic coefficients at all because that has been done before using these multi-blade coordinates that Mr. Biggers told us about, and 
when you use constant coefficients, you also find that instability. There's a lot of literature available from Lockheed and from the 
people at Boelkow and from Westland and so on where they determined the lag instability and now it is also known from the literature 
that the inflow has a tremendously large effect. Now, you used a very, very simple inflow model that has no non-uniformity in the 
longitudinal direction. Now it has been established in literature that the non-uniformity of the inflow has a large effect on the stability 
of the inplane motion. And also it has been established in the literature, that the elastic lag-flap-torsion coupling is one of the main 
parameters that enters into the stability problem and you have also neglected that so I think, really, you are still quite a bit away from 
really getting close to the actual blade stability, but nevertheless, the thing that you are doing is very valuable because it is certainly 
including more than many other people have done. 

Jan Drees, Bell Helicopter Co.: Since this radial flow is such an important parameter — I still don't understand what it is and I don't 
have a gut feeling for what you mean there — could you explain what it is? 

Friedmann: Let me tell you what the term is. If you want to think in terms of an articulated blade, it's simply a beta times zeta term 
times m 2 cos 2 ip term which arises in the aerodynamic term associated with a lag degree of freedom. 

Hohenemser: I think it's a term that most people have included. It's just a new name for it. 

Maurice Young, University of Delaware: Just one last point. I think in some of these individual blade dynamic studies, there's always 
a temptation to say that the blade is a cantilever, as you have done and many of us have done before. I'd like to suggest the concept of 
reduced mass, that is, not assuming that the fuselage is of infinite mass. In other words, a moderate refinement will produce the same 
sorts of changes that some of these other parameter variations produce. So, I'd suggest that when you have these infinite hub 
mass calculations, with a little bit more work you can re-interpret them in terms of a real hub mass. 

Bob Ormiston, AMRDL, Ames Directorate: I think we are really a very long way from understanding the true nature of the problem 
and qualitatively whether stability or instability exists in forward flight even for the flap-lag system without the torsion. I think in that 
sense that we have to look at these results and say we really don't know what's going on yet. I'd like to amplify a little bit what Dave 
Peters was remarking about. He was talking about some of his work. Using somewhat different assumptions on trim and some of the 
same basic assumptions, some grossly different stability behavior was observed. So I think there is some question about the interpreta- 
tion of these results and there's going to be a considerable amount of effort that has to be done to really resolve some of these. One 
more point is that in the paper it's mentioned that droop has a large effect on stability in hover and I don't understand the reason for 
this, it's quite contrary to some of our results. We've used several approximations for flap-lag stability in hover, and we definitely do 
not see a strong sensitivity to droop or, as we define it, preconing. 

Friedmann: The Lockheed people claim that droop has a very large effect. That's all I can tell you. 

Ormiston: I'd like to say that the Lockheed work, as far as I understand it, does include the torsion degree of freedom and that is 
probably the most important factor here because of the way it couples with the flap bending and lag bending and directly influences 
stability. So droop and precone are very important factors if you're considering the full flap-lag-torsion problem. But for flap-lag — I 
can't say anything about forward flight — but in hover, our results show it doesn't have an effect that's appreciable. 

346 



QUESTIONS AND ANSWERS 

Session II — Helicopter Vibration and Loads — Theory 

Chairman: E. Roberts Wood, Lockheed-California Company 
Co-chairman: G. Alvin Pierce, Georgia Institute of Technology 

Paper No. 7 

CORRELATION OF FINITE-ELEMENT STRUCTURAL DYNAMIC ANALYSIS WITH MEASURED FREE VIBRATION 
CHARACTERISTICS FOR A FULL-SCALE HELICOPTER FUSELAGE 

I. /. Kenigsberg and M. W. Dean 
Sikorsky Aircraft 

R. Malatino 
Naval Air Systems Command 

David Peters, AMRDL, Ames Directorate: You mention in the beginning, you used the philosophy of building up the fuselage from 
components. I take it you reduce out degrees of freedom and then add them together. Had you thought of calculating the mode shapes 
of the smaller components, and then using modal composition to get the overall mode shapes of the structure? 

Kenigsberg: Well, that's the method of substructuring. What I was referring to is the structural modeling techniques and the allocation 
(and the location) of dynamic degrees of freedom to each component, under the assumption that you had correlation. Then you 
added something like a sponson. Model that, represent your mass distribution, add that to your structure either as a complete addition 
or by substructural analysis, and compare the correlation. If now there's a degradation in correlation it has to be an error in the 
modeling technique for the last element added, because everything else is correct. But this can be applied whether you use substructur- 
ing or not. This was a test and correlation approach. 

Marcel Kretz, Giravions Dorand: The first question, did you make any attempt to evaluate the damping of your structure? 

Kenigsberg: No, we did not during the course of this shake test, but during other tests that we have run on other vehicles, we have 
found the percent critical damping on fuselage modes varies from about 3% to about 6%. 

Kretz: It's purely experimental? 

Kenigsberg: Based on experiment, that's correct. 

Kretz: The second question, what is the accuracy of your theoretical evaluation, in general? 

Kenigsberg: On the fuselage first vertical bending mode, first lateral bending mode, and transmission modes, we had an average error 
between predicted frequency and measured frequency of 3% to 4%. 

Bill Nettles, AMRDL, Eustis Directorate: Irwin, I don't know exactly how to paraphrase this but in looking at the calculation 
accuracy you've got with the system without the appendages [stripped vehicle] , you had a relatively complex system where the 
individual elements were of significance to the problem. Now did you in fact, when you added these other lump masses in given areas, 
actually diminish the significance of the smaller components, and in essence, create a simpler problem, thereby allowing a more 
accurate solution, if you will. Did you actually go from a very complex problem of small elements to a new superimposed problem 
which included the small elements but which became predominated by the large masses? 

Kenigsberg: Well the answer to your question is yes, we did by adding the ballast, attack a simpler problem. We did not change the 
structural representation, we added the ballast. What we were trying to analyze in the stripped configuration, was a shell structure with 
a limited number of dynamic degrees of freedom. We only applied 200 dynamic degrees of freedom and we were getting modes well 
over 2,000 cpm. We were trying to develop modeling techniques which could be applied to the helicopter. The helicopter has a large 
transmission, a concentrated mass. It has a large concentrated mass at the tail rotor, so we can safely say that the modeling techniques 
we've evolved as a result of this will apply when analyzing the helicopter, but if you went Out and tried to analyze the shell of a space 
vehicle, they would not apply. 

Nettles: My question is, I guess, do the two different configurations actually represent two different problems. One in which the 
ability to predict accurately is a good deal more difficult than the other case? 



347 



Kenigsberg: Yes, in the stripped vehicle we would have to go to a much higher number of dynamic degrees of freedom in order to get 
the level of correlation that we achieved. They are two different problems. 

Bob Donham, Lockheed California Co.: In your model analysis, you mention the essentially plate restraint that you placed on the 
forebody which you think affected the behavior of the pitch degree of freedom of the transmission. I think you might have said 
something about the mass effect, and I was thinking more of the elastic properties of the root restraint of the transmission. You might 
comment on that. 

Kenigsberg: The point I was trying to make is that from our experience, the transmission pitch mode is to a great extent controlled by 
the frames immediately below the transmission and the mass and inertia of the transmission. And if you completely ! strip a vehicle and 
go to a shell representation, and try to predict the mode shapes and the frequencies there, then the point I'm trying to make is that it's 
not fair to take a shell structure and make a beam out of it and try and get an accurate prediction of that mode. However, as we kept 
that same modeling assumption in the analysis, representing the full fuselage as a beam; when we added the fully assembled 
transmission, or the equivalent of a fully assembled transmission, in that case we got excellent correlation. 

Donham: Yes, I think that clarifies the question. I guess really the point that I understand you to make is that you have to get a 
sufficient distance from the transmission base to be able to get the elastic properties of that support. 

Kenigsberg: Maybe I should reemphasize that if you were analyzing a vehicle without a rear cargo ramp, then all you need consider is 
a detailed finite element model, for example, spanning about 8 bays about the transmission region. That would give you an accurate 
prediction of the fuselage mode and the transmission mode. 

Donham: You mention that you had considered the panels to be non-buckled. I noticed in some of the maneuvers that you've made 
with the aircraft, that the panels are heavily buckled. What's the relevance here to your prediction, or have you made some 
calculations to assess that effect? 

Kenigsberg: Well, vibration levels always change in maneuvers, but when we design the vehicle from a vibration standpoint, we're 
designing for the 1 g level flight condition which is basically an unbuckled state. 

Bob Wagner, Hughes Helicopters: Presumably, you had a tremendous amount of detailed information relative to section properties. 

Kenigsberg: That is correct. 

Wagner: Now, would you care to comment on the accuracy, that you might have in the course of preparing a proposal where you 
haven't got so much detail information. 

Kenigsberg: All right, I should have emphasized then and I didn't, and I'll do it right now, that I believe that in order to get this level 
of correlation you have to have accurate structural data. And in developing the input for these analyses we actually went through the 
drawings one by one and generated the frame properties, the stringer areas, the skin thicknesses, the moment of inertia of the frames, 
and input it to the analysis. I believe that during a proposal effort, you would not have that type of accuracy, and I think therefore, 
from a time standpoint, we all have to give consideration to integrating our structural design analyses, and our dynamic analyses, so 
that as rapidly as possible, the structural data base required for the dynamic analysis, could be updated. This is why we're going away 
from the FRAN analysis and beginning to use the NASTRAN analysis for dynamics because we're using that for stress as well, and if 
we can combine the two we can get an accurate structural data base in the most rapid fashion. But it does require an accurate 
representation of the structure. 

Paper No. 8 

COUPLED ROTOR/AIRFRAME PREDICTION METHODS 

/. A. Staley and J. J. Sciarra 
Boeing Vertol Company 

Dick Bennett, Bell Helicopter Co.: Are your externally applied air loads treated only in terms of harmonics of 1 per rev and 2 per rev? 

Staley: Yes, this analysis as indicated here would be performed one harmonic at a time, in other words we would only feed back one 
harmonic of hub motion into this analysis, and ideally we should consider effects of other harmonics, their effects on the vibratory 
hub loads. 

Vincent Piarulli, Pi-Rad, Inc.: In your optimization techniques, if you start looking at more than just getting the response to 4 per rev, 
what is your criteria going to be for optimizing? 



348 



Staley: Well, the more things you try to satisfy at one time, the more difficult the problem becomes. The approach that was basically 
mentioned here was an attempt to identify structural members which could be changed to move particular resonances either above or 
below a particular exciting frequency. Conceivably, modes that would be excited at say, 4 per rev on a 4-bladed helicopter would not 
necessarily be involved in response at 1 per rev so that if we could identify members which could be changed to modify the frequency 
of one mode without affecting the frequency of another mode then that would be the optimum condition. That's not always possible, 
but usually there are some members which affect a particular mode without having a great affect on another mode. At this point, I 
don't think the solution to this problem has been completely put into closed form, and people are using a more or less trial and error 
approach to this problem. 

Bill Nettles, AMRDL, Eustis Directorate: You said that there were something like 1849, 1 think, structural mass degrees of freedom. 
My question [is] , since numbers like this stick in people's minds long after the subtleties of your program: Do you in fact have a 
dependency in your so-called structural degrees of freedom, that you can somehow reduce down such that you are using the term 
degrees of freedom loosely, are they truly degrees of freedom? Do you really have 50 degrees of freedom to begin with? 

Staley: Well, I'm not certain I completely understand your problem. As pointed out in the earlier paper, in the 3-dimensional model 
there are certain constraints that you have to try to work against and these are usually defining the shape of frames in the model and 
things of this nature which require the large number of degrees of freedom in the original structural model, merely to get the stiffness 
representation due to effects of shape of the model. Once those are obtained, then reduction of many degrees of freedom can be 
accomplished without too much effect on the lower modes of the final system. Am I answering your question? I believe your question 
is do we need all these structural degrees of freedom if we are only going to use 200 of them in the final analysis? Is that what you are 
asking? 

Nettles: My question was, since you go through this reduction, do you in fact use 1849 degrees of freedom? 

Staley: Oh, yes, if we break down the matrix equations into a set which has — if we look at the modal representation with no forces 
on the right hand side, and we partition these equations into a set with no masses on the left hand side and a set corresponding to the 
masses, you can see that the only things that couple the equations are terms in the stiffness matrix, so that we can more or less express 
a final reduced set of equations in terms — in other words there is a relationship between the reduced degrees of freedom and the final 
retained degrees of freedom in the final set of equations and that allows us to eliminate some of these degrees of freedom from the 
matrix to bring it down to a size which we can handle in the eigenvalue solution. So all of them are actually used, nothing is lost in the 
structural representation, assuming that the mathematics can be performed accurately. 

Piarulli: There was one more question I had that I also had wanted to ask the previous speaker. In your simulation of the fuselage, or 
body of the aircraft, do you take advantage of any symmetries of the structure? 

Staley: Well, NASTRAN can take advantage of symmetry. I'm not certain exactly how far we have gone with this, but there are 
features within NASTRAN which can analyze a structure which is symmetrical. In other words, you would only have to provide 
structural mass input data for half of the structure say, if it were symmetrical, and the program would then fill in the remaining 
portion of the final stiffness matrix based on the assumption of symmetry. So those features are in NASTRAN I understand. I'm not 
certain that we've taken advantage of them or whether we have complete symmetry in our structures so that we can. 

Paper No. 9 

HELICOPTER GUST RESPONSE INCLUDING UNSTEADY AERODYNAMIC STALL EFFECTS 

P. J. Arcidiacono and R. R. Bergquist 
Sikorsky Aircraft 

W. T. Alexander, Jr. 
USAAMRDL, Eustis Directorate 

A. Z. Lemnios, Kaman Aerospace Corp.: Peter, let me congratulate you on a very well presented paper, I think it was a very 
interesting piece of work. I did have a question on your gust penetration work. Did you look at the combination of advance ratio and 
rotor rpm whereby you could actually get a blade going into and out of the gust or did you treat the gust penetration as a solid 
actuator disc which then penetrates the gust? 

Arcidiacono: Well, we looked at several advance ratios, the gust profiles were as you saw them. Most of the results were for the sine 
squared gusts, and there was a fixed wavelength on that gust. We looked at advance ratios up to .5 , so we in effect varied the rate of 
penetration. It was a four bladed rotor and those parameters define when a particular blade will penetrate what part of the gust. And 
[if] each blade was considered separately, the results showed that the blade would actually travel downstream faster than the gust 
front was proceeding across the disk, the blade would move out of the gust front. 



349 



Lemnios: In other words, the gust alleviation factor that you quoted, for 50% did include the effect of perhaps a blade moving in and 
out of the gust in any one revolution. 

Arcidiacono: Yes, it did. 

Jan Drees, Bell Helicopter Co.: In reading over your paper, the effect of dynamic staUwas mentioned as one of the important factors. 
In our study that parallels pretty much your study, we came up with similar conclusions. We were using the Wagner and ' 
Kussner functions on each blade in order to simulate the dynamic effects of sudden encounters of each blade. And of course this 
alleviated the load from the blades and on the helicopters. I wonder if you've looked at that to see if that compares with the dynamic 
effects that you incorporated in your model? 

Arcidiacono: As I understand the work that was done at Bell, the effects that you just described were included, but those are 
primarily potential flow effects and they tend to alleviate the buildup of loads in potential flow. It was my understanding that the 
dynamic stall effects, the non-potential flow effects, were not included in the Bell work. 

Drees: That is correct. 

Arcidiacono: And that works against you, it allows the rotor to build up a larger load factor and we had, in effect I think, both effects 
included in our analysis through the use of the a, A, B method. Your second question? 

Drees: This is not so much a question as a comment. You've just looked at single-rotor helicopters and came to the conclusion that 
the airframe motions were not all that important. In our study, we also included the effect of a tandem rotor and I think there we 
found a much bigger effect since during the penetration of the gust the front rotor will first experience the gust and the aircraft will 
have time to pitch up a little bit and we found an increase in load level because of that. Maybe somebody from Vertol could comment 
on this. And then we looked at alleviation factors and we found that 5 3 and a bobweight in the collective system could substantially 
reduce the effects of gust. 

Arcidiacono: I think those are two good points. We didn't consider the tandem at all and obviously it will take more time for the gust 
to encompass the entire aircraft. As far as the bobweight part of the problem is concerned, we did the entire thing controls fixed 
because we felt that was a more general, and perhaps easier problem. 

Hohenemser, Washington University: My question really has been answered already, by what Jan Drees said but I wanted to ask you 
what really is the difference between your analysis and the analysis that they did there. I thought they were doing that with the first 
versions of C-81 and I wasn't aware that stall wasn't involved in it. I thought they had also stall in it. 

Arcidiacono: I pointed out, that as far as I understand, we have included unsteady stall aerodynamics in the problem. 

Hohenemser: That's the main difference? 

Arcidiacono: Yes, as far as I can see. Of course, as you see on Friday, there may be other differences that aren't so apparent. 

Dave Brandt, Boeing Vertol: Pete, to the best of your knowledge, has this analysis or any analysis that's similar to this, ever designed 
so much as one rivet on any piece of flightworthy hardware in the helicopter industry? 

Arcidiacono: To the best of my knowledge, I think the answer is no. And that's why I pointed out that, however, for commercial 
applications the ride comfort part of the problem still is of interest. 

Brandt: Then you need a control system in the SAS. Well, may I just make the comment that before we fall into the trap that our 
fixed-wing brothers have, let us dispense with this topic before somebody drives that idiotic power spectral density jazz down our 
throats, there are better ways to waste a lot of money than this topic. 

Peretz Friedmann, UCLA: I have a question of curiosity. You use a longitudinal gust, and for case for rotor blade which has rotational 
symmetry, it always bothered me — why not consider simultaneously a lateral gust and a longitudinal gust? Maybe that is a worse case 
than the one we have been looking at. What is your feeling about this? 

Arcidiacono: Perhaps you are right. Looking at those effects was beyond the scope of our study and concerning Dave Brandt's 
comment, I'm not sure that they should be looked at. 



350 



Paper No. 10 

APPLICATION OF ANTIRESONANCE THEORY TO HELICOPTERS 

F. D. Bartlett, Jr. and W. G. Flannelly 
Kaman Aerospace Corporation 

Irwin Kenigsberg, Sikorsky Aircraft: Is it absolutely necessary for the anti-resonant isolator to have both an inertial and elastic element 
or can you just have an elastic element? 

Bartlett: You will have to have both inertial and elastic elements because it is from the stiffness type element and the mass type 
element that the frequency is defined. 

Kenigsberg: Could the inertial element be the transmission itself? If the elastic element were mounted under the transmission? 

Bartlett: Well, the anti-resonant isolator is designed such that it is independent of the, say, the non-isolated body. So that it would 
have its own internal degrees of freedom. 

Leo Kingston, Bell Helicopter Co.: I wonder, is there a significant mathematical difference between your description of the anti- 
resonant system and a shaker system which has a closed loop feedback? Or several shakers, of course, which may be placed in several 
places on the fuselage with a closed loop. 

Bartlett: I don't have an answer to that question at this time. 

Bob Wagner, Hughes Helicopters: I don't want to appear stupid certainly, but maybe I'm speaking for more of the members than for 
fewer of the members if I would ask if there is a simpler way of explaining the principles of the anti-resonance devices that we have 
just heard? 

Bartlett: Admittedly, the subject of anti-resonance theory is complicated. Since joining Kaman and studying anti-resonance theory, I 
have found that no matter how many times I read the subject, I learn something new. Now, as to whether or not we can simplify this 
theory, I see no reason to simplify it. We do not simplify conventional matrix iteration techniques when we solve for resonances and 
normal modes and this technique is simply an extension of that by which we solve for those points in the structure that have zero 
motion for forces along generalized coordinates which we are interested in. So, if you look at the DAVI system itself as an 
independent and decoupled dynamic system, the resulting equations are not as difficult to understand. However, the overall picture 
looking at the isolated structure and the non-isolated structure and taking all degrees of freedom into account, is more complicated, 
yes. 

Leonard Foster, Lord Kinematics: By way of further explanation, are you not talking similar to earlier papers on structural analysis 
using NASTRAN to tune the structure, in that you are placing critical stations at nodal points? 

Bartlett: We're not placing nodal points, we are forcing nodal points. 

Foster: Well, that's what I'm referring to. I'm trying to get at an explanation of actually — You have an association between what 
you're talking about and doing a structural analysis or a finite element analysis, in reducing the response at different stations. Am I off 
base there or is there a correlation between tuning the structure, and what you are talking about with anti-resonances? 

Bartlett: Well, we do tune the structure to have an anti-resonance using the vibration isolator. 

Foster: O.K. Well, that's the point I wanted to make. 

Bob Jones, Kaman Aerospace Corp.: I think to maybe clarify the issues a little bit — and I'm not really up on the theory even though I 
work with Bart - I sort of agree with the gentleman who just spoke before. It really is a structural dynamics problem. Bart was only 
using say, anti-resonance isolators or dynamic absorbers to illustrate the problem. But it is a structural dynamics problem from the 
aspects that eigenvalues for antiresonances can be obtained and by structural dynamics you could force a nodal point at a given point 
of the structure, such as Bart was saying, at the pilot seat. Therefore, in design, this capability of calculating anti-resonances and 
putting anti-resonance at a specific point, could be of major value. So it is a structural dynamics problem, and not really an 
anti-resonant vibration isolator or vibration absorber problem, per se. 

Al Pierce, Session Co-Chairman: If there are no other questions, I would like to comment, myself. And that is in the light of simplified 
explanations. Simplified explanation is only arrived at through understanding. And I think this paper deserves some congratulations as 
much as it has clarified many techniques of resonance and anti-resonance approaches in one organized piece of work. Thank you, Bart. 



351 



Paper No. 1 1 

THE EFFECT OF CYCLIC FEATHERING MOTIONS ON DYNAMIC ROTOR LOADS 

K. W. Harvey 
Bell Helicopter Company 

Dick Gabel, Boeing Vertol: Can you comment on what portion of the load increase is due to the aerodynamics? I gather that there are 
two effects here, the aerodynamics and the dynamic effect of this adverse tuning at 3 per rev. Well, I presume then that you did not 
have flapping aerodynamics from cycling, resulting from cycling, prior to this. Can you comment on what portion of that load increase 
is aerodynamic? 

Harvey: Yes. There was a single set of aerodynamic forces that were taken from C81 and of course the aerodynamic cyclic effect is 
treated in there. But our representation in that program, consists of an airfoil shell that oscillates around a non-feathering elastic blade 
structure. The collective pitch is represented and a built-in twist is represented for the elastic blade structure, but the cycling motion 
of the elastic structure is not. The airfoil shell does oscillate which is required to balance the air load moments to maintain steady-state 
flight or precess the rotor to some other position. But just a single set of air loads were used for both cases, with and without cyclic. 

Peretz Friedman, UCLA: I'm wondering why in the plots which you have of the natural frequency versus rpm, you have no numbers 
on the vertical coordinates, both in your slides and in the paper? 

Harvey: You can obtain those directly. The scale is labeled from zero to 300 rpm and then the 1, 2, 3, and 4 per rev lines are 
shown so that you can construct the values very readily. I didn't feel that lends anything to the significance of the placement. But the 
magnitudes can be obtained directly. 

Gene Hammond, AMRDL, Langley Directorate: In the data that you present, you present out-of-plane degrees of freedom and 
in-plane degrees of freedom, but nothing on the torsional degree of freedom. Is there a reason first of all for leaving out the torsion, and 
secondly, do you expect to see any effect on the torsional response from your cyclic feathering? 

Harvey: Are you talking about blade torsion degree of freedom? 

Hammond: Yes. 

Harvey: Well, I included strictly the rigid body degree of freedom in blade torsion mode. The reason that it's not in there is just 
because of the development cycle of the program. I intended that that would be one of the next items included. And yes, I do expect 
some effect when that is included, particularly if that elastic motion derives from say, a control deflection as opposed to a blade twist, 
because it does seem to be that the most important consideration is the geometric angle of the blade structure that's closest to the 
feathering bearing. That's the principal item affecting this effect. 

Marcel Kretz, Giravions Dorand: You showed on one of the slides that the feathering modified the bending moment in the blade. Can 
you give the configuration of the rotor and especially the advance ratio of the rotor for this case? 

Harvey: I'm not sure I got the question. The flight condition for this set of air loads was a King Cobra which is our upgraded Huey 
Cobra flying at 1 50 knots, and the rotor rpm was 300, so I think that's about .2 or .25 advance ratio, somewhere in there. 

Kretz: Can you give an idea of the influence of the advance ratio on the bending moment in the blade? 

Harvey: No, I cannot, I've just run that one case. Most of the study to date has been to try to gain some sort of decision as to whether 
or not this effect was significant and we have not run very many cases. We're not automated to the point that we can very easily, at 
this point. 

Dave Brandt, Boeing Vertol: I don't understand why the aerodynamic damping doesn't show up in your analysis so that you would 
predict your damped resonance at something like 3.1 per rev. What's missing in the analysis that doesn't permit that? 

Harvey: We have that represented in the analysis, but it's a question of the trade-off between computer time and so forth. We have 
certainly aeroelastic effects and aerodynamic damping in the present C8 1 family of programs. In this particular study, I excluded it so 
that it would not mask the structural damping effects, the apparent damping effects due to structural feathering. 



352 



Paper No. 12 

CONTROL LOAD ENVELOPE SHAPING BY LIVE TWIST 

F. J. Tarzanin, Jr. 
Boeing Vertol Company 

P. H. Mirick 
USAAMRDL, Eustis Directorate 

Kip Cheney, UARL: Paul [Mirick] , that was a good paper, I really enjoyed it. One question that I had that you didn't address was the 
frequency of response. I was busy counting bumps on your curve there. I know on some of the 5 and 7 per revs, it looked like it was 
responding at 5 and 7 per rev, but the 3 and 4 per rev blades weren't responding at 3 or 4 per rev. 

Tarzanin: If you look at all the wave forms, you can see for the lower Cj/a's evidence of the lower frequency spiking. But as you get 
deep into stall you tend td see a higher frequency appearing which is traceable to the second torsional mode. So you have participation 
of both the first torsional and second torsional mode in the very high Op/a conditions. 

Peter Arcidiacono, Sikorsky Aircraft: Frank [Tarzanin] , and Paul, you've varied the torsional stiffness and torsional inertia indepen- 
dently — it looked like — on one of your figures. It looked as though the reduction in the torsional inertia led to higher control loads, 
did I interpret that figure correctly? 

Tarzanin: Earlier we did some work where we varied pitch inertia, control system stiffness, and blade torsional stiffness, and 
combinations thereof, and we got a very gross trend which showed that the lower frequency blades reduced the loads and the high 
frequency blades had higher loads. But there was a great deal of scatter, especially down at the low frequency end. You're right here in 
that when we went to the low frequency end and looked at changing pitch inertia there were actually increases. 

Arcidiacono: I raise that point because we had a companion study to this and we found that — our analysis at least — predicted that 
reducing the torsional inertia reduced the torsional loads, and was a more dramatic effect than the torsional frequency. 

Mirick: On your analysis you also — when you looked at torsional frequency — you coupled it with control system stiffness, didn't 
you, Pete? 

Arcidiacono: We made independent changes and combined changes as well. We're going to get into it on Friday, the comparison of 
various analyses, and this is one opportunity to at least examine the effect, the qualitative effect, or prediction of the two analyses 
with regard to a particular design parameter change and I was interested in that one point. 

Tarzanin: We may not disagree. It depends I think on where you started from. Because if you start at the higher end, if you go to 
around 7 per rev and then reduce the inertia, I think you can reduce the load. 

Arcidiacono: I think we differ in some areas. 

Leo Kingston, Bell Helicopter Co.: I'd first like to congratulate you on an excellent paper. At least I enjoyed it very thoroughly. I 
wonder whether you can help me a little bit. In the beginning of the paper you differentiated between the effects that you could 
achieve by changing control stiffness or by changing torsional stiffness of the blade. You showed that you apparently got the best 
results by varying the blade stiffness. I wonder whether you could elaborate a little bit to give me the rationale for why this was the 
best trend. 

Tarzanin: I wish I knew. I'm looking into it now and I just fall back to some of the work we did earlier where, when we changed 
everything we were able to get loads to go up and down using all methods. However, why torsional stiffness reduces loads the most, I 
don't know. 

Ray Piziali, Vizex Inc.: I may have missed a point. For each of the different torsional stiffness blades that you examined — and made 
comparisons — were the rotor trim conditions of forces and rolling and pitching moments maintained? 

Tarzanin: As the study was first done, the cyclic pitch boundary was matched and thrust was matched. In other words, side force and 
propulsive force could vary. We went back at the extreme conditions, and retrimmed the rotors, and found that the trends didn't 
change but it significantly affected the power. Originally, when we left the cyclics the same, it started showing that the softer blades 
had a great advantage in power. But when we went back and retrimmed them, they all appeared to have the same power boundaries. 

Maurice Young, University of Delaware: I'm puzzled and a little concerned at the presentation of the results in absolute terms, rather 
than using reduced frequency. I have a concern that what looks good on this particular rotor with its particular load distribution and 
chord and so forth, might be ruinous on another rotor system. In effect, have you tried to sift these factors out in terms of that 
standard parameter? 

353 



Tarzanin: No, I haven't. 

Jim McCroskey, AMRDL Ames Directorate: I presume some of that problem of the rise in pitch link loads at the low natural 
frequency is due to compressibility phenomena, and hence would tend to go away if you reduced the rpm. But of course, that would 
make stall flutter worse on the retreating blade. And I wonder if you have looked any at the trade-off in that direction? 

Tarzanin: No, again I haven't looked at that. 

McCroskey: Do you have any feel for how that would go? 

Tarzanin: It seems reasonable, it's clearly the advancing blade problem which is associated with speed. You can see that it is very, very 
speed sensitive, so yes, it makes sense. 

McCroskey: Well, I think it has to be that, because I don't believe the pitching moment characteristics of a cambered airfoil flying 
upside down are that much different from one flying right side up, in low speed flow. So it must be the fact that when the cambered 
airfoil is flying at negative angle of attack, it has pretty bad shock wave characteristics on the lower surface. 

Tarzanin: Oh, you are saying it's just due to drag. 

McCroskey: No, it's not drag. It changes the center of pressure. 

Tarzanin: No, it's not the pitching moment. The pitching moment, per se, is restoring, it's goodness. 

McCroskey: But Paul said it was a nose down pitching moment, and I presume he means [that] nose down for the negative angle of 
attack is like a divergent moment. 

Tarzanin: The load results from the fact that on the advancing blade, the tip is lifting down. And you have a shape like this [blade tip 
deflected down] . Then you take your drag [moment] , and you bring that thing nose down again, well that nose down twist makes the 
lift more down again. Which again gives you more moment, so if your torsional stiffness is too low you just wind this thing right up. 
It's very similar to divergence, yes. 

Bill Nettles, AMRDL, Eustis Directorate: I think with regard to Jim's question, even if what he's saying is true, I don't think your 
analysis accounts for it in the aerodynamic data at the tip. So I don't think — well I may be wrong, but I don't believe it accounts — 
you remember all the bit about the aspect ratio effect and all that — so there's no major change in center of pressure as your data is 
now put together. 

Tarzanin: We've looked into aspect ratio effects at the tip and we do have tables that account for it and that doesn't seem to be a 
significant factor here. I think if we got to very high Mach numbers that would be true. 

Nettles: O.K. There's another question that I want to ask you. Your conclusion, and I'm not sure I necessarily agree, is that you can — 
using this rather complex mathematical model as a design tool, a useful design tool at this point - qualitatively predict something with 
regard to the stall flutter boundaries and the torsional characteristics. Do you think, that now in retrospect perhaps understanding the 
problem a little better, that you could prepare a simplified model which would accomplish the same particular objective? 

Tarzanin: Do you mean a non-computer model like some kind of parameter? 

Nettles: Something along the lines of the simplified models that we've discussed, that perhaps could be done by hand or maybe would 
only require minimal usage of the computer. 

Tarzanin: It certainly sounds feasible. I guess it is a matter of work and time to work something like that out. 



354 



QUESTIONS AND ANSWERS 

Session III — Rotor/Vehicle Dynamics 

Chairman: Peter J. Arcidiacono, Sikorsky Aircraft 
Co-Chairman: William E. Nettles, USAAMRDL, Eustis Directorate 

Paper No. 14 

ROTOR AEROELASTICITY COUPLED WITH HELICOPTER BODY MOTION 

Wen-Liu Miao 
Boeing Vertol Company 

H. B. Huber 
Messerschmitt-Bolkow-Blohm Gmbh 

Kurt Hohenemser, Washington University: I have two questions. First, what is really different in your present conclusions, from the 
ones you had in your various previous works? I have the superficial impression that everything is really more or less the same, you just 
have more data and have correlation with your new tests, but it seems that there are no surprises in it. But maybe I didn't quite catch 
it. And the second question is this was all based on the rotating system, you have very low damping ratios of 1% or so, I assume this is 
all based on the rotating reference system. Now those are the damping values without any structural damping, is that right, you have 
eliminated structural damping in your presentation? 

Mao: O.K. The second question first. The model blade has 1/2% inplane structural damping and all the data presented here have 1/2% 
inplane structural damping. And because of the construction of the model, we know the blade has less damping than the full-scale 
blade. We have tested the BO-105 full-scale blade, it has more damping, at least 1%. It has data scatter but can go as high 3%, so the 
data shown here is, in a way, not representative of a full-scale blade. Now, to answer your first question, as far as I remember, there are 
two papers [that were] presented on this. One is for very small scale model, of 28 inch diameter. Another is the BO-105 full-scale data 
and on the BO-105 the parameter variation is impossible. This one is 5-1/2 ft in diameter and I believe the data is more reliable. 
Something that is always of concern, say what is the cyclic [pitch] influence on the stability, do you really have to include cyclic 
pitch in the analysis of air resonance? But this test data seems to indicate that the air resonance mode is not really influenced by that. 
And we also varied the control system stiffness. We showed that we will have a favorable trend versus collective pitch, when the 
collective pitch increased to a certain amount, because there is blade elastic pitch-flap-lag coupling. Then at high collective, because of 
the induced drag, the trend turns again. Now the control system stiffness will minimize that, for the high control system stiffness. This 
has never been shown before. 

Hohenemser: But this is all based on the same analysis. You could have gotten all this data with what you did a year and two years 
ago, is that right? 

Miao: No, the analysis I am showing here, is mostly the Blama analytical model [Messerschmitt-Bolkow-Blohm] which Huber used to 
correlate with the BO-105 data. 

Hohenemser: Yes, but he used that in the presentation of the last program, too. 

Miao: Yes. 

Bob Wagner, Hughes Helicopters: Would you conclude that the modal damping ratios that you show are proportional to the 0.5% 
damping that you showed about the virtual lead lag hinge, is that a fair statement? In other words, does the modal damping ratio 
appear as a product of some other dampings of which the lead-lag damping would be a component? 

Miao: If you refer to the old Coleman's criteria — 

Wagner: That's right. 

Miao: - the product of damping from the oleo and from the blade lag damping - we really didn't look at it that way. But I believe 
that because the blade has a very powerful flap damping, in a way you can think of it as oleo damping, and the lag damping in here is 
more or less very minimized. I don't know what would happen if I had zero blade lag damping in there. But we did do the analysis 
theoretically by varying the blade inplane damping, I can go down to zero damping and the mode is still stable. 

Wagner: That would indicate then that the modal damping is not only a product of components, but a sum of say the flapwise 
damping perhaps. 

355 



Miao: Well, I think the way that Coleman reached the conclusion on the product of damping, he ignored all the aerodynamic effect in 
there. 

Wagner: That's right, he did. 

Miao: Right. But I think the air is important in there. You can never get zero damping. 

Bob Ormiston, AMRDL, Ames Directorate: I have two questions, Jerry. First, the trim condition in forward flight. Was the model in 
trim and propulsive force, or what were the conditions during those tests? 

Miao: Let's see, trim and propulsive force. 

Ormiston: In other words, it was matched at some drag coefficient for the air frame. 

Miao: Right. 

Ormiston: I have another general question about some of the past work of both yourself and Mr. Huber. If I'm correct in remember- 
ing, I think some of the past work for stability in hover, showed that in your results there is an instability in the lead-lag mode at high 
collective pitch. And in some of the work that Mr. Huber has done, this instability does not appear unless the effects of stall are 
included. I wonder if you can comment on this, and if I have stated it correctly, maybe resolve the point for me. 

Huber: This was a question already at Washington [AHS Forum] last year, and up to now we have no exact [explanation] of these 
differences. In our theory, which brings only the stall instability in the very high thrust region, this new instability reached at 
moderate g levels without stall, does not come out. We are looking for this effect — I cannot give you a clear answer to that. But it 
seems that this curve [versus] g, which Jerry [Miao] showed, compares better to test results. 

Ormiston: In other words there still is a question that could be resolved? 

Huber: Yes. 

Miao: But, I think in a way, actually, the data we show there versus very high collective pitch, very high, I mean, and it shows a 
downward trend. We have eliminated the stall effect in there — you still will have it and it is caused by induced drag. But I think 
maybe if you included stall this thing would happen earlier. Could be. 



Paper No. 1 5 

AN APPLICATION OF FLOQUET THEORY TO PREDICTION OF MECHANICAL INSTABILITY 

C. E. Hammond 
USAAMRDL, Langley Directorate 

Dick Bielawa, UARL: From your results, it sort of looks to me like what you have with one blade damper inoperative, that you have a 
one-bladed rotor and the rest of the blades comprising part of the pylon mass. Do you think it might work to apply the Coleman 
analysis to an equivalent system, where you now have a one-bladed rotor without a damper and the rest of the blades not responding 
and comprising part of the pylon mass? 

Hammond: Let me give you an answer to it and see if I can give you a satisfactory answer. The answer is — directly to your question, 
I'm not sure, but one thing that comes to mind is the fact that one of the criteria that Coleman gives for completely doing away with 
the mechanical instability is to have the lag frequency of the blade greater than 1 per rev. That criteria is applied to a blade which has 
no damper on it. If you put damping on it, that's a very conservative requirement. All right, so to treat the situation that you ask 
about, it would say that if I had my blade with no damping, if I can somehow get its frequency above 1 per rev, maybe I can eliminate 
the problem. I have some results that show that you can eliminate the problem before you get that high, so I'm not really sure how to 
answer your question other than to tell you that I have some results where I've changed the lag spring on the blade to try to get the 
frequency up above 1 per rev to see what effect it has. And the instability actually goes away before you get as high as a 1 per rev lag 
frequency. 

Bielawa: I don't think you understand my question. I don't expect an answer. I'm just sort of asking you to consider it, that you 
might get an approximation to the problem very simply by treating the whole system as one wherein you have a reduced mass rotor, 
where you only have a one bladed rotor mass. It does not have any damping so your mass ratio will be reduced from what it would be 
if you considered all of the rotor mass. 

356 



Hammond: You want me to run a one bladed rotor now? 

Bielawa: Well, it sort of looks like you might be able to get an approximation to the solution using the Coleman results for an 
equivalent system. I don't know whether it will work or not, but it is an idea that was suggested to me by the responses that you 
showed. 

Maurice Young, University of Delaware: As a follow-up on the previous question, it seems to me that the system in either case is a 
linear one and that the loss of one damper is, in effect, a perturbation in one of the system coefficients. So that the solution should 
form a continuum and if you use the perturbation technique alluded to during one of yesterday's sessions - if you have a stable 
system to begin with, the first perturbation should give you a direct estimate of the loss in damping for the system, due to the failure 
of one damper. 

Hammond: I think that that's possibly true. The thing with that approach is that you have to go through and set up the perturbation 
equations and do the perturbation analysis, whereas if you use the Floquet results everything comes out from the basic equations. 

Young: Well, perturbation analysis doesn't really require any additional analysis because the characteristic part of the equation is 
identical from perturbation to perturbation. In effect, the perturbation produces a forcing function for the correction term. So it's a 
relatively straightforward calculation. 

Hammond: Right. 

Vince Piarulli, Pi-Rad, Inc.: Gene, to go back to the question before this one, maybe you could look at that by taking a good look at 
the modal response that you are getting on the other blades. If they are all moving in phase then he's probably right that you could 
simulate it by one blade and lump the other blades in with the hub. But if not, if you are seeing that they are acting differently, then 
that won't be true. 

Hammond: Right. One of the things that I have not done, is to look at detail at what each blade is doing. Even though the time history 
doesn't really indicate instability, I think that it would show the phasing between the blades. 

Bart Kelley, Bell Helicopter Co.: Maybe I didn't get it through my thick skull as you were explaining, but the collective drag mode 
with all the motion in the blades — this was the symmetrical hub - would be torsional then on the mast, or on the shaft, right? And a 
torsional stiffness should be included in the analysis there. That's one part of the question, and the other part is, this might also 
modify the anisotropic hub case. I just wondered whether that consideration is included or whether you were assuming a "cast iron" 
rpm that absolutely cannot vary — an infinitely stiff shaft driving this thing. 

Hammond: It's a "cast iron" rpm. It's exactly the same as the Coleman and Feingold analysis, except that now I take one of the 
dampers off, and look at a non-isotropic rotor. I think the rpm will change and certainly the effect of a torsionally soft mast will have 
some effect on the problem. i 

Paper No. 16 
THEORY AND COMPARISON WITH TESTS OF TWO FULL-SCALE PROPROTORS 

W. Johnson 
USAAMRDL, Ames Directorate 

A. Lemnios, Kaman Aerospace Corp.: I may have missed the point, Wayne, but did you mention what the wing incidence was during 
these tests? 

Johnson: The wing incidence for the theoretical and most of the experimental results presented, was zero. In the tests and in the 
theory we have varied wing incidence up to about 20 degrees. It has relatively little effect on high inflow proprotor dynamics. The 
main influence of the wing aerodynamics is on the damping due to angle of attack change when you have vertical bending motion. 

Lemnios: I was just wondering if perhaps it did have a significant effect on the stabilizing characteristics at the resonant frequencies 
that you pointed out in there, in other words, at the low speeds zero incidence condition, you indicated that you had an instability 
occurring at the 50 kt condition on the Boeing rotor, I believe it was. 

Johnson: O.K. Well, that is simply a resonance with the vertical bending frequency and the only effect that it might have is as you 
change the orientation of the inplane motion with respect to the wing vertical bending and chordwise bending direction. You might 
change that a little bit. I don't believe that we investigated that experimentally as I recall. I would not expect it to have a significant 
effect over the range of angle of attacks involved. 



357 



Bob Donham, Lockheed California Co.: I think the paper is really a very fine piece of work and I think it's very interesting. I do have a 
question relative to the effect you observed in the powered flight versus the autorotation flight case. The question is, do you believe 
there is an effect of whether or not the thrust vector on the rotor is positive or negative, so that if we had used power and put a large 
negative thrust vector in the sense of a braking action, would that have produced stability? 

Johnson: The influence of the lift and drag on the blade in high inflow aerodynamics is relatively small compared to the lift changes 
due to angle of attack perturbations. In fact, that's one of the things I meant by the "rotor" aerodynamic model that I referred to. 
The comparison of the autorotation and powered cases that were shown in the curves you saw today in fact, neglected the thrust 
terms on the blade. In general, the conclusion is that the effect you mentioned is small for high inflow aerodynamics. On the other 
hand of course, for hover aerodynamics, that may be another matter, certainly. 

John McCloud, NASA Ames: Wayne, could you comment on the experimental accuracy of determining damping? Your last slide 
showed some pretty horrible scatter for a condition where we were trying to find out if we were stable at 192 knots in our wind 
tunnel. Is this related to the difficulty referred to earlier of getting information on stability from time histories? 

Johnson: It's not the same thing that was referred to in the last paper, because the distinction that was made in the last paper was in 
the behavior of constant coefficient and periodic coefficient systems. In the matter of interpreting the periodic coefficients. In this 
case, it is involved in the question of interpreting experimental transient traces for the wing motion. That is what is involved. The way 
these damping levels were measured in this case, was to excite the wing motion with an aerodynamic vane; we'd oscillate the vane on 
the wing tip until we got a sizeable wing motion then we would stop the vane and then watch the decay of the strain gauge [output] 
measuring the wing motion. Now at high speeds you have enough turbulence in the tunnel to excite these big proprotors quite a bit, 
and that together with the rather low damping level indicated for the air resonance behavior, made the interpretation of the decay rate 
of these transient curves rather difficult. Now there are techniques available to get better experimentally determined damping ratios 
out of either the transient motion or out of transfer functions. This simply indicates that it would be a good idea to look seriously at 
these. 

John Bloomer, Bloomer Flying Saucer Co.: I wanted to ask you about the possible application of your theory to a jet flap proprotor. 
Would it be, do you think, applicable to a jet flap proprotor of about the same dimensions? 

Johnson: The only thing that comes to mind is the important factor, the change in the lift on the blade section with angle of attack. 
That's the dominant factor in the rotor aerodynamics and I'm not really that familiar with jet flaps to know whether you could just 
get by with using a higher effective lift curve slope or maybe you have to do a little bit more detail there. Basically, I think the 
dynamics involved would probably remain the same. 

Paper No. 17 

EXPERIMENTAL AND ANALYTICAL STUDIES IN TILT-ROTOR AEROELASTICITY 

R. G. Kvaternik 
NASA/Langley Research Center 

Bob Thresher, Oregon State University: We have a particular interest at Oregon State in wind machines and both you and the last 
speaker operated your rotors as essentially windmills. I wondered if you would have any comments on potential problems, particularly 
at low inflow ratios which might be the regime of operation for a wind machine? 

Kvaternik: Very low inflow ratios? We, experimentally of course, have not done anything there and although I've done a lot of 
analytical studies, even in the windmilling case they've all been essentially restricted to the fairly large inflow ratios typical of airplane 
operation of the tilt rotor. 

Thresher: You don't have any intuitive feel then for potential problems in that area? 

Kvaternik: No, I wouldn't want to venture a guess. 

Jeff MacDonald, AMRDL, Eustis Directorate: I was wondering if you had an opportunity to make lateral control inputs on the model 
to determine if the rotor inplane motion coupled with the wing motion and provided any high frequency tendency for the pylon to 
couple in the model? 

Kvaternik: I haven't put any oscillatory inputs into the equations, at this point. 

MacDonald: Or to the model? 

Kvaternik: No, we didn't have that particular capability in the model at that time. 

358 



Paper No. 18 

COMPARISON OF FLIGHT DATA AND ANALYSIS FOR HINGELESS ROTOR REGRESSIVE 

INPLANE MODE STABILITY 

W. D. Anderson and J. F. Johnston 
Lockheed California Company 

Jerry Miao, Boeing Vertol: I find your paper very interesting because it is a stiff inplane rotor and is different from a soft inplane 
rotor. Now I have two questions. The first one is a definition of your regressive inplane mode. Is that a mode which is the rotor speed 
subtracted by the lag frequency? And the second question is, on the last slide you showed test data of the regressive mode damping 
and also of roll mode damping. How do you obtain those test data? Do you measure it from the roil trace and one from the lag trace? 

Anderson: Well the first question. The inplane frequency in the rotating coordinate system on a stiff inplane system is, of course, 
above rotational speed so if you looked at a regressive mode in the rotating system, you would subtract rotational speed from it and 
you'd still end up with a regressive mode in the stationary system and the frequency of that mode would be the inplane frequency less 
rotational speed. So if you have a 1.5 P inplane frequency, the frequency would be 0.5 P. The advancing mode would come out at 
2.5 P or the inplane frequency plus rotational speed. Does that answer the first question? 

Miao: Yes, it does. But the result we provide is also the regressive inplane mode. 

Anderson: O.K., right, on the soft inplane system, the frequency is below rotational speed of course, and in the rotating system it's a 
regressive inplane mode, whereas, when you go through the transformation of that mode into the stationary system, you'll find that it 
turns out to be advancing in stationary coordinates. I mean I hope you find that out. 

Miao: The second question is where do you get the damping from? From the last slide. [Fig. 25] 

Anderson: The experimental technique there is done by the pilot instituting usually a control reversal of the lateral stick. We call it a 
stick doublet and this may be an inch, inch and a half or two inches of lateral stick, at different frequencies. Usually we have him do it 
at one-half cps, one cps, and one and one half cps, and the pilot we've had flying, for that particular test anyway, seems to hit these 
things right on. From that we take the data and first do a hand analysis but then we use a process that we call a moving block analysis 
which tries to isolate the single frequency. I reported a little bit on that at the AHS Forum last year. Using Fourier transform 
techniques we try to isolate the specific frequencies and also the amplitude of those frequencies with time in order to determine the 
log decrement or damping of each mode. 

Miao: Well, my question is here, when you analyzed the damping from the roll mode, does it have a frequency in the fixed system? 
And when you analyze the regressive lag mode, the frequency is in a rotating system — from a different trace? 

Anderson: O.K., I understand what you mean. No, the damping in each case indicated has been resolved to the fixed coordinate 
system. For the roll damping, basically we look at roll rate to just get that; for the inplane however, we take the blade measurements 
in the rotating coordinate system and resolve or sum them together in such a way that we end up with a cyclic inplane bending 
moment by appropriate summation of the blades in the rotating coordinate system. If you sum blades 1 and 3, you will end up with — 
I'm sorry — if you difference blades 1 and 3 you will end up with a cyclic mode and likewise on blades 2 and 4. So using that data we 
have a cyclic which does include 1 per rev and 3 per rev and 5 per rev and so on but we do a Fourier analysis on that at different 
intervals of time and plot the amplitude of that versus time. 

Kurt Hohenemser, Washington University: This is just a comment I want to make, with respect to the definition of droop and sweep. 
These are not dimensionless quantities and I think it is really basically wrong to use them. There is a big pitfall in it. It should be really 
expressed as a percentage of the chord shift, say the eg shift, in percentage of chord. Isn't it true that in the XH-51, you had no 
problems at all with this system and I think you had 2-1/2° sweep or something, and in the AH-56 you did get into problems because 
the percentage chord shift was not the same, but rather the sweep was the same? I think one should try to really use non-dimensional 
quantities in the description of the systems. That is really my comment. 



Anderson: I think you have a very good point, Dr. Hohenemser. 



359 



Paper No. 19 

HUB MOMENT SPRINGS ON TWO-BLADED TEETERING ROTORS 

W. G. O. Sonneborn and J. G. Yen 
Bell Helicopter Company 

Jim Hay den, U.S. Army Aviation Systems Test Activity: Thanks for an excellent paper. It was my understanding that the original 
undershing feathering axis was put in to reduce Corioiis loads. Will not the additional underslinging bring back that problem? 

Sonneborn: The question of Corioiis loads in a two-bladed rotor is at least a difficult one. In our analysis, we did include the 
computation of Corioiis loads in using the assumption that the vector of rotation is lined up with the rotor mast; you will find one 
gentleman in the audience, Mr. Harvey, who gave a paper yesterday, who would violently disagree with this assumption, and assume 
because of the relative inertias between engine and the rotor and the relative softness of the mast, that 1 per rev Corioiis motions 
arising out of coning and flapping, do not apply to a two-bladed teetering rotor. I have in the paper an additional figure that shows 
how the chord loads are influenced by additional underslinging and hub restraint. The change in chord loads appear to be very small 
compared to what we see being induced by airloads. I would think that the small amounts of additional underslinging that are 
required to balance the moments about a point below the hub, are insignificant in changing the chord loads. 

Ken Amer, Hughes Helicopters: This is probably a related question. You said that when you flew the OH-58, although you did not 
change the underslinging, you did not get a significant increase in vibration because the blades were acting like a dynamic absorber by 
inplane bending. Was there a significant effect on blade chordwise stress in your flight test of the OH-58? 

Sonneborn: Well, the inplane bending was changed. If you look in the paper you can see some results that show that the inplane 
bending was changed. However, once airloads and the inplane bending due to the dynamic absorber response are considered together, 
there is a phase difference, and for small amounts of hub restraint the chord loads actually go down because they are out of phase with 
those induced by airloads. And again, if we compare the chordwise loads that are introduced by airloads at Vg level flight conditions, 
they are significantly higher than those required for the dynamic absorber action to balance this 23% additional control power. Again, 
the chord loads are changed, yes, but the change is insignificant, it does not produce a structural problem. 

Kip Cheney, United Aircraft Research Laboratories: Two questions. What effect did spring have on the fundamental flapping 
frequency and also did you measure root stresses, flatwise stresses, what were they? 

Sonneborn: In answer to the first question, I do not know exactly. However, the amount of restraint that we provided I would guess 
would be in the order of 1 .0 1 or maybe 1 .02 per rev, very small. And the flatwise stresses are increased, correct. 

Cheney: To what level? 

Sonneborn: I am not able to answer that right now, but if you go to a 50% increase in additional control power, they have to be 
considered in the design, definitely. 

Marcel Kretz, Giravions Dorand: I would like to ask a question concerning the possibility to build a rigid two-bladed rotor. Once you 
introduce a spring constraint, it's a question of degree, to what extent the two-bladed rotor becomes a rigid rotor. In your opinion, is 
it feasible to have a rigid rotor, in this case? 

Sonneborn: A completely rigid two-bladed rotor? 

Kretz: That's right. 

Sonneborn: That would be unstable on a soft pylon. That is a matter of degree. I gave a reference in my paper that might help you in 
that respect. In our hybrid computer studies we could go to very high spring stiffnesses and even in the model we could lock out the 
teetering hinge completely, the flexibility provided by the coning flexures was sufficient to prevent instability. So you can go rather 
far. It doesn't appear to be a practical consideration. 



360 



QUESTIONS AND ANSWERS 

Session IV — Helicopter Vibration and Loads — Applications 

Chairman: James J. O'Leary, Boeing Vertol Company 
Co-Chairman: William G. Flannelly, Kaman Aerospace Corporation 

Paper No. 21 

VERTICAL-PLANE PENDULUM ABSORBERS FOR MINIMIZING HELICOPTER VIBRATORY LOADS 

K. B. AmerandJ. R. Neff 
Hughes Helicopters 

Irwin Kenigsberg, Sikorsky Aircraft: In general, when applying a dynamic absorber to a system, the absorber will drive the attachment 
point to zero with the tuned frequency. Since the attachment point is outboard on the blade, it's not immediately evident that the 
root shear must go to zero. Could you explain further why the root shear goes to zero, and why you believe a more outboard station 
would be more effective? Question number two is, in your paper you discuss effective mass. Could you define that more completely? 

Amer: I don't believe I said that the root shear was literally driven to zero. I said it was reduced significantly. You're right that because 
the pendulum is not literally at the flapping hinge, it couldn't literally drive it to zero, I don't believe. So it does not drive it to zero, 
but it does reduce it theoretically by roughly two thirds — practically, by roughly a half. Your question about mass ratio, Jim, why 
don't you take it. 

Neff: Your question was about the effective mass? 

Kenigsberg: The mass ratios that are in the paper are not clearly defined. 

Neff: The mass ratio shown there is the relative proportion of the modal mass, just looking at it from a modal standpoint. 

Kenigsberg: Is that the ratio of the mass of the absorber to the generalized mass of the mode? There are two mass ratios given. One is a 
mass ratio and one is a modal mass ratio. 

Neff: The modal mass ratio is the portion of the modal mass contributed by the absorber. The other one is just the actual mass of the 
absorber and the actual mass of the blade. 

Kenigsberg: I see. Two other questions. The modes that you had were quite close to 3 and 5 per rev. In a hypothetical case if it were a 
three-bladed rotor system and you put the absorbers on at 2 per rev. and at 4 per rev with your modes at 2.7 and 4.8, would the 
absorber have been as effective, or do you believe it would have been more sensitive to tuning, or off tuning, from the principal 
harmonic? 

Amer: I would say the effect of the absorber would be less significant because you're not near a resonant frequency. I could also say 
that from practical experience, we also build a three-bladed rotor system on our 269 helicopter, and it is much less of a vibration 
problem. In other words, we did not have, to go to pendulum absorbers in there to achieve a smooth helicopter. So it takes a 
combination of both proximity to blade resonance as well as a specific number of blades to result in the vibration problem . 

Kenigsberg: One other question, if I may. What was the effect on your 3 per rev and 5 per rev flatwise stresses following the addition 
of the absorber? 

Neff: There was very little effect. There could have been a slight increase, it was negligible overall. 

Jim Staley, Boeing Vertol: I also have two questions. You mention the problem in the approach to hover. What was the experience 
with the pendulums in vibration in that condition, and also, what direct measurements do you have of the effectiveness in reducing the 
3 and 5 per rev components of the rotor blade root shears? 

Amer: You mean during the approach to hover? 

Staley: No, in all conditions. But you didn't cover the approach to hover condition. What was that experience, and then in general, 
what was your experience in the 3 and 5 per rev components of blade root shear reduction as far as measurements indicated? 

Amer: Well, answering the second question first, we never did measure root shears specifically. The only measurements that were made 
were blade stresses and cockpit responses. As far as the first question, during the approach to land condition, to my knowledge, we 
only had qualitative pilot opinion. 

361 



Staley: Did you have any measurement of pendulum amplitude during steady state and transient conditions? Pendulum flap 
amplitude? 

Amer: To this extent, as originally installed, we would occasionally - you notice the pendulums are below the blades — as originally 
installed, we would occasionally see evidence of contact between the pendulum and the underside of the blade. We then had to put 
some stops. So I would say 95+% of the time it's less than about 20° amplitude. Occasionally, during a severe maneuver, the 
pendulums will try to exceed 20° and will hit their stops. 

Glid Doman, Boeing Vertol: This is perhaps in the line of an historical note and on your comment Ken about the use of these 
pendulums being a first. It ties back to the question about effective pendulum driving its attach point to zero. In 1944, we flew these 
things on the R6, on the swashplate. A pendulum at the base of each pushrod. They were tuned to enforce a zero for 2 per rev and 
4 per rev. We elected to use two and they enforced absolute zero conditions at the base of the pitch link until we switched blades and 
put some of our worst blades upon the machine and the pendulum was overpowered and put out of tune. We added another half 
pound and we were back to utter enforcement of no motion. This was, of course, on a machine with no servos, no irreversibles, and it 
was a wonderful thing for the pilot. The advent of servos had made us forget this. 

Amer: Well, it sounds like an interesting way to get your control force vibration down to zero. What I was talking about was getting 
the overall helicopter vibration reduced. 

Paper No. 22 

THE EVALUATION OF A STALL-FLUTTER SPRING-DAMPER PUSHROD IN 
THE ROTATING CONTROL SYSTEM OF A CH-54B HELICOPTER 

W. E. Nettles 
USAAMRDL, Eustis Directorate 

W. F. Paul and D. O. Adams 
Sikorsky Aircraft 

A. Z. Lemnios, Kaman Aerospace Corp.: Bill [Nettles] , and also Bill Paul, and Dave Adams from Sikorsky, I think you have done very 
fine work on this damping in pitch, I find it quite interesting. I do have a three-part question, however. I was wondering, had you 
considered in your analysis what the sensitivity would be to mismatched pitch dampers, what would happen if one of the pitch 
dampers did fail, and in fact, did you measure the pitch damping rates of the dampers to see how well they were matched? 

Nettles: I'll answer your last question, first. We built something like nine dampers. We took the dampers which were best matched. 
Frankly, I don't recall the numbers, but they were relatively close. In the whirl test and in the flight test of the vehicle, we were able 
to get a rotor system which tracked reasonably well. We had a little trouble in the whirl test, because we didn't know at first how you 
track such a rotor, but we learned how to do it, and in flight test the rotor system was well tracked. With regard to a mismatch 
occurring in flight, we talked about it, but I'm not sure how you would analyze it. Certainly it would cause a problem But we don't 
believe it would cause a catastrophic loss of the aircraft. As far as the total loss of one, that's a good question. Now I will say this, in 
order to check out what would happen if we lost damping, in the whirl test we bled the dampers. We were flying on nothing but the 
spring. Never noticed a thing. 

Session Chairman O'Leary: Any other questions? Frank Harris. 

Nettles: Frank, you told me you wouldn't do it. 

Frank Harris, Boeing Vertol: I guess I should just say Hi, Bill — is that what you are saying? I think it's an excellent example of 
exploratory testing and I'm pleased to hear that you are going to gather more points with which to look at this problem because 
acquiring stall data in straight and level flight is fairly non-repetitive kind of situation, and then to also be tackling the problem in a 
banked maneuver makes it even more of a chance for data scatter, but I'm very encouraged by your results. There was one area that I 
thought was interesting. I have sort of calibrated eyeballs on the torsion waveform, and I noticed that in the advancing quadrant of the 
disc, the pitch link is shown to be in tension. Maybe my eyeballs are uncalibrated, is there a sign problem there or is the tension load 
on the advancing side of the disc consistent with that kind of a rotor system? 

Nettles: Let me ask Pete Arcidiacono what he thinks about that. The reason that I ask for Pete is that this becomes a critical point 
with regard to the accuracy of the analysis, more so than the significance of the effect on the damper itself. 

Pete Arcidiacono, Sikorsky Aircraft: I think it's reasonably consistent. It's an uncambered blade and usually we have found with our 
waveforms that there is a nose-down pitching moment that occurs as you go into stall on the retreating side, and as the average 
moment due to stall relieves itself, the blade tends to spring back and overshoot on the advancing side and then comes back and finally 
enters into stall again. It would be different with a cambered blade. 

362 



Nettles: I wouldn't hang my hat on the quantitative accuracy of this analytical data. I wouldn't hang my hat on the quantitative 
accuracy of any of the analytical data from this type of program. But we went through a great deal of soul searching, and it is at this 
point when you do the proper amount of soul searching and use the proper ingenuity — whether you're talking Normal Modes or for 
that matter, C-60 — that you can use it as a design tool. In this particular case I think it was quite adequate for accomplishing the 
purpose intended. 

Paper No. 23 

MULTICYCLIC JET-FLAP CONTROL FOR ALLEVIATION OF HELICOPTER BLADE STRESSES AND 

FUSELAGE VIBRATION 

/. L.. McCloud, III 
NASA I Ames Research Center 

Marcel Kretz 
Giravions Dorand 

Bob Wagner, Hughes Helicopters: On one of the last slides, the second from the end, I believe, where you showed the comparison 
between the blade stresses and the vibrations. Do you feel a little uneasiness when you see the blade stresses going down but the 
vertical vibration going up? Is there not some kind of implied contradiction there, John? 

McCloud: No, I don't think there's a contradiction there - it isn't the greatest, this isn't what I would like — I'd like them both to go 
down together. But in my experience with other systems, and I was going to ask this question of Bill Paul and Bill Nettles, when I've 
added dampers in the system, it has chased the loads someplace else, into the blades for example. That's all I can say — I don't find it 
contradictory. 

A. Z. Lemnios, Kaman Aerospace Corp.: In order to perhaps amplify somewhat on John [McCloud's] statement, we're in the process 
of completing a study of putting in some multi-cyclic flap control on our controllable twist rotor. And with the dual control system 
such as that, we do have an optimization problem whereby we can maintain trim with many, many combinations of collective and 
cyclic on both the inboard and onboard end. We have been putting in 2, 3, and 4 per rev on the servo flap and the blade and we can 
get reductions of the shears transmitted to the hub. We do get some slight increases in bending moments, perhaps a little bit of increase 
in horsepower, 1 or 2 percent increase in horsepower, and peak angles of attack stay relatively about the same. What we are talking 
about here in trying to maintain optimum contours of many parameters at the same time. Now, you are going to benefit some of these 
parameters, but you will also get some detriments in the other parameters. 

McCloud: I think the real reason, Mr. Wagner, is, that we really haven't made the load uniform all over the disc. To do this, we would 
need a system which would give us radial control that is, variation in a radial sense as well as azimuthally. And I think until we do that, 
we will not be able to get rid of all the vibration at one time, all of the bad things. 

Paper No. 24 

IDENTIFICATION OF STRUCTURAL PARAMETERS FROM HELICOPTER DYNAMIC TEST DATA 

N. Giansante and W. G. Flannelly 
Kaman Aerospace Corporation 

Gene Hammond, AMRDL, Langley Directorate: First of all, is the method you describe applicable to rotating structures. Secondly, 
have you ever tried it on such a structure? 

Giansante: No, I'm sure it's applicable to rotating structures but I think this is one of the main points of interest for most helicopters 
right now, is getting impedance data for the rotor, or mobility data for the rotor. 

Hammond: When you say you are checking the structure sinusoidally, is there any particular frequency you are using or are you 
sweeping the frequency? 

Giansante: We are sweeping. 

Irwin Kenigsburg, Sikorsky Aircraft: Having defined the mass matrix and the stiffness matrix from the test data, do you have any 
plans for determining the modifications required for the stiffness matrix to reflect changes in the structure during the vibration 
reduction program? 



363 



Giansante: Well, I think perhaps this might be accomplished from substructure analysis. 

Kenigsberg: You mean integrating analytical methods into the system identification technique? 

Giansante: I think the two would complement one another, yes. 

David Peters, AMRDL, Ames Directorate: I have two questions. First, what would you consider to be the difference in using white 
noise input rather than sweeping in sinusoidal excitations, using this method? Would it be feasible? 

Giansante: Well, we would have to define frequencies, we would have to define the natural frequencies of the system, first of all — at 
least approximate natural frequencies. This method can be implemented only using data proximate to the natural frequencies. So you 
have to force at a particular frequency of excitation which is very close to each of the natural frequencies. 

Peters: Of course, white noise contains all frequencies. 

Giansante: Well, we would be exciting all the modes one time then. 

Peters: Right, and then by frequency analyzing the output, you can find what the output is at that frequency. 

Giansante: I suppose that might be an application. 

Peters: The second question is, in one of the earlier slides, you mention an iteration technique of matrices in connection with the 
pseudo inverse, do you mean you're iterating on the pseudo inverse? 

Giansante: No. We're using the pseudo inverse as a vehicle for the iteration. 

Peretz Friedmann, UCLA: You mentioned when Gene Hammond asked you a question that you are sweeping and when Peters asked 
you a question you told him that you are exciting at the natural frequency. So, either/or, you can't have both. 

Giansante: We are sweeping from the point of view that we are sweeping to discrete frequencies and taking data at these discrete 
frequencies, which are in proximity to the natural frequencies. That's true, we are not sweeping over the complete frequencies. 

Friedmann: O.K., because I saw that on the first slide, you had S equal to the natural frequency and the exciting frequency, both 
statements. 

Giansante: The S matrix, as I defined it there, was only a function of the natural frequencies and forcing frequencies. 

Friedmann: So, do I understand correctly that if you want to find the natural frequencies of a structure, where you don't have the 
slightest idea where they are, you can't use this method. 

Giansante: That's right. 

Bill Wilson, Bell Helicopter Co.: We ran a correlation study about a year or so ago, on an AH-1 J, using NASTRAN and modeling 
NASTRAN with the shake test of the AH-1 J. We were able to do pretty well out to about 22, 23 Hz. The modes in between there and 
100 Hz as you have there on the model, are considerably more complex than we have yet defined, I mean we're still modeling to try to 
get higher in frequency. Do you feel that damping in these higher frequency modes, and you showed in your beam — a very undamped 
beam by the way and I don't think any fuselage structure is going to be that way — but do you foresee that this method will handle 
the complex damping terms that are going to arise, in the higher frequencies? 

Giansante: Yes, that should present no problem at all. As a matter of fact, the beam representation we used had approximately 5% 
structural damping. At each of the modes, we inserted that much, 5% structural damping. 

Paper No. 25 

ENGINE/AIRFRAME INTERFACE DYNAMICS EXPERIENCE 

C. A. Fredrickson 
Boeing Vertol Company 

John Bloomer, Bloomer Flying Saucer Company: How did you match the torque in the three engines? 

Fredrickson: We are working the problem in the fuel control unit and we feel we have a satisfactory solution. 

364 



Bob Donham, Lockheed California Company: I think the paper and the presentation were excellent. I have a question about how you 
balance the engine-to-combiner gearbox shafting in view of the fact that it is 6 in. in diameter and operating at 1 1 ,500 rpm. 

Fredrickson: We are having some problems with vibration but not necessarily in balancing the shafts. End fixity at the combining 
transmission contributed to the shaft vibration problem but balancing is not as great a problem. The shafts are now roughly only 25 in. 
to 30 in. long and therefore are operating further from the bending critical speed. 

Donham: I asked the question because we experienced a problem with a similar shaft on a test stand that was not successful. This shaft 
was about 6 in. in diameter, 40 in. long, and operated at 13,500 rpm. The problem was in balancing the thin walled tube. It was 
balanced by adjusting weight on the end fittings but this did not prevent vibrations due to dynamic unbalance caused by imperfection , 
in the tube wall and transient torque loads. We eventually went back to a small diameter precision ground shaft. This is feasible 
because it doesn't really have to be any larger than the torque meter shaft. 

Fredrickson: We are having some problems right now with an installation on one of the transmission test stands. It's a problem of end 
fixity. I was just wondering how many bending critical speeds did you go through with this new shaft installation? 

Donham: Our shaft, the new installation, you mean on the stand? 

Fredrickson: Yes. 

Donham: The final one was operating above the bending critical speed. And the problem we ran into originally was that we ended up 
with a first torsion mode between these two opposite ends, which had been shaved off to balance them, and the shaft, when it got to 
twisting, started oscillating and it just shook the devil out of everything. 

Fredrickson: Very interesting. 

Bob Wagner, Hughes Helicopters: Are the HLH engines soft mounted or not? 

Fredrickson: The plans for the prototype right now are for a hard mount installation. But we have been doing work to find out what 
type of mounts we should put in there, in case of any problems. In other words, we do have a back-up design including a de-tuned and 
isolated engine installation. But the plans right now are for a hard mounted installation. 

Paper No. 26 

HINGELESS ROTOR THEORY AND EXPERIMENT ON VIBRATION REDUCTION BY PERIODIC VARIATION 

OF CONVENTIONAL CONTROLS 

G. J. Sissingh and R. E. Donham 
Lockheed-California Company 

John Bloomer, Bloomer Flying Saucer Co.: Yes, I was wondering^ I understand there were several Lockheed rigid rotor helicopters 
that did come apart in the wind tunnel or experimentally. Is this theory going to help alleviate the problem? 

Donham: No, I think the areas of activity you are talking about are not at all related to this information. A$ a matter of fact, in a 
somewhat different vein however, I would like to point out that the magnitudes of these angles we are talking about, relative to the 
control harmonics, appear to be small. In an autopilot functional mode you are not committing very much authority to such systems, 
and I think in that sense, they look like they could be very, very attractive with minimum risk to the aircraft. 

Bloomer: You mean to say that this system that you are describing could potentially improve the stability of the rigid rotor that was 
used in that particular helicopter? 

Donham: No, our subject is not related to stability. We are really related to forced response here, for a system that is already a stable 
operating system. 

Sissingh: I would like to make some general comments about this subject and about the idea. We have in principle, our rotor, which 
gives us our vibration — put this rotor in a box. Out comes vibrations of all types, all the frequencies and so on. And we have six 
control inputs available, the sine 4# and cosine 4^ variations of collective and two cyclics. And we have the choice to reduce six 
quantities to zero. So you can select what you want. In our cases, we selected the vertical forces and the pitching and rolling moment 
as we describe them. It is not necessary that these vertical forces and pitching and rolling moments be made to go to zero at the same 
point of the helicopter, in principle. In general, you would do that. But in principle, you could reduce the vertical forces at point A to 
zero, the rolling moments at B, and for pitching moments at C. We did not calculate response data and all our investigations are based 
on experimental data. We use our frequency response data to calculate the input to obtain the required output. In this case, we try to 
reduce, or to counteract the existing vertical vibration, pitching moment, and rolling moment. 

365 



Rod Balke, Bell Helicopter Co.: I may have missed it, but what was the consideration of torsional stiffness of the model blades? And is 
this a factor? 

Donham: This particular model is extremely stiff in torsion. I can't give you a figure but I suggest that it's over 8 per rev. It's quite 
high. My comment was — well, it's a good comment - that the model behavior, I think, was not at all influenced by the torsional 
rigidity of the blades, so we were giving this blade angle to the full span of the rotor blade. 

Dick MacNeal, MacNeal Schwendler Corp.: I'd be interested in a comment regarding your opinion of the feasibility of an adaptive 
control system to perform these functions. 

Donham: Well, I'll be honest in telling you that I haven't made such a study. I think, to be quite frank, Dick, the sense of our work to 
date has been to establish this: if I knew how to steer it, would I be able to make it? I think John McCloud is shedding some light on it 
and I think he believes he could make it, and I think the majority of the people here believe that you can do it. I think though the 
basic issue is, can you do it and handle the varied conditions you encounter? I'm kind of optimistic that you could actually make an 
adaptive control system that would automatically relieve this moment, and to be honest again, the reason I believe that, Dick, is that 
the authority levels you have to have, I think, are sufficiently low that you can actually, I guess the word is, play with it. I think it can 
be done. 

Ray Rose, Honeywell Inc.: I understand from what Dr. Sissingh stated, that you can concentrate on a particular load at a frequency 
that you would like to reduce. I was wondering if you do concentrate on one particular load, is there some limit of control that you 
have to exercise to be sure that you don't cause another load to diverge at another frequency? 

Sissingh: Theoretically, you can do anything. What is practical, of course, I don't know. If the theoretical problem can be solved, you 
can concentrate on six quantities which reduce to zero, in theory. 

Donham: I guess I would add one remark to that. I can't imagine from a gut feeling — that's been used a lot but I think it is a good 
expression - how could it be the case that we could experience loads? I think most of us know full well, that if the rotor system is 
actually "exceeding its endurance limits in trimmed flight it is in a great deal of difficulty. So it seems rather surprising to me to imagine 
— unless we had a very unusual event here — that we could get into serious difficulty in that area. 



366 



LIST OF ATTENDEES 



David O. Adams 
Sikorsky Aircraft 

H. R. Alexander 
Boeing Vertol Company 

Kenneth B. Amer 
Hughes Helicopters 

William D. Anderson 
Lockheed California Company 

Peter J. Arcidiacono 
Sikorsky Aircraft 

Rodney W. Balke 

Bell Helicopter Company 

Jewel B. Barlow 
University of Maryland 

Lawrence D. Barrett 
Boeing Vertol Company 

Felton D. Bartlett, Jr. 
Kaman Aerospace Corp. 

William G. Basham 
U. S. Army 

Paolo Bellavita 

Costruzioni Aeronautiche Giovanni Agusta 

Richard L. Bennett 
Bell Helicopter Company 

Dale F. Benton 
Tenneco, Inc. 

Marvin S. Berger 

U. S. Army Aeronautical Depot Maintenance Center 

Richard L. Bielawa 

United Aircraft Research Laboratories 

James C. Biggers 
NASA-Ames Research Center 

B. J. Bird 

Bell Helicopter Company 

John H. Bloomer 
Bloomer Flying Saucer Co. 

H. Peter Borie 
Sikorsky Aircraft 

William G. Bousman 

Ames Directorate, USAAMRDL 



David E. Brandt 
Boeing Vertol Company 

Herbert H. Brimmer 

Lockheed Missiles and Space Company, Inc. 

James H. Brown, Jr. 
U. S. Army 

William R. Bryant 

U. S. Army Foreign Science and Technology Center 

Arthur E. Bryson 
Stanford University 

James A. Burke 
HQ, USAAMRDL 

Marvin W. Buss 

U. S. Army Aviation Systems Command 

Raymond G. Carlson 
Sikorsky Aircraft 

E. S. Carter 
Sikorsky Aircraft 

James R. Chadwick 
Chadwick-Helmuth Company, Inc. 

DavidP.Chappell 

U. S. Army Aviation Systems Command 

M. C. Cheney 

United Aircraft Research Laboratories 

Sing Chu 

NASA-Ames Research Center 

J.J. Costes 

Office National d'Etudes et de Recherches 
Aerospatiales 

Samuel J. Craig 
Systems Technology Inc. 

W. L. Cresap 

Bell Helicopter Company 

John P. Cress 



H. C. Curtiss 
Princeton University 

Roland Dat 

Office National d'Etudes et de Recherches 
Aerospatiales 



367 



Gary L. Davis 
U. S. Army 

John M. Davis 

Bell Helicopter Company 

Robert C. Ditrich 
Lord Kinematics 

Glidden S. Doman 
Boeing Vertol Company 

Robert E. Donham 
Lockheed California Company 

Jan M. Drees 

Bell Helicopter Company 

Frank DuWaldt 
Calspan Corporation 

David Victor Ellis 

Westland Helicopters Limited 

Robert F. Forsyth 

U. S. Army Aviation Systems Command 

Leonard W. Foster 
Lord Kinematics 



W. Earl Hall, Jr. 
Systems Control, Inc. 

C. E. Hammond 

Langley Directorate, USAAMRDL 

Franklin D. Harris 
Boeing Vertol Company 

Gary Hartwick 
Pi-Rad Inc. 

Keith W. Harvey 

Bell Helicopter Company 

James Hay den 

U. S. Army Aviation Systems Test Activity 

Robert E. Head 



David H. Hickey 

NASA— Ames Research Center 

Dewey H. Hodges 

Ames Directorate, USAAMRDL 

Kurt H. Hohenemser 
Washington University 



David J. Fowler 
U. S. Army 

Charles A. Fredrickson 
Boeing Vertol Company 

Juanita Frick 

Ames Directorate, USAAMRDL 

Peretz Friedmann 

University of California, Los Angeles 

Richard Gabel 

Boeing Vertol Company 

Troy M. Gaffey 

Bell Helicopter Company 

G. H. Gaonkar 

Southern Illinois University 

Nicholas Giansante 

Kaman Aerospace Corporation 

Donald P. Gleiter 

U. S. Naval Air Development Center 

Mike Gockel 

The MacNeal-Schwendler Corporation 

Jerry H. Grantham 
Lord Kinematics 



Helmut Huber 
Messerschmitt-Bolkow-Blohm GmbH 

Shukry K. Ibrahim 
Honeywell, Inc. 

F. H. Immen 
HQ USAAMRDL 

Nelson K. Itterly 

U. S. Army Aviation Systems Command 

David S. Jenney 
Sikorsky Aircraft 

Wayne Johnson 

Ames Directorate, USAAMRDL 

J. F. Johnston 

Lockheed California Company 

Robert Jones 

Kaman Aerospace Corporation 

Clarence T. Jones 
U. S. Army 

Daniel D. Kana 

Southwest Research Institute 

Gerd Kanning 

NASA— Ames Research Center 



368 



David H. Kaplan 

Bell Helicopter Company 

Neville F. K. Kefford 
Sikorsky Aircraft 

Bartram Kelley 

Bett Helicopter Company 

Mark W. Kelly 

NASA-Ames Research Center 

Irwin J. Kenigsberg 
Sikorsky Aircraft 

Paul Kesling 

Lockheed California Company 

David Key 

HQ, USAARMDL 

Edward J. King 
GTE Sylvania, Inc. 

Leo Kingston 

Bell Helicopter Company 

Douglas G. Kirkpatrick 

U. S. Naval Ship Research and Development Center 

Marcel Kretz 
Giravions Dorand 

Raymond Kvaternik 
NASA-LangJey Research Center 

Anton J. Landgrebe 

United Aircraft Research Laboratories 

A. Z. Lemnios 

Kaman Aerospace Corp. 

Richard B. Lewis 

U. S. Army Aviation Systems Test Activity 

Karl Luken 
U. S. Army 

H. I. (Jeff) MacDonald 

Eustis Directorate, USAAMRDL 

Richard H. MacNeal 

The MacNeal-Schwendler Corporation 

George Maggos 

Naval Air Systems Command 

Martin D. Maisel 

Ames Directorate, USAAMRDL 

Ray Malatino 

Naval Air Systems Command 



John L. McCloud III 
NASA— Ames Research Center 

Barnes W. McCormick 
Pennsylvania State University 

W. J. McCroskey 

Ames Directorate, USAAMRDL 

Joseph H. McGarvey 

Eustis Directorate, USAAMRDL 

Dennis P. McGuire 
Lord Kinematics 

Ronald B. Mcintosh 
U. S. Army 

Walt Mclntyre 



Wen-Liu Miao 

Boeing Vertol Company 

Paul Mirick 

Eustis Directorate, USAAMRDL 

Richard L. Mohr 
Systems Control, Inc. 

Gilbert G. Morehouse 

Ames Directorate, USAAMRDL 

H. Andrew Morse 

Ames Directorate, USAAMRDL 

David E. Mowbray 
British Embassy 

Mike Munguia 

U. S. Army Aeronautical Depot Maintenance Center 

A. W. Myers 

Bell Helicopter Company 

James R. Neff 
Hughes Helicopters 

William E. Nettles 

Eustis Directorate, USAAMRDL 

James J. O'Leary 
Boeing Vertol Company 

Robert A. Ormiston 

Ames Directorate, USAAMRDL 

William F. Paul 
Sikorsky Aircraft 

Robert E. Peskar 



369 



David A. Peters 

Ames Directorate, USAAMRDL 

Wayne Petrie 
U. S. Army 

J. J. Philippe 

Office National d'Etudes et de Recherches 
Aerospatiales 

Vincent J. Piarulli 
Pi-Rad Inc. 

G. Alvin Pierce 

Georgia Institute of Technology 

Raymond Piziali 
Vizex, Inc. 

M. F. Platzer 

U. S. Naval Postgraduate School 

Richard K. Pollock 
U. S. Army 

James A. Poore 

U. S. Army Hughes Plant Activity 

Curtis J. Poree Jr. 
U. S. Army 

Robert D. Powell, Jr. 

Eustis Directorate, USAAMRDL 

A. LeRoy Pucker 
Lord Kinematics 

John P. Rabbott, Jr. 

Ames Directorate, USAAMRDL 

B. M. Rao 

Texas A&M University 

Kenneth R. Reader 

U. S. Naval Ship Research and Development Center 

Gunther Reichert 
Messerschmitt-Bolkow-Blohm GmbH 

Barclay T. Resler 



Stephen A. Rinehart 
Pi-Rad, Inc. 

Raymond E. Rose 
Honeywell, Inc. 

Samuel W. Russell 



Warren E. Schmidt 
Lord Kinematics 

Fredric H. Schmitz 

Ames Directorate, USAAMRDL 

Daniel P. Schrage 
U. S. Army 

Robert G. Schwendler 

The MacNeal-Schwendler Corporation 

Michael P. Scully 

Massachusetts Institute of Technology 

David L. Sharpe 

Ames Directorate, USAAMRDL 

John L. Shipley 

Langley Directorate, USAAMRDL 

G. J. Sissingh 



Charles E. Smith 
Oregon State University 

Walter G. O. Sonneborn 
Bell Helicopter Company 

M. E. Southgate 
British Embassy 

Richard Spivey 

Bell Helicopter Company 

Francis E. Spring, Jr. 

U. S. Army Aviation Systems Command 

James A. Staley 
Boeing Vertol Company 

Irving C. Statler 

Ames Directorate, USAAMRDL 

Robert E. Steinwedel 
Lord Kinematics 

W. Z. Stepniewski 
Boeing Vertol Company 

K. G. Sunding 
Lord Kinematics 

Brian E. Swan 

NASA-Ames Research Center 

Peter Talbot 

Ames Directorate, USAAMRDL 



370 



Frank J. Tarzanin 
Boeing Vertol Company 

Robert W. Thresher 
Oregon State University 

Andrew T. Trenka 
Vizex, Inc. 

William J. Twomey 

Office National d'Etudeset de Recherches 
Aerospatiales 

George. P. U»ge.r HI 

Naval Air Systems Command 

V. V. Utgoff 

U. S. Naval Academy 

R. J. van der Harten 

KLM Nordzee Helikopters B.V. 

Henry R. Velkoff 
HQ USAAMRDL 

Robert A. Wagner 
Hughes Helicopters 

John F. Ward 

N ASA-Langley Research Center 

George A. Watts 

Lockheed California Company 

Lawrence E. Wessman 
U. S. Army 



Ricnard P. White, Jr. 

Rochester Applied Science Associates, Inc. 

Bradford H, Wick 

NASA- Ames Research Center 

Jostoh B. Wilkerson 

U. $. Naval) Ship Research and Development Center 

Robert M Williams 

U. S. Naval Ship Research and Development Center 

Robert Elliot Wilson 
Oregon State University 

William F. W jn 

Bell Helicopx^r Company 

Robert A. Wolfe 

U. S. Army Aviation Systems Command 

Edward R. Wood 

Lockheed California Company 

Clifton Wrestler, Jr. 

U. S. Army Materiel Command 

Jing G. Yen 

Bell Helicopter Company 

Maurice I. Young 
University of Delaware 

Steven E. Zalesch 

Naval Air Systems Command 



•k U.S. GOVERNMENT PBINTING OFffrCE: 197*~635-Q43/25 



371 



NATIONAL AERONAUTICS AND SPACE ADMINISTRATION 
WASHINGTON, D.C. 20546 



OFFICIAL BUSINESS 
PENALTY FOR PRIVATE USE *300 



SPECIAL FOURTH-CLASS RATE 
BOOK 



POSTAGE AND FEES PAID 

NATIONAL AERONAUTICS AND 

SPACE ADMINISTRATION 

451 




POSTMASTER : 



If Undeliverable (Section 158 
Postal Manual) Do Not Return 



"The aeronautical and space activities of the United States shall be 
conducted so as to contribute . . . jo the expansion of human knowl- 
edge of phenomena in the atmosphere and space. The Administration 
shall provide for the widest practicable and' appropriate dissemination 
of information Concerning its activities and the results thereof." 

— National Aeronautics and Space Act of 1958 



NASA SCIENTIFIC AND TECHNICAL PUBLICATIONS 



TECHNICAL REPORTS: Scientific and 
technical information considered important, 
complete, and a lasting contribution to existing 
knowledge. 

TECHNICAL NOTES: Information less broad 
in scope but nevertheless of importance as a 
contribution to existing knowledge. 

TECHNICAL MEMORANDUMS: 
Information receiving limited distribution 
because of preliminary data, security classifica- 
tion, or other reasons. Also includes conference 
proceedings with either limited or unlimited 
distribution. 

CONTRACTOR REPORTS: Scientific and 
technical information generated under a NASA 
contract or grant and considered an important 
contribution to existing knowledge. 



TECHNICAL TRANSLATIONS: Information 
published in a foreign language considered 
to merit NASA distribution in English. 

SPECIAL PUBLICATIONS: Information 
derived from or of value to NASA activities. 
Publications include final reports of major 
projects, monographs, data compilations, 
handbooks, sourcebooks, and special 
bibliographies. 

TECHNOLOGY UTILIZATION 
PUBLICATIONS: Information on technology 
used by NASA that may be of particular 
interest in commercial and other non-aerospace 
applications. Publications include Tech Briefs, 
Technology Utilization Reports and 
Technology Surveys. 



Details on the availability of these publications may be obtained from: 
SCIENTIFIC AND TECHNICAL INFORMATION OFFICE 
NATIONAL AERONAUTICS AND SPACE ADMINISTRATION 

Washington, D.C. 20546