Rotary Wing Structural Dynamics and Aeroelasticity
Rotary Wing Structural Dynamics and Aeroelasticity
Rotary Wing Structural Dynamics and Aeroelasticity
Structural Dynamics
and Aeroelasticity
Second Edition
Richard L. Bielawa
EDUCATION SERIES
Joseph A. Schetz
Series Editor-in-Chief
Virginia Polytechnic Institute and State University
Blacksburg, Virginia
Published by
American Institute of Aeronautics and Astronautics, Inc.
1801 Alexander Bell Drive, Reston, VA 20191
1 2 3 4 5
Copyright © 2006 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
Printed in the United States. No part of this publication may be reproduced, distributed, or transmitted,
in any form or by any means, or stored in a database or retrieval system, without the prior written
permission of the publisher.
1-56347-698-3
Data and information appearing in this book are for information purposes only. AIAA is not responsible
for any injury or damage resulting from use or reliance, nor does AIAA warrant that use or reliance
will be free from privately owned rights.
Editor-in-Chief
Joseph A. Schetz
Virginia Polytechnic Institute and State University
Editorial Board
Takahira Aoki David K. Holger
University of Tokyo Iowa State University
It is a privilege and pleasure to help bring this second edition to the attention of
the widest possible readership. It follows a very well received first edition and, as
did its predecessor, provides a comprehensive yet accessible account of many of
the important concepts and developments in the field by a master of the subject.
Dr. Bielawa is superbly qualified by his experience in industry and academe to
contribute an authoritative yet clear account of this fascinating set of topics.
Those of us who have on occasion worked in rotorcraft structural dynamics and
aeroelasticity, but whose primary work has been in fixed wing aircrafl including
turbomachinery, have only the highest respect for those engineers who have taken
up the high challenges that rotorcraft provide. These challenges emerge across
a spectrum of activities from the intellectual, computational, and experimental
to perhaps the greatest challenge of all in design. They arise from a wide range
of disciplines including fluid dynamics, structural dynamics, dynamic response,
control, and stability. To bring coherence and structure to this wide range of relevant
activities and disciplines in the context of rotorcraft is a major challenge in itself
and one that is well met in this volume.
This book will be of great value to all those who participate in one or more
of these activities and disciplines. The scope of the treatment is broad yet deep.
Experienced rotorcraft engineers will benefit from an account of new ideas as well
as from a ready reference for now classical material. The beginning student, with
the help of an experienced guide such as Dr. Bielawa or one of his peers, will find
this a reliable and engaging introduction to this fascinating topic. Self-study is also
possible and frequently this book will be one of the best places to begin the study
of a topic in rotorcraft structural dynamics and aeroelasticity.
Dr. Bielawa has again performed a great service by bringing forth this new
edition. I am confident that it will find a well deserved and highly valued place in
the libraries of rotorcraft engineers, both present and future.
Earl Dowell
Duke University
ix
Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Rotary Wing vs Fixed Wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 719
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 729
Richard L. Bielawa
October 2005
xv
The material presented herein thus deals with fundamental issues that form a
matrix into which all of the more detailed and newly emerging engineering studies
can be placed. To be certain, much of the text deals with diverse methodologies
relating to the rotational dynamics of rotors. At an even more fundamental level
basic mathematical tools are presented which relate to the analysis requirements
of multiple-degree-of-freedom dynamic systems. For many practicing engineers
this mathematical material might be of limited novelty, but for others its inclusion
should prove to be quite useful.
Because the subject technology has two main drivers, the minimization of vibra-
tion and the assurance of aeromechanical and aeroelastic stability, this text has been
organized sequentially as much as possible into these two basic areas. As required
for both of these areas, Chapters 1–4 present various basic analysis tools which lay
the necessary groundwork for the remainder of the text. Beyond the basic tools pre-
sented in these first chapters, additional methodologies are presented throughout
the text on an as-needed basis. Chapters 5–8 are focused on the airframe vibration
problem with an intended emphasis on tools for designing vibration out of a new
design and/or fixing unaccounted-for vibration problems once they have surfaced.
The rotor dynamic environment is rich in possibilities for a variety of aeronauti-
cal and aeroelastic instabilities. Indeed, those instability phenomena presented in
Chapters 10–13 are not intended to constitute an exhaustive exposition. Instead,
the phenomena presented were selected on the bases that they either constitute the
more pertinent phenomena driving the rotorcraft design, or they form the bases for
other more complex phenomena. Finally, because the available analytical tools are
as yet not completely reliable (much less “user-proof”), successful integration of
the technology into real-world applications must have experimental verification.
Consequently, Chapters 8 and 14 are respectively directed at experimental tech-
niques found to be useful in each of the two basic technology areas. The text has
been organized in this manner so that, hopefully, it will be a practical and conve-
nient teaching resource for either a two-semester or three-quarter course sequence
at the graduate level. The objective of providing a vehicle for the formal teaching
of the material in a university environment consequently required the composition
and incorporation of suitable exercise problems. The problems compiled include
a variety of relatively simple ones as well as more difficult ones; in all of the prob-
lems presented the intent is to teach and thereby to help readers achieve mastery
of the material.
The objective of addressing the emerging, more advanced rotary wing concepts
required the inclusion of material relating to propeller-nacelle whirl flutter, as well
as some analysis material relating to multiple rotor dynamic systems. The extensive
analytical development presented in Appendix D was primarily formulated as
a rudimentary analysis tool for helicopter ground resonance and air resonance.
However, its inclusion in this text was justified on the grounds that it also serves
the useful purpose of defining a basic, practical aeroelastic rotor module, to be
used as a building block for formulating larger analyses of multiple rotor (tilt-rotor
and/or tilt-wing) aircraft aeroelastic stability.
This textbook was not written in a vacuum and in many instances the text is
a selective compilation of the findings of many other researchers. Furthermore,
without the encouragement of friends and colleagues it would not have been writ-
ten. Special acknowledgments in this category go to William Twomey, William
Richard L. Bielawa
June 1992
At its basic roots, the subject matter of rotary wing structural dynamics
and aeroelasticity is concerned with the following problem: How can structural
integrity and passenger comfort be attained in the vibratory environment pecu-
liar to rotary wing aircraft? From this broad problem definition has evolved the
two principal areas of concern in rotary wing structural dynamics: vibrations and
aeroelastic stability. The emphasis is currently on vibration reduction, but with
close attention being paid, nonetheless, to the fact that rotors are subject to a
variety of potentially unstable response phenomena. The advent of new rotor con-
cepts increases the probability of an otherwise benign characteristic rotor response
becoming a major aeroelastic instability issue. In any practical design the insta-
bilities must be well understood, and ways must be found to suppress them in all
conceivable flight conditions. Last, it should be noted that one area of concern
that closely relates to structural dynamics and aeroelasticity is that of noise. The
high degree of noise characteristic of most rotorcraft has spawned the growing
technical areas of rotor far-field noise acoustics and structure-borne interior noise
acoustoelasticity. Although both these subjects are important and timely, this text
is limited to considerations of only vibrations and aeromechanical and aeroelastic
instabilities.
are “structural lightweights” compared with their fixed wing counterparts. They
achieve a large measure of their stiffening from the tension induced by the rota-
tional centrifugal force field. The rotational environment of the rotor blades also
gives rise to a host of rotation-related phenomena: gyroscopic characteristics,
Coriolis forces, and a variety of nonlinear inertial loadings. While still governed
principally by linear operators, the resulting aeroelastic description of rotor blade
elastic responses is consequently fraught with nonlinearities that modify the results
obtained using only linear analysis.
The relatively high degree of rotor flexibility also drives the significant interac-
tion occurring between the rotor and the also flexible rotorcraft airframe. At present,
the stubbornness of typical modern-day helicopter airframes to yield to accurate
dynamic analysis continues to pose a very important challenge to the rotorcraft
structural dynamist. Furthermore, contrary to the case of fixed wing aircraft, where
aeroelastic stability characteristics are the priority issue, the dynamic analysis of
rotorcraft airframes must be relatively much more accurate to enable a reasonably
accurate calculation of the vibration characteristics.
1.2 Methodology
Stated in the context of an engineering solution, the structural dynamics problem
of rotorcraft entails three general concurrent avenues of approach: 1) Knowledge
of the vibratory environment of the rotorcraft must be acquired. Principally, this
means knowing the essentially unsteady aerodynamic characteristics of rotorcraft.
2) The extent to which the structure responds to the environment must be calcu-
lated. 3) The resulting responses must be judged for acceptability; if they are not
acceptable, ways must be found to make them so.
helicopter fuselage by an unbalanced rotor, and the (number of blades) per rev
accelerations resulting in the fuselage from a periodic main rotor wake impinge-
ment on the horizontal stabilizer. An example of an unstable response is the flutter
of a wing or rotor blade. Customarily, a resonant response is ideally thought of
as the excitation of a structure at one or more of its natural frequencies by an
external energy source, whereas an unstable response is thought of as a structure
driving itself from an internal energy source. It is a complication with rotorcraft
that, in practice, it is often difficult to ascertain which of these responses is in fact
occurring. This is because both types of response can exhibit measured behavior
that “blossoms” for periods of time, and furthermore, both types can exhibit limit
amplitudes of sinusoidal motion.
these problem areas, i.e., vibration and aeroelastic stability, chapters are presented
dealing with appropriate experimental procedures.
The third section of the text consists of Chapters 15 through 18, which deal with
material that has special applications to both of the two main areas. Chapter 15
presents material relating to elastomeric devices, as these have found a permanent
niche in rotorcraft structural dynamics applications. Chapter 16 presents specifics
for characterizing the blade section with an emphasis on the application of compos-
ite materials. Chapter 17 describes instances of cross-over interactions of “outside”
dynamics-related material with the more standard material of rotorcraft structural
dynamics covered in the previous chapters. Finally, Chapter 18 presents some
concluding thoughts together with an abbreviated description of key milestones in
the development of rotary wing structural dynamics and aeroelasticity.
It is to be hoped that this text will provide an in-depth introduction to the subject
material as well as a useful reference resource both for actual formulations and
bibliographies. The theory of linear differential equations is the primary requisite
mathematical discipline employed in this text; most of the analytical formulations
involve only the integrating of a handful of basic, relatively simple concepts.
The essential tools needed for analyzing the structural dynamics of rotorcraft
must address a variety of issues. The principal problem areas are characterized by
1) small motions, usually sinusoidal in nature; 2) several degrees of freedom of
motion; and 3) a variety of loadings arising from rotational effects. Much attention
is therefore paid to identifying and reviewing pertinent methods for analyzing
linear dynamic systems, starting from the simplest configuration (single degree
of freedom) and progressing to general configurations involving several degrees
of freedom. Basic techniques dealing with “frequency-domain” problems are also
identified and reviewed.
The appropriate dynamic equation is obtained by taking the mass as a free body
and examining and equilibrating the forces acting on it. One indispensable tool for
analyzing dynamic systems is D’Alembert’s principle, which allows the conversion
of dynamic problems to static ones. Typical usage of this principle consists of
representing the statement of Newton’s second law (that F = ma) by an inertial
load equal to ma, but directed opposite to the acceleration a. With this approach
the forces acting on the mass can be equilibrated, as shown in Fig. 2.2, where the
5
xh (t) = x̄eλt
(mλ2 + cλ + k)
x̄eλt = 0
(not, in general, = 0)
Therefore,
mλ2 + cλ + k = 0 (2.6)
This is the basic characteristic equation whose solution (the two values of λ) is
given by
λ1 , λ2 = −c/2m ± i k/m − (c/2m)2 (2.7)
Thus,
Thus, the homogeneous solution for x(t) can then be written in the following form:
where
and C1 and C2 are arbitrary constants determined from the initial conditions. For
the undamped case c = 0, the homogeneous solution becomes a simple sinusoid:
x(t) = C1 cos k/m t + C2 sin k/m t (2.12)
Note: The frequency increases with spring stiffness k and decreases with inertia m.
At this condition both of the characteristic values λ1 and λ2 become equal, and the
appropriate solution then becomes
For systems with damping levels less than the critical value, the responses will
always be oscillatory. Those systems with damping values equal to or in excess of
the critical value will always be aperiodic, and the response with the critical value
can be shown to approach quiescence aperiodically most quickly. These points are
depicted in Fig. 2.4.
Although it is unlikely that a new class of structures will exhibit damping char-
acteristics anywhere approaching critical damping in the near future, the concept of
critical damping is nonetheless quite useful because it separates periodic responses
from aperiodic ones. This feature provides a clear-cut yardstick for measuring
damping. Because the dynamic response is seen to depend on the interrelationship
of inertia, damping, and stiffness, some meaningful measure is needed to describe
how much damping is present in any given system, irrespective of the scale of the
system (as defined by the mass and stiffness).
One method for achieving this measure of damping is afforded by first recasting
the form of the basic differential equation of motion by dividing it by the mass:
where ζ is denoted the critical damping ratio for the system and ωn is the undamped
natural frequency. With these two concepts the basic equation of motion can then
be put into the following useful form:
σ = −ζ ωn (2.18a)
ωD = 1 − ζ 2 ωn (2.18b)
To summarize,
where
U( f ) = cycles/s = (Hz); U(ω) = rad/s (2.21)
initial conditions for a complete time-history solution requires that both homo-
geneous and particular solutions be taken together before the values of the
undetermined coefficients in the homogeneous equation can be ascertained.
A lot of practical engineering use can be made of both the homogeneous and
particular solutions separately without resorting to constructions of complete solu-
tions at this point. Problems requiring the detailed construction of such complete
time-history solutions do arise, however, and will be treated in later chapters as
needed.
Because we have assumed that the forcing function has been operating on the
linear spring-mass-damper system for a sufficiently long time, the homogeneous
solution can therefore be assumed to have died out. The resulting response
produced will then also be sinusoidal and at only one frequency: the same frequency
as the applied load frequency . Therefore, the particular solution for this case
will have the following form:
Consequently,
Substitution into the basic differential equation and collecting terms yields
P0
(ωn2 − 2 )A + 2ζ ωn B = (2.29a)
m
−2ζ ωn A + (ωn2 − 2 )B = 0 (2.29b)
Therefore,
A = X cos φ (2.31a)
B = X sin φ (2.31b)
so that
and
X= A2 + B2 (2.31d)
P0 /m
X= ≡ response amplitude (2.32)
(ωn2 − )2 + (2ζ ω
2
n )
2
2ζ ωn
φ = tan−1 ≡ phase angle (2.33)
ωn2 − 2
Plots of these two basic functions with variations in forcing frequency and
damping are given in Fig. 2.5.
X 1 P0 P0
= 2 −→ X= = (2.34a)
( P0 /m) ωn mωn2 k
For = 0, φ=0
For → ∞, φ=π (2.34b)
The preceding material summarizes the basic material used for simple harmonic
(sinusoidal) motion. The next section generalizes these concepts to the case
wherein the simple single-degree-of-freedom system is excited by multiple
frequencies, which are also integral multiples of a basic frequency. As we shall
see in later sections, this situation is of prime importance to rotorcraft dynamic
vibration problems.
Again, because the dynamic system is linear, the response will also be at the same
frequency:
where
If the excitation were now to consist of the steady load as well as a summation
of such harmonic excitations, wherein n is varied with increasing integer values
(n = 1, 2, 3, . . .), the excitation is then expressible as a Fourier series:
N
F(t) = F0 + [Fnc cos nt + Fns sin nt] (2.38)
n=1
N
xp (t) = A0 + [An cos nt + Bn sin nt] (2.39)
n=1
where the An and Bn coefficients derive from the Fnc and Fns excitation Fourier
coefficients as given earlier.
1 T
F0 = F(t) dt (2.40b)
T 0
so that
Thus, for the discretization shown in Fig. 2.6 the following expressions can be
written:
F0 = f0 (2.45a)
.. ..
. .
FN = fN (2.45h)
Note that, by virtue of Eq. (2.43), G0 and GN are identically zero. The approximate
resulting Fourier series is then expressed as
N−1
f (t) = A0 + [An cos n(2π t/T ) + Bn sin n(2π t/T )]
n=1
The Fourier coefficients in this representation are obtainable from the F and G
functions as follows:
1 1 1
A0 = F0 + F1 + F2 + · · · + FN−1 + FN (2.47)
N 2 2
2 1
An = F0 + F1 cos nx1 + F2 cos nx2 + · · ·
N 2
1
+ FN−1 cos nxN−1 + (−1)n FN (2.48)
2
1 1 1
AN = F0 − F1 + F2 − · · · + (−1)N−1 FN−1 + (−1)N FN (2.49)
N 2 2
2
Bn = [G1 sin nx1 + G2 sin nx2 + · · · + GN−1 sin nxN−1 ] (2.50)
N
BN = 0 (2.51)
where
xi = 2π ti /T (2.52)
1 +∞
F(ω) = f (t) e−iωt dt (2.55)
2π −∞
The primary usefulness of the Fourier transform is that it provides a very general
method of determining particular solutions for a wide range of (aperiodic) exci-
tations. Some problems for which the excitation cannot be expressed in the time
domain, but rather only in the frequency domain (and, even then, sometimes only
using statistical concepts), are intractable with any solution method other than the
Fourier transform. In particular, assume for the moment that the applied excitation
to the dynamic system F(t) is expressible only in the form of its frequency content
[i.e., F(ω), a Fourier transform of the excitation]. This is often the case with exper-
imentally acquired data wherein the Fourier transforms are produced in software
by the test equipment. The frequency response function (FRF) of the dynamic
system H(ω) is a frequency-domain representation of the sinusoidal output of the
dynamic system caused by a sinusoidal input. For the single-degree-of-freedom
system, H(ω) would be expressed as
P0 /m
H(ω) = (2.56)
ωn2 − ω2 + i2ζ ωn ω
Then, the Fourier-transform method could be used to obtain the Fourier transform
of the product response X(ω):
Finally, the actual time domain solution x(t) is then obtained using the inverse
Fourier transform:
+∞
x(t) = H(ω)F(ω)eiωt dω (2.58)
−∞
a step removed from the details of the actual configuration and hence from an
understanding of the physics of the resulting dynamics.
Use of the Lagrangian approach hinges on the concepts of kinetic and potential
energy, T and U, respectively, as defined for the total system. The Lagrangian
approach utilizes a single differential equation that can be used for any dynamic
system, linear or nonlinear, and is applicable to every one of the (M) degrees of
freedom:
d ∂T ∂T ∂U ∂D
− + + = Fi i = 1, 2, . . . , M (2.63)
dt ∂ q̇i ∂qi ∂qi ∂ q̇i
where
T ≡ kinetic energy,
U ≡ potential (strain) energy,
D ≡ dissipation function,
qi ≡ ith generalized coordinate, and
Fi ≡ generalized force for ith generalized coordinate
where the generalized coordinates and generalized forces are simply given as
q1 ≡ x1 ; q2 ≡ x2 ; F1 ≡ 0; F2 ≡ F(t)
x1 = x 1 eλt (2.66a)
x2 = x 2 eλt (2.66b)
With these substitutions in the equations of motion, the equation set then becomes
a set of algebraic equations and, in essence, an eigenvalue problem:
and solving the resulting polynomial for λ. This polynomial is called the
characteristic equation.
3) Once the various values of λ are known, for each λ a mode shape x̄1 /x̄2 can
be found. That is, only the relation between x̄1 and x̄2 is obtainable; either x̄1 or x̄2
can have any arbitrary value we might select for it. Typically, the selected value is
unity.
Undamped case. For the general damped case the resulting characteristic
equation is a quartic and is typically not solved using simple analytical formulas.
Also, both the eigenvalues and eigenvectors are usually complex valued. However,
the undamped case leads to a quartic polynomial characteristic equation devoid
of odd powers that can then be readily solved using the binomial theorem. Thus,
the characteristic equation (λ) = 0 has two imaginary root pairs given by the
following solutions:
1/2
2 1/2
m2 (k1 + k 2 ) + m 1 k2 k1 k2 m2 (k1 + k2 ) + m1 k2
λ = ±i ± − +
2m1 m2 m1 m2 2m1 m2
(2.69)
The two quantities multiplying (±i) in the preceding equation constitute the
resulting natural frequencies for the dynamic system ω1 and ω2 . The mode shapes
can then be found from either of the system differential equations: If we arbitrarily
set x̄1 = 1, then
or
m1 λ2 + (k1 + k2 )
x2 = (2.71)
k2
m1 = m2 = m (2.72a)
k1 = k2 = k (2.72b)
Thus,
ω1 = 0.618 k/m (2.74a)
ω2 = 1.618 k/m (2.74b)
These are then the two (undamped) natural frequencies for this two-degrees-
of-freedom spring-mass system. If we alternatively let x̄2 = 1, then for each of
the two natural frequencies (m = 1, 2),
λ2m + (k/m) √
(m)
x1 = =1− 1
2 3± 5 (2.75)
(k/m)
Thus,
(1)
x 1 = 1 − 0.382 = 0.618 (mode 1) (2.76a)
(2)
x 1 = 1 − 2.618 = −1.618 (mode 2) (2.76b)
These results are given pictorially in Fig. 2.8. This figure depicts the relative
positions the masses have with respect to each other when they are oscillating
freely at the respective natural frequencies. The positions of x̄1 and x̄2 (relative to
each other) are the natural mode shapes.
this simple problem, but can easily be generalized to other more complicated, but
linear, dynamic systems. We begin by rewriting the two homogeneous equations
as follows:
mode 1 (2.78)
(1) (1) (1)
m2 ω12 x 2 = −k2 x 1 + k2 x 2
(for ω = ω2 )
m1 ω22 x 1 = (k1 + k2 )x 1 − k2 x 2
(2) (2) (2)
mode 2 (2.79)
(2) (2) (2)
m2 ω22 x 2 = −k2 x 1 + k2 x 2
Each of the two equations of motions can then be written symbolically (for the ith
equation and pth mode and for n = 2) as
n
( p) ( p)
mi ωp2 x i = Kij x j (2.80)
j=1
n
(r) (r)
mi ωr2 x i = Kij x j (2.81)
j=1
n n n
(r) ( p) (r) ( p)
ωp2 mi x i x i = Kij x i x j (2.82a)
i=1 i=1 j=1
n n n
( p) (r) ( p) (r)
ωr2 mi x i x i = Kij x i x j (2.82b)
i=1 i=1 j=1
Because for conservative systems Kij = Kji , the sums on the right-hand sides
of these two equations are equal. Therefore, if the two equations are differenced,
the following result is obtained:
n
( p) (r)
(ωr2 − ωp2 ) mi x i x i =0 (2.83)
i=1
Because, for p = r, the first term is clearly nonzero, the second term must be zero.
This result is summarized as follows:
n
( p) (r) 0, r = p
mi x i x i = (2.84)
Mp , r=p
i=1
r = 1, p = 2 :
r = 2, p = 1 :
r = 2, p = 2 :
Thus, these results indeed demonstrate that the orthogonality principle is real.
The property is no less real for larger systems and actually proves to be a good test
for the accuracy of modes, as calculated by any given eigenvalue solution scheme.
Furthermore, although the orthogonality principle is explicitly formulated earlier
for undamped systems, the principle is still valid for damped systems (both sym-
metric and unsymmetric terms), but requires a different formulation. The reader
is referred to Hurty and Rubenstein for a discussion of both proportional and non-
proportional damping cases. (The term proportional damping refers to damping
terms in the equations of motion that are proportional to linear combinations of
already existing inertia and/or stiffness terms.)
( p)
where x̄m represents the mth component (degree of freedom) of the pth mode
( p)
shape. [Note that here we have assigned the value of 1 to x̄2 .]
2) This coordinate transformation is then substituted into the differential
equations of motion.
3) For each of the two values of p (1 and 2), the first equation is multiplied by
( p) ( p)
x̄1 and the second equation by x̄2 .
4) For each value of p, this multiplication operation is then followed by a
summation of the two equations. By virtue of the orthogonality principle, each
of the two resulting equations ( p = 1, 2) is then uncoupled from each other to
produce equations of the following form:
where
qp (t) ≡ pth modal variable, a scalar quantity denoting how much the pth natural
mode is being excited
Mp ≡ pth generalized mass corresponding to the pth natural mode
ωp ≡ pth natural frequency
p ≡ pth generalized forcing function or excitation, the measure of how
much the pth mode is being excited
Note that a complete solution of the equations is then afforded by solving for the
generalized (modal) variables first and then combining them as per the coordinate
transformation just given. The aforementioned development, although formulated
in terms of only two degrees of freedom, can be easily generalized to any number of
degrees of freedom. Thus, by means of natural modes multiple-degree-of-freedom
systems can be reduced to an ensemble of single-degree-of-freedom problems. For
systems with (viscous) damping similar (but not identical in detail) techniques can
be used to uncouple the modes one from the other.
The procedure just given is an example of a more generalized procedure known
as the Galerkin method, which will be covered in more detail in a later section.
The Galerkin method is a powerful solution technique that can be used not only
for devising modal solution approaches to linear systems, but for treating modal
Then, in the absence of damping the two degrees of freedom x1 and x2 can be
similarly defined:
P0 k2
X1 = (2.90a)
m1 m2 (ω12 − 2 )(ω22 − 2 )
P0 (k1 + k2 − m1 2 )
X2 = (2.90b)
m1 m2 (ω12 − 2 )(ω22 − 2 )
Plots of the amplitude variations of these two responses are given in Fig. 2.9.
Points to be noted are as follows:
1) Similar to the single-degree-of-freedom results, the system exhibits
resonance conditions (i.e., the responses become unbounded) when the forcing
frequency approaches either one of the two natural frequencies ω1 and ω2 .
2) At a certain frequency ω̂3 , situated between ω1 and ω2 , the response of X2
becomes zero. Such a condition is typically denoted an antiresonance condition.
Physically (at this frequency) the first mass m1 is acting as a “vibration absorber”
to the second mass. The reason that the response of the second mass can be zero
(despite the fact that it is receiving the full brunt of the application of the external
load) is that the first mass, together with the springs, develops an equal and oppo-
site load “in tune” with the forcing frequency . From the equation for X2 , this
frequency is seen to be
ω̂3 = (k1 + k2 )/m1 (2.91)
From a consideration of the physics of the configuration, this is also the natural
frequency of the first mass operating as a single-degree-of-freedom system with
the second mass constrained to have zero motion, x2 ≡ 0.
3) For excitation frequencies between ω1 and ω̂3 and from ω2 to ∞, the X2
response curve shows negative values. Similarly, the X1 response curve is negative
between ω1 and ω2 . Generally, a negative value is to be interpreted to mean that
the response is 180 deg out of phase with the excitation F(t).
4) Apart from the antiresonance issue, each of the response curves can be
interpreted to be comprised of an ensemble of single-degree-of-freedom-system
response curves overlapping each other. These single-degree-of-freedom systems
are clearly the modal response degrees of freedom treated in a preceding section.
Damped case. The reinstatement of the dampers shown in Fig. 2.7 to
nonzero values would result in a system of four algebraic equations in the four
(real) coefficients needed to describe the response of two degrees of freedom. The
solution cannot be stated easily in a way so as to glean out the physics. However,
resorting to a modal interpretation produces the result that the inclusion of damp-
ing modifies the results given earlier for the undamped case in ways suggested by
the single-degree-of-freedom results:
1) The singular or unbounded results at the resonance conditions disappear to be
replaced by high (but bounded) responses that attenuate with increased values
of damping.
2) The antiresonance condition develops a finite (but still small) response level,
depending on the damping level. At the minimum response level the phase
angle of the second mass would be near 90 deg because the response would be
resisted principally by the damping.
Thus, it can be inferred that the optimal operation of vibration absorbers based
on an antiresonance principle requires a minimization of damping in the system.
Therefore,
Fz = F(t) (2.92d)
where the ( ) and ( ) underlined terms represent, respectively, the inertia and
stiffness coupling terms between the two degrees of freedom.
Free-free systems. Often the dynamic system is not rigidly fixed to a sta-
tionary point. In such a case a rigid-body mode appears that has zero frequency and
involves zero spring deflections. Furthermore, for free vibrations the system then
has a center of gravity that does not move in space and hence becomes a system
node point. Consider the example given in Fig. 2.11 of two flywheels connected
dumbbell-like to each other by a flexible shaft:
Let us find the natural frequency and node point for this system. The equations
of motion can be easily determined to be
yields the following simultaneous equations for the modal response values:
ω4 J1 J2 − ω2 K(J1 + J2 ) = 0 (2.97)
The zero frequency corresponds to a rigid-body mode, whereas the nonzero one
corresponds to the only elastic mode possible. Solving for this elastic mode gives
then
or
1 / = J2 /(J1 + J2 ) (2.102)
where
= 1 + 2 (2.103)
The same result can be obtained by considering the node point to be built in, as
shown in the following sketch:
Here the natural frequency squared calculation can be made on the basis of a
single-degree-of-freedom system:
where
K1 = GJ/1 (2.105a)
K2 = GJ/2 (2.105b)
mf ( yg + 2 φf ) + mt ( yg − 1 φt ) = 0 (2.107)
which results in a dependence of the lateral translation of the gimbal on the fuselage
and transmission roll angles and, as a consequence, a reduction in the number of
independent degrees of freedom needed to describe the dynamics:
mt 1 φt − mf 2 φf
yg = (2.108)
mf + mt
Using Lagrange’s equation, one can then write the appropriate kinetic and potential
energy relationships:
The generalized “forces” appropriate to the transmission and the fuselage Fφt and
Fφf , respectively, are to be found by considering the incremental work accom-
plished by the forces acting on the two masses. In this case these loads are the
gravity loads Wt and Wf , which are equilibrated by the total rotor lift L. Note that,
in this case, because work is accomplished only by virtue of the two rotations the
generalized forces will actually be moments.
Therefore,
mt mf
T= (1 φ̇t + 2 φ̇f )2 + 21 (It φ̇t + If φ̇f )2 (2.113)
(mt + mf )
Application of Lagrange’s equation then yields the following preferred form of the
equations of motion:
mt mf 21
+ It φ̈t + (Kr + Kg + Lt − Wt 1 )φt
mt + mf
mt mf 1 2
+ φ̈f − Kg φf = 0 (2.114)
mt + mf
mt mf 1 2 mt mf 21
φ̈t − Kg φt + + If φ̈f + (Kg + Wf 2 )φf = 0 (2.115)
mt + mf mt + mf
These equations are then similar in form to those of the other two-degree-of-
freedom systems considered earlier in this chapter, wherein both inertia and
stiffness couplings are present.
2.4.1 Definition
The Laplace transform is similar to the Fourier transform in that it transforms
functions of time into a new variable. In this case the new variable p is complex-
valued. The complete definition of the Laplace transform of the function f (t) is
given by
∞
L [f (t)] ≡ F( p) = f (t)e−pt dt (2.116)
0
where f (0) and df (0)/dt are the initial conditions relating to f (t).
In practice the completion of the solution is to reduce the transformed variable
F( p) into a linear combination of specific forms that have known transform pairs
and then use the pairs to get back into the time domain. Although many such pairs
are available in the literature (e.g., Churchill), four transforms that find a lot of
application in structural dynamics and aeroelasticity are
1
L[1] = (2.118a)
p
1
L[eat ] = (2.118b)
p−a
a
L[sin at] = 2 (2.118c)
p + a2
p
L[cos at] = 2 (2.118d)
p + a2
N( p) Bm pm + Bm−1 .pm−1 + · · · B0
L [f (t)] = F( p) = = (2.119)
D( p) pn + An−1 pn−1 + · · · A0
This form can be reduced to elementary forms by first factoring the denominator
function D( p). This chore amounts to finding the roots of the denominator function
(using any of a variety of algorithms available, see Press et al.). All of the roots of
the denominator polynomial can be placed in the complex plane, which we now
refer to as the Laplace transform domain. Equation (2.119) then becomes
N( p) Bm pm + Bm−1 · pm−1 + · · · B0
F( p) = = (2.120)
D( p) ( p − a1 )( p − a2 ) · · · [( p − σj )2 + ωj2 ] · · ·
n
−1 −1 N( p) N(aj )
L [F( p)] = L = exp(aj t) (2.121)
D( p) D (aj )
j=1
It is to be understood that if any of the roots are complex, Eq. (2.121) must be
evaluated in complex arithmetic. Also, if roots are repeated, the use of continued
fractions can still be used with appropriate modifications. The complete menu of
techniques available for using partial fractions to inverse transform expressions is
beyond the scope of this text, and the reader is referred to Churchill or an equivalent
for such usage.
where c̄ = gωn = 2ζeq ωn , g is the structural damping coefficient, and ζeq is the
viscous equivalent critical damping ratio.
ẍ + (1 + ig)ωn2 x = 0 (2.123)
to (coupled) modes that are complex valued rather than real valued (which are
observed) because these damping terms do not generally obey the same ortho-
gonality relationship that the mass and stiffness terms do. However, again because
of the ease of implementation, these formulations are still extensively used in the
analyses of multiple-degrees-of-freedom systems despite such anomalies.
Using the results given earlier for the homogeneous solutions for the single degree
of freedom, this logarithmic decrement can be related to the critical damping ratio
ζ , as well as to the structural damping coefficient g:
g= ω/ω0 (2.126)
(i.e., in./lbf-s), the circle, as plotted from the admittance data, has these same units.
At the point of maximum velocity magnitude, the following conditions hold:
1) The frequency for maximum polar magnitude ω̂ is given by
ω̂ = ωn (2.127)
2) The radius (and center location) of the resulting circle R is related to the
equivalent viscous damping coefficient ceq and, hence, to the equivalent critical
damping ratio as follows:
1 1
R= = (2.128a)
2ceq 4ωn ζeq Mn
or
1
ζeq = (2.128b)
4ωn Mn R
where ωn and Mn are, respectively, the natural frequency and generalized mass
for the single-degree-of-freedom system being considered. Note that whereas the
natural frequency is directly available using Eq. (2.127), the generalized mass must
be obtained from some other method.
3) Near the resonance condition wherein the polar amplitude is maximum,
the rate of change of the argument of the locus vector (i.e., phase angle) is also at
a maximum condition. The point of maximum rate of change of phase angle occurs
at frequency ω̃ as given by
1/2
ω̃ = ωn 2 1 − ζeq
2 −1 (2.129)
2.6 Matrices
2.6.1 Basic Definitions
A matrix is a rectangular array of elements of any size for which specific math-
ematical operations have been prescribed. These elements are usually numbers,
real or complex, whose established mathematical operations define a set of unique
characteristics. Because of these characteristics, matrices prove to be highly useful
in formulating and solving structural dynamics problems. The elements compris-
ing matrices can be quite general and can be functions of one or several variables. A
matrix is generally written as a bracketed quantity [A] or as a boldfaced variable A.
Basic definitions of matrices, both general and specific are given as follows:
a11 a12 ... a1n
a21 ... a2n
[A] = A = [aij ] =
... .. (2.130)
.
am1 ... amn
where aij represents an element in the ith row and jth column. The matrix A in the
preceding equation is said to be of order m × n.
The most frequently encountered matrices are square matrices. Square matrices
are matrices in which the number of rows equals the number of columns. If, for
example, A, the preceding matrix, were square, then m would be equal to n. In that
case the matrix A would be a matrix of the nth order. A few special types of square
matrices are listed in the following.
Diagonal matrix
0, i = j
aij = (2.131)
aii , i = j
Example:
$ %
a11 0 0
[A] = 0 a22 0
0 0 a33
Example:
$ %
1 0 0
[A] = 0 1 0 = [I]
0 0 1
Scalar matrix
0, i = j
aij = (2.133)
α, i = j
Example:
$ %
α 0 0
[A] = 0 α 0 = α[I]
0 0 α
Symmetrical matrix
Example:
$ %
3 7 1
[A] = 7 2 6
1 6 5
Skew-symmetrical matrix
Example:
$ %
0 1 −3
[A] = −1 0 2
3 −2 0
Note that the diagonal elements of the skew-symmetric matrix must necessarily
be zero because aij = −aji only for a value of 0.
Null matrix
aij = 0 (2.136)
Example:
$ %
0 0 0
[A] = 0 0 0
0 0 0
If a matrix consists only of a single row, it is termed a row vector, has order of
1 × n, and is denoted as
and
then
where
Matrix subtraction
where
Matrix multiplication
Scalar multiplication
where α is a scalar.
To multiply a matrix by a scalar, all elements of the matrix must be multiplied
by the same scalar.
The basic properties are as follows:
where
that is,
n n
c11 = a1k bk1 , c12 = a1k bk2 , etc. (2.145d)
k=1 k=1
b11 b12 ···
a11 a12 a13 · · · a1n
a21 b21 ···
a22 a23 ··· = [C]
.. ...
.
bn1
(a11 b11 + a12 b21 + · · · + a1n bn1 ) · · ·
= (a21 b11 + a22 b21 + · · · + a2n bn1 ) · · · (2.145e)
..
.
Example:
(2 × 3) (3 × 2) (2 × 2)
$ %
1 0
2 1 −1 7 −1
3 −1 =
0 −3 1 −11 3
−2 0
Pre- and postmultiplication. Given that [A][B] = [C], then [B] is said to be
premultiplied by [A], and [A] is said to be postmultiplied by [B].
Continued product. Provided that they are conformable, matrix products can
be cascaded:
[A][B][C][D] . . . [N] = [R] (2.151a)
[A] [B] [C] [D] . . . [N] = [R]
(2.151b)
m×n n×p p×q q × ··· ··· × y m×y
Also, matrix multiplication is associative:
([A][B])[C] = [A]([B][C]) (2.152)
In algebra,
ab = c
b=d (2.153a)
ad = c
With matrices,
[A][B] = [C]
[B] = [D] (2.153b)
[A][D] = [C]
then
Determinants. If [A] is any matrix, then the determinant of [A] exists only if
[A] is square. Det[A] = |A| implies a specific operation on the elements of [A]. A
minor of any element aij is the determinant of the matrix [A] wherein the ith row
and the jth column have been deleted:
a
11 a12 a13 · · · a1j · · · a1n
a
21 a22 a23 · · · a2j · · · a2n
· · · · ·
(minor), Mij ≡ · · · · · (2.157)
ai1 ai2 ai3 aij ain
· · · · ·
a
n1 an2 an3 · · · anj · · · ann
For any element aij whose minor is Mij , there exists a cofactor defined by
The solution is
|ai |
xi = (2.161)
|a|
where |ai | is the determinant of the coefficients with the ith column replaced by
the column of constants yj and |a| is the determinant of the coefficients.
where Cij are the cofactors of aij . Note that every square matrix has an adjoint.
Properties of the adjoint matrix are as follows:
1) [A] (adj[A]) = (adj[A])[A] = |A|[I] (commutative)
2) |adj[A]| = |A|n−1
[Proof: From property a, (adj[A])[A] = |A|[I]; therefore, |adj[A]| |A| = ||A| I| =
|A|n . Thence, a division by |A| produces the desired result.]
where
then
where
It can then be verified that this expansion allows for the following inverse for the
M matrix to be written:
−1 [I] −[A]−1 [B] [I] [0] [I] [0] [A]−1 [0]
[M] =
[0] [I] [0] [E]−1 −[C] [I] [0] [I]
(2.172)
Thus, the total inverse requires actual inverses to be calculated only of the A and
E matrices, each of which is of smaller dimension than M, thereby resulting in
savings in CPU time. Furthermore, if several inverses of M are required, wherein
only one or the other of the A and E matrices changes with each such inverse, then
the inverse of that submatrix need not be repeatedly calculated, resulting in still
further CPU reductions.
R = R1R (2.178)
|1R | = 1 (2.179b)
Thus,
1R = R/R (2.180)
B · C = BC cos θ (2.181)
and the angle between them θ are known), the following definition forms the basis
for evaluating the product when the components of the respective vectors are
known:
B · C = (bx i + by j + bz k) (cx i + cy j + cz k)
= bx cx + by cy + bz cz (2.183)
B × C = −C × B = 1n BC sin θ (2.185)
The result is a vector quantity with the unit vector 1n being perpendicular to both
the B and C vectors.
2) The cross products of the various unit vectors are as follows:
Triple products. Because the vector cross product is itself a vector, it can
then be used in vector multiplications with other vectors. Two such products are
then defined: a scalar (dot) triple product and a vector (cross) triple product.
1) The scalar triple product is the dot product of one vector with the result of
a vector cross product of two other vectors:
A · (B × C) = (A × B) · C = B · (C × A)
ax ay az
= [A B C] = bx by bz (2.189)
c c c
x y z
2) Similarly, the vector triple product is the cross product of one vector with
the result of a vector cross product of two other vectors:
R = R1R (2.191)
Therefore,
dR dR d1R
Ṙ = = 1R + R (2.192)
dt dt
dt
magnitude direction
change change
dR dR
= 1R + ω R × R (2.197)
dt dt
This equation is the basic tool that we can now use to bridge the gap between vector
algebra and vector calculus. It shows that an intrinsic relationship exists between
the differential of a vector and its orientation. A more complete understanding of
the vector differentiation process is afforded by considering rotations of vectors,
generally, as developed in the following section.
Fig. 2.17 Euler angles defining one set of three sequential coordinate transformations.
The sequence of angle rotations selected is not unique; there are 12 sets of Euler
angles depending on the order in which they are taken: three axes available for
the first rotation, two axes available for the second rotation, and two available for
the third rotation. Now let the angles α, β, and γ be “small,” and, further, let us
redefine these angles to be explicit rotations about each of the three principal axes:
α = θx (2.202a)
β = θy (2.202b)
γ = θz (2.202c)
[Note that if we had selected any one of the other five Euler angle sets involving
x, y, and t axes, the same coordinate system matrix equation for small angles, that
is, Eq. (2.203), would still have resulted.]
Then, after making the following shifts to vector notation,
x4
R4 = y4 (2.204a)
z4
x1
R1 = y1 (2.204b)
z1
θx
θ = θy (2.204c)
θz
R4 = R1 − θ × R1 (2.205)
This important vector equation forms the basis of the following subsection.
pS R = pV R + ωSV × R (2.208)
where ωSV is the rotational velocity vector of the variable coordinate frame rela-
tive to the stationary one. This equation is the statement of Coriolis’ theorem for
coincident origins. As such, it defines the inertial velocity in terms of velocities
measured in the variable coordinate frame.
A similar equation can be derived for the inertial acceleration of the velocities
and acceleration measured in the variable coordinate frame:
mp21 R = mG
+ F
(2.210)
gravitational external force
field intensity applied to mass, m
Let us now define gravity to include the gravitational field intensity and the
centrifugal force caused by Earth’s rotation:
g = G − ωie × (ωie × R) (2.211)
where ωie is the Earth’s rotational velocity vector relative to inertial space and
measured in the Earth-fixed coordinate system. Using the equation of Coriolis
for inertial coordinate system accelerations, we can formulate an adjusted version
of Newton’s law wherein the accelerations are now measured in the Earth-fixed
coordinate system:
mp2e R = mg + F − 2mωie × pe R (2.212)
This equation thus relates the forces acting upon a mass (gravity mg and some
other external applied force F ) to the acceleration and velocity measured relative
to the Earth.
References
Section
2.1–2.3 Ewins, D. J., Modal Testing: Theory and Practice, Research Studies
Press, Herts, England, U.K. 1984.
Gaukroger, D. R., Skingle, C. W., and Heron, K. H., “Numerical
Analysis of Vector Response Loci,” Journal of Sound and Vibration,
Vol. 29, No. 3, 1973, pp. 341–353.
Hurty, W. C., and Rubinstein, M. F., Dynamics of Structures, Prentice–
Hall, Englewood Cliffs, NJ, 1964.
Kennedy, C. C., and Pancu, C. D. P., “Use of Vectors in Vibration
Measurement and Analysis,” Journal of the Aeronautical Sciences,
Vol. 14, No. 11, 1947, pp. 603–625.
Scanlan, R. H., and Rosenbaum, R., Introduction to the Study of
Aircraft Vibration and Flutter, Macmillan, New York, 1951.
Tong, K. N., Theory of Mechanical Vibrations, Wiley, NewYork, 1960.
Problems
2.1 Consider the following differential equation for a single-degree-of-freedom
response variable X(t):
Ẍ + 25X = F(t)
a) F(t) = cos 5t
b) F(t) = 0.1Ẋ
c) F(t) = −0.1Ẋ
d) F(t) = cos 5t − 0.1Ẋ
and all cases have the same initial conditions on X and Ẋ as follows:
Find
(a) the undamped natural frequency, damping ratio, and fundamental
forcing frequency;
(b) the zeroth and first harmonic Fourier coefficients of F(t) using exact
and approximate numerical methods; and
(c) the zeroth and first harmonic amplitude responses of x(t) using the
exact Fourier coefficients for F(t).
The degrees of freedom are the plunge displacement z and the rotation angle θ
about the elastic center of the airfoil section. Also, let the section be quantified
by the following numerical values for the parameters:
Find
(a) the equations of motion for free oscillations;
(b) the natural frequencies and mode shapes; and
(c) the nodes (points with zero displacement) for each mode.
(Note: These points need not be located within the airfoil section.)
2.4 Derive the expressions used for each of the three methods of determining the
structural damping coefficients from experimental data.
2.5 For the radius of curvature method, derive the appropriate damping coeffi-
cient for the case of a complex stiffness representation for structural damping.
2.6 Derive the Euler angle transformation matrix relating an aircraft’s body axis
system {x, y, z}4 to the Earth-fixed coordinate system {x, y, z}1 (horizontal
plane with a normal vertical direction). The aircraft body-axis system is
defined by first, a heading angle ψ, as measured by compass heading; second,
a pitch attitude angle θ, that is, the angle the aircraft centerline makes with
the horizontal; and third, the aircraft roll angle φ, as measured about the
aircraft centerline from a wings-level position.
where X, Y , Z denote the inertial frame and x, y, z denote the rotating frame.
where
d
pI = pS ; pV =
dt
β0 = const; = const
2.8 Part 1. Show that the equations of motion (for small oscillations) of a freely
swinging pendulum can be written as
where
Part 2. Solve for the ( finite) frequencies of oscillation that exist. This is the
Foucault pendulum effect. [Remember that ωie is generally small compared
with the square root of (g/r).]
∂2
M(x, t) = EI z(x, t) (3.2)
∂x 2
yield the following basic differential equation for the beam in tension:
∂2 ∂ 2z ∂ ∂z
EI − T = fz (x, t) (3.3)
∂x 2 ∂x 2 ∂x ∂x
∂ 2z ∂ 3z
z(0, t) = (L, t) = (L, t) = 0 (3.4)
∂x 2 ∂x 3
Additionally, a fourth boundary condition must be imposed at the beam root based
on the type of fixity at that point. If the beam represents an articulated rotor blade,
then the beam has the boundary condition appropriate to a hinge, and the bending
moment is taken to be zero:
∂ 2z
(0, t) = 0 (3.5)
∂x 2
If the beam represents a hingeless rotor blade, then the beam root has the canti-
levered boundary condition, and the root slope is taken to be a constant equal to
some built-in coning value βB :
∂z
(0, t) = βB (3.6)
∂x
Distributed tension. The tension in the beam arising from the centrifugal
force field can be obtained by integrating the axial load distribution fx (=m2 x):
L
T (x, t) = 2
mx1 dx1 (3.7)
x
Correspondingly, the transverse load distribution for the in-plane motion case is
in the y direction and is similarly interpreted. Thus, for out-of-plane and in-plane
motions z and y, respectively, the transverse loadings are given as follows:
Out-of-plane bending:
∂ 2z
fz (x, t)inertial = −m (3.9a)
∂t 2
In-plane bending:
∂ 2y
fy (x, t)inertial = −m + m2 y (3.9b)
∂t 2
The second term in the expression for the in-plane inertial load distribution arises
from the fact that the centrifugal force field is radial; hence, a component of this
force field will be in the direction of the in-plane deformation y.
where γj (x) is the jth natural mode shape (which satisfies the boundary
conditions) and qj (t) is the jth generalized coordinate or modal response
variable.
One mathematical significance of the γj (x) function is that it satisfies the
following orthogonality condition:
L
Mj , for k = j
mγj γk dx = (3.12)
0 0, for k = j
This uncoupled equation for qj (t) then has the following general form:
where the generalized structure stiffness KSj and generalized centrifugal stiffness
KCFj follow directly from Eq. (3.13). Note that neither of these generalized stiff-
nesses, taken separately, demonstrates orthogonality, but when taken together with
the appropriate value of 2 they do, in fact, demonstrate orthogonality.
1) Isolate the linear portion of the equation as a left-hand side of the equation,
and (optionally) solve for the normal modes and natural frequencies.
2) Using the obtained normal modes or other approximate ones (which at least
satisfy the boundary conditions), premultiply the equation, in turn, by each of
these mode shapes.
3) Integrate the resulting equation over the length of the structure (beam), and
use the orthogonality condition to remove terms that are zero.
4) Solve the resulting ordinary equations of motion as either an assemblage of
single-degree-of-freedom equations or as a system of coupled equations.
The key to the use of Galerkin’s method is that the equations be “reasonably
linear.” Mathematically, this means that the nonlinear terms, if present, can be
expressed as a linear combination of the mode shapes (and modal variables) and that
any one of the resulting modal contributions is small relative to the corresponding
retained linear term. Thus, from a practical standpoint the nonlinear terms are
generally grouped together and put on the right-hand side of the equation as part
of the excitation. The remaining linear terms can either be made to yield a modal
solution or can be approximated by one.
where ωNRj is the jth mode nonrotating blade natural frequency and Kwj is an
additional term to account for centrifugal stiffening and is alternatively denoted
herein as the rise factor or Southwell coefficient for the jth mode.
Then, after the orthogonality principle has been applied a similar modal equation
for in-plane bending can be obtained:
ωk2 = ωNR
2
k
+ 2 Kvk (3.19)
where
Note that in this case Kwk represents the rise factor of the same beam but rotated
so that the motion would be out of plane. Thus, the approximate expression for
the natural (rotating) modal frequency for in-plane bending is the same as the
corresponding one for out-of-plane bending except that the factor multiplying the
rotor speed squared, the rise factor, is reduced in value and can be taken to be that
for the corresponding kth out-of-plane bending mode, but reduced by unity.
Fig. 3.3 Typical Southwell coefficient plot for a rotating beam in bending.
in this section is to look at the limiting cases of the effective Southwell coefficient
at the two extremes of rotor speed variation: = 0 and → ∞. We will actually
look at these cases in reverse order, for simplicity.
Limiting case of the Southwell coefficient for in-plane modes. For the free
vibration in the in-plane modes, the appropriate differential equation, subject to
the same assumptions of spanwise uniformity and very high rotor speed, becomes
ωk 2
2 + 1 ȳk + [(1 − x 2 )ȳk ] = 0 (3.25)
From this equation it can then be inferred that the solution is again expressible
in terms of Legendre polynomials. Thus, the limiting values of the Southwell
coefficient for in-plane motion are given as
k(2k − 1) − 1; cantilevered root condition
Kvk →∞ = (3.26)
(2k + 1)(k + 1) − 1; hinged root condition
1 6 5 1 0
2 15 14 6 5
3 28 27 15 14
4 45 44 28 27
where ωj and γwj are, respectively, the natural frequency and normal mode shape
for the blade at zero rotor speed. The next step is to differentiate Eq. (3.27) with
respect to 2 and then set 2 to zero:
∂γwj ∂γwj ∂ωj2
2 −ωj m 2 + EI
2
−2 mγwj
∂ ∂2 ∂2
→0
− m[(1 − x 2 )γw j ] = 0 (3.28)
Equation (3.28) can then be solved using the Galerkin technique wherein the
equation is first multiplied by the jth mode shape and then integrated over the span
of the blade. The derivative of the mode shape with respect to 2 , ∂γ wj /∂2 can
be written as a linear combination of all of the mode shapes:
∞
∂γwj
= bjm γwm (x) (3.29)
∂2
m=1
Then, when Eq. (3.28) is multiplied by γwj and integrated over the blade span, the
orthogonality principle will leave the first term in braces equal to zero because it
defines the eigenvalue problem for ωj and γwj . The rest of the equation, after the
Galerkin technique has been invoked, becomes
1
γwj (x) 2(Kwj0 )γwj (x) + γw j (x) 1 − x 2 dx = 0 (3.30a)
0
where
∂ωj2
Kwj0 ≡ (3.30b)
∂2
→0
which, after an integration on parts with the second term in Eq. (3.30a), yields the
following simpler form in terms of modal integrals:
1 1
2(Kwj0 ) γwj (x) dx −
2
γw2j (x)(1 − x 2 ) dx = 0 (3.31)
0 0
Again, it can be readily verified that, for both hinged and hingeless rotor blades,
the similar limiting case for in-plane modes is
−mωNR
2
γ + [EIγwj ] = 0
j wj
(3.34)
This equation is linear with constant coefficients and possesses an exact general
solution:
γwj (x) = Aj cos ω̄j x + Bj sin ω̄j x + Cj cosh ω̄j x + Dj sinh ω̄j x (3.35)
where
ω0 = EI/mR4 (3.36b)
Solutions values:
ω̄1 = 3.516015
ω̄2 = 22.034490
ω̄3 = 61.697208
ω̄4 = 120.901922
Constants defining the mode shapes:
ω̄j + sinh ω̄j
sin
Aj = −Cj = − (3.38a)
2(sin ω̄j cosh ω̄j − cos ω̄j sinh ω̄j )
cos ω̄j + cosh ω̄j
Bj = −Dj = (3.38b)
2(sin ω̄j cosh ω̄j − cos ω̄j sinh ω̄j )
where the differentiations are with respect to x̄ (=x/R) and ω̄ and ¯ are,
respectively, the natural frequency and rotor speed nondimensionalized by ω0 .
Zero rotational speed frequencies. Table 3.2 presents the results for zero
rotational speed for hinged and hingeless blades in terms of the nonrotating natural
frequencies and Southwell coefficients for zero rotor speed for in-plane and out-
of-plane motion.
ω̄α2 m = ω̄NR
2 ¯ 2 Kαm (
+ ¯ 2) (3.42)
m
where Kαm can be expressed in terms of the limiting values given in Tables 3.1 and
3.2 and a rise factor variation function ξm , defined as
¯ 2 ) − Kαm
Kαm (
¯ )=
ξm (
2 0
(3.43)
KαmI − Kαm0
The variations of the rise factor variation function with nondimensional rotor speed
squared are given in Figs. 3.4 and 3.5 for hinged and cantilevered root boundary
conditions, respectively. Equation (3.43) can then be rearranged to give the desired
variation of rise factor with rotor speed:
¯ 2 ) = Kαm + ξm (
Kαm ( ¯ 2 )(Kαm − Kαm ) (3.44)
0 I 0
Note that, again, the rise factors for in-plane modes (at any rotor speed) are the
corresponding out-of-plane rise factors minus one. The results of this section are
“exact” in that no simplifying mathematical assumptions were made. The results
of Figs. 3.4 and 3.5 required the use of an appropriate numerical method for
solving Eq. (3.41). Two basic numerical methods for accomplishing this solution
are presented in a subsequent section.
¯ 2 )]
χm = f [ξm ( (3.45)
Fig. 3.4 Rise factor variation functions for first four elastic bending modes, hinged
blades.
The functional variations of χm with the ξm values given in Figs. 3.4 and 3.5
are presented in Figs. 3.6 and 3.7, respectively, for the hinged and hingeless blade
configurations:
The mode shape adjustment factors are then used to “adjust” the nonrotating
mode shapes [i.e., Eq. (3.35)] to produce approximate mode shapes at rotor speed.
Fig. 3.5 Rise factor variation functions for first four elastic bending modes, hingeless
blades.
Fig. 3.6 Mode shape adjustment factors for first four bending modes, hinged blades.
¯ 2) ∼
γm ( ¯ 2 )] γ̄m
= γm0 + χm [ξ( (3.46)
The normalized mode shape adjustment functions are presented in Figs. 3.8 and
3.9 again, respectively, for the hinged and hingeless blade configurations. Note
that in each of these figures the functions all achieve a maximum value of either
Fig. 3.7 Mode shape adjustment factors for first four bending modes, hingeless
blades.
Fig. 3.8 Normalized mode shape adjustment functions for first four bending modes,
hinged blades.
Fig. 3.9 Normalized mode shape adjustment functions for first four bending modes,
hingeless blades.
plus or minus unit value somewhere along the span and are then scaled by the
mode shape adjustment factors. Also, in each of the figures the spanwise vari-
ations of the maximum points over the nondimensional rotor speed range are
indicated.
Example 3.1
Consider a longitudinal string of masses such as that given in Fig. 3.10. Let us
examine in Fig. 3.11 the forces acting on a single spring-mass element. Assuming
oscillating motion (based on a trial value of frequency), we can relate the forces
and displacements across this element:
F1 = m1 ω2 , x1 = 1 (3.51)
The preceding matrix equation defines a transfer matrix between the state con-
ditions at either end of the spring-mass element. This basic equation can then be
used repeatedly, starting with n = 1 and an assumed trial value for ω2 , until the
left-hand side of the structure is reached. The correct value of ω2 will produce
the correct boundary conditions at this end of the structure. In general, any given
trial frequency will be incorrect, and an iteration must be used to reach a suitably
accurate value. Iteration in this case consists of plotting or otherwise “tracking”
the value of the left-hand end boundary condition vs trial values of ω2 , as shown
in Fig. 3.12. Iteration consists of varying the frequency squared value and testing
for a crossing of the frequency squared axis, thereby indicating a zero value.
Fig. 3.11 Free-body diagram for a typical spring-mass element (Example 3.1).
Fig. 3.12 Variation of left-hand end boundary condition with trial frequency
(Example 3.1).
Again, the solution technique is to form a transfer matrix statement relating the
state variables at the node point defined on one beam element to those at the
similarly defined point on the next beam element. Note that the sketch shows
the beam element to be lopsided, with the concentrated mass located at the left-
hand end; this will be discussed in more detail later. For the time being let the
appropriately defined point on the beam element be denoted { }L . Then the transfer
matrix statement we require, with n increasing from right to left, can be written as
The various transfer matrices can then be cascade multiplied to relate the boundary
conditions at the root to those at the tip:
where the required boundary conditions are incorporated in { }LN and { }L1 , as
discussed earlier.
Consider now the makeup of a typical transfer matrix as defined over an arbi-
trary spanwise element. Let us divide the element into two distinct parts: a rigid
concentrated mass-inertia part (associated with the coincident mass mn+1 and iner-
tia In+1 ) and an elastic massless part, each attached to each other at the spanwise
point just to the right of the concentrated mass (denoted { }Rn+1 ), as shown in the
following sketch:
n
Fn = mi xi 2 (3.54a)
i=1
Then, upon balancing the transverse forces and moments on the mass element
and equating the displacements and rotations across the point mass, we obtain the
following recursive relationships:
L
Sn+1 = Sn+1
R
+ mn+1 ω2 zn+1
R
(3.55a)
L
Mn+1 = R
Mn+1 + In+1 ω 2 R
ψn+1 (3.55b)
L
zn+1 = zn+1
R
(3.55c)
L
ψn+1 = R
ψn+1 (3.55d)
Elastic massless part: field transfer matrix. Let us now turn our attention
to the massless elastic portion between { }Rn+1 and { }Ln (see Fig. 3.15).
The equation for elastic deformation of this arbitrary elastic element can be
obtained by considering the elastic deformation characteristics of a simple canti-
levered beam in tension F, of length L, and with constant section properties m
and EI. This beam element is supported at the left-hand end so that it has inboard
boundary conditions of z = z0 and z = z0 , and is loaded at the right-hand end
by the vertical shear and axial loads ST and FT , respectively, and by a bending
moment MT . Upon taking the spanwise variable x to start at the left-hand end,
it can be verified that the differential equation for transverse deformation is then
given by
where zT is the transverse deflection at the tip (right-hand end) of the elastic
element. Solution of this equation is facilitated by a simple change of variable. Let
ζ = x − x0 and η = z − zT , so that
EIη − FT η = ST (L − ζ ) + MT (3.58)
where ζ varies from 0 to L and where the solution η(ζ ) is subject to the following
boundary conditions:
The condition that η(L) = 0 follows automatically from the definition of η. It can
then be verified as an exercise that the solution of this equation is given by the
tanh λL ST L tanh λL
zT = z0 + z0 L + 1−
λL FT λL
MT 1
+ 1− (3.60)
FT cosh λL
1 ST L 1 MT
zT = z0 + 1− + [λL tanh λL] (3.61)
cosh λL FT cosh λL FT
√
where λ = F/EI, and for λL → 0 (zero-tension case), these expressions
simplify to the more elementary expressions for beam deflections:
The elastic transfer function description is then completed by relating the shears
and moments at the two ends to each other:
M0 = MT + ST L − FT (zT − z0 ) (3.64)
S0 = ST (3.65)
Equations (3.60) through (3.65) can then be rearranged and combined into a matrix
format. In doing so, let us rewrite the left-hand and right-hand quantities in terms
of the variables defined in Fig. 3.15. The resulting matrix equation is then given by
R L
z 1 −Ln fn (Ln /Fn )( fn − 1) ( gn − 1)/Fn z
ψ 0
gn −( gn − 1)/Fn
−fn hn ψ
= (3.66)
S
0 0 1 0 S
M n+1 0 −Fn Ln fn Ln fn gn M n
=[Fn ]
where
sinh λL
fn = (3.67a)
λL n
gn = cosh λLn (3.67b)
hn = Ln /EIn (3.67c)
and
λn = Fn /EIn (3.68)
When the preceding results, that is, for both the point and field transfer matrices
[Pn+1 ] and [Fn ], respectively, are combined, the required transfer matrix for the
total (nth) element [Tn ] is obtained:
1 −Ln fn (Ln /Fn )( fn − 1) ( gn − 1)/Fn
0 gn −( gn − 1)/Fn −fn hn
−mn+1 [1 + mn+1 ω2 [mn+1 ω 2
[Tn ] = mn+1 ω2
×ω2 Ln fn ×(Ln /Fn )( fn − 1)] ×(gn − 1)/Fn ]
[In+1 ω2 gn [−In+1 ω2 ( gn − 1)/Fn [−In+1 ω2 fn hn
0 − Fn Ln fn ] + Ln fn ] + gn ]
(3.69)
As described earlier, all of the spanwise transfer matrices are then used in a
cascade multiplication to relate the state conditions at the root of the beam to
those at the tip. Because each of the transfer matrices is frequency dependent, an
eigenvalue problem is thereby defined:
The rest of the eigenvalue problem consists of rewriting the transfer matrix
equation [Eq. (3.70)] utilizing a statement of the boundary conditions that will
result in a standard (matrix) form of an eigenvalue problem. To this end, let us first
review appropriate statements of the boundary conditions:
At the tip:
z1
ψ
1
{ }1 = (3.71a)
0
0
At the root:
0
0
ψ
0
N
{ }N = (3.71b)
SN
SN
0 MN
articulated hingeless
rotor rotor
Then, if these vectors are inserted into the transfer equation and only those
nonzero terms are retained, the following equations are formed:
0 = t11 z1 + t12 ψ1 (3.72a)
(ψN ) = t21 z1 + t22 ψ1 (3.72b)
SN = t31 z1 + t32 ψ1 (3.72c)
(MN ) = t41 z1 + t42 ψ1 (3.72d)
where one or the other (but not both) of the terms in parentheses is set to zero.
Thus, for each type of rotor, articulated as well as hingeless, the explicitly written
matrix equation will contain two scalar equations involving z1 and ψ1 [Eq. (3.72a),
and either Eq. (3.72b) or (3.72d)], each of which is equal to zero.
t11 (ω2 ) t12 (ω2 ) z1 0
2 2
= (3.73)
ti1 (ω ) ti2 (ω ) ψ1 0
(ω2 ) = 0 (3.75)
Once the various roots are known, the mode shape calculations follow in a
convenient manner. Because the various n transfer matrices will have already
been evaluated for the solution root (in the process of obtaining the elements of the
characteristic determinant), they are available to relate the inner element responses
to those at the tip. The steps required to find the mode shape for each root are as
follows:
1) Establish the normalizing condition, that is, set z1 or ψ1 equal to some
reference value (usually 1.0).
2) Solve for the other unknown variable at the tip using t12 ψ1 = −t11 z1 . (Note
that the shear and moment are zero at the tip.) This then completely defines the
state vector at the tip.
3) Use the individual transfer matrices for the interior elements to find the state
vectors at the other spanwise stations.
Clearly, knowledge of the state vectors at the several spanwise stations inher-
ently constitutes a solution for the natural mode shape in terms of displacement and
slope. However, even more importantly, as we shall come to appreciate later, the
state vector also provides knowledge of the shears and moments at the spanwise
stations for unit deflection of the mode.
where δU, δT , and δW are, respectively, the variation of strain energy, the vari-
ation of kinetic energy, and the virtual work done. Generally, these expressions
are subject to considerable diversity among analysts because of the variety of
assumed coordinate systems used for defining the deformed state of the blade and
of assumed coupling effects to be included.
2) For the rotor-blade problem the deformed state of the blade typically consists
of three or four components: a) an out-of-plane displacement component w; b) an
in-plane displacement component v; c) a torsion component φ; and d) occasionally
a radial displacement component u. All of these components are assumed to be
continuous functions of span.
3) The blade structure is discretized into spanwise finite elements, wherein the
boundaries of these elements or spanwise segments are referred to as the nodes.
The values of the deformed state components (i.e., w, v, φ, etc., and their respective
spanwise derivatives) at these nodes are then taken to be the degrees of freedom
of the system (see Fig. 3.16).
4) The distributions of the deflections (w, v, φ, etc.) over each element are then
represented in terms of the nodal displacements using appropriate shape functions.
One common approach to this is to use Hermite interpolation polynomials for the
shape functions. Thus, for the ith spanwise element the deflection components
would be expressed as
w
v = [H]{qi } (3.77)
φ i
Fig. 3.16 Typical identification of nodal variables for a blade finite element modeling.
where i is the length of the ith element and xi is the local axial coordinate for that
element, as measured from the left end of the element.
5) By use of the preceding formulation, the virtual displacements δw, δv, and
δφ over the ith element can then be defined:
δw
δv = [H]{δqi } (3.81)
δφ i
= i =0 (3.82)
where
t2
i = (δUi − δTi − δWi ) dt (3.83)
t1
[δqi ][Mi (qi )]{q̈i } + [δqi ][Ci (qi )]{q̇i } + [δqi ][Ki (qi )]{qi } − [δqi ]{Qi } = i
(3.84)
where [Mi (qi )], [Ci (qi )], and [Ki (qi )] represent the inertia, damping, and stiff-
ness matrices, respectively, and {Qi } represents the element load vector for the ith
element.
7) The global matrices are obtained from the assembly of the element matrices.
We denote the global degree-of-freedom vector as {q} and the global virtual
displacement vector as {δq}. Likewise, we denote the global load vector as {Q}. All
three of these vectors are formed by combining the respective components from
each element. When the components of the statement of Hamilton’s principle
are added together, the following expression results:
Because the virtual displacements {δq} are arbitrary, this equation then leads to
the final desired equations of motion:
8) The matrix equation of motion (3.86) is typically quite banded, and compu-
tational techniques should be employed to eliminate the need for storing elements
that are always equal to zero and to effect the eigensolution of such banded sys-
tems. The method outlined earlier has elements of a collocation technique in that
a solution to a differential equation is obtained at discrete points with approximate
analytic functionality in between. Further, the use of Hamilton’s principle in this
manner defines a generalized Galerkin method of solution wherein the assumed
shape functions replace the use of normal modes in the more conventional use of
the Galerkin method.
Fig. 3.17 Nonuniform beam approximation for use with the Yntema charts (no tip
weight).
the rise factors for the rotating beam in the in-plane and out-of-plane directions.
Similar to the development given earlier in Sec. 3.1.5 and 3.1.6, the formulations
implicit in the Yntema charts include the assumption that the mode shapes are
invariant with rotor speed. Consequently, the results given for rise factor include
no variation with rotor speed.
The quantities an , K0n , and K1n are given in the charts formulated by Yntema for
each of the first three bending modes (n = 1, 2, 3); excerpts of these charts are
presented in Appendix B.
Configuration 2. The beam has constant values of mass and stiffness distri-
butions, but has a variable concentrated tip weight MT , as shown in Fig. 3.18.
This configuration implicitly addresses those blade designs wherein the mass
characteristics are dominated by the tip weight. For this case (using the original
nomenclature) the nonrotating natural frequency is given by
#
EI0
ωNRn = θn2 (3.89)
m0 L 4
Fig. 3.18 Uniform beam approximation for use with the Yntema charts (with tip
weight).
EI equivalencies over short sections of the blade. Over any section of the
blade of total length , wherein the EI distribution is undergoing rapid variations
with span, an approximation can be formulated based on the maintenance of
similar rotation per moment characteristics. To this end, we note that the basic
beam bending equation is given by the following usual form:
M = EIw (3.90)
which over a series of small span lengths i (all within the basic length , wherein
the moment can be expected to have a relatively constant value), the relationship
can be written by equating the sum of the bending rotations to the total angular
rotation:
Mi i M
θ= θi = =M = (3.92)
EIi EIi EIequiv
i i i
Therefore,
$
i i
= ; → EIequiv = $ i (3.93)
EIequiv EIi i (i /EIi )
i
EI distributions over the entire length of the blade. In some cases we require
that an arbitrary EI distribution over the entire blade span be reduced to an equiv-
alent linear variation, as discussed earlier. One basis for calculating an equivalent
linear variation would be to equate the bending strain energy of the real blade result-
ing from the application of a load at the blade tip, with that strain energy resulting
for the blade with the linear EI variation. In effect, this amounts to having both the
real blade and the linear EI equivalent blade have the same tip deflection per load
characteristics. This equivalency condition does not, however, produce a unique
answer in that the linear EI variation blade needs the values of EI at both the
root EI0 and the tip EIt [or, equivalently, a root EI(EI0 ) and a tip-to-root EI ratio,
REI (=EIt /EI0 )]. A further equivalency condition would be to require that both the
real blade and the equivalent blade have the same tip slope per load characteristics.
The following material presents derivations resulting in a final useful equivalency
relationship. The starting point is to solve the basic (nonrotating) beam equation
wherein a linear variation of EI is assumed and a moment distribution is taken
appropriate to a load at the beam tip. Let us further define the spanwise location
in terms of a nondimensional variable η(= x/L) and formulate expressions for the
resulting EI and moment distributions:
Then
% &
M FL (1 − η)
w = = (3.95)
EI EI0 1 − (1 − REI )η
Subsequent integration yields the following expressions for slope and deflection
at the end of the beam (θt and wt , respectively):
FL 2 /EI0
θt = [REI nREI + 1 − REI ] (3.96)
(1 − REI )2
FL 3 /EI0
wt = [−REI
2
nREI + 1/2(1 − 4REI + 3REI
2
)] (3.97)
(1 − REI )3
These two equations can then be combined to give the following expression for
REI in terms of the blade tip deflection and slope:
1/2 − wt
REI = (3.98)
θ t − wt
where
EI0
wt ≡ wt (3.99a)
FL 3
EI0
θ t ≡ θt 2 (3.99b)
FL
The previously stated equivalency condition can also be used to obtain an
expression for the tip value of EI:
1/2 − EI0 wt /FL 3
EIt = (3.100)
θt /FL 2 − wt /FL 3
The use of the preceding equations still does not constitute a unique solution
for EI0 and EIt , as required for using the Yntema charts. For any given application
the (nonrotating) static deflection characteristics of the blade (wt and θt ) would
be known, as well as the length L. However, the equivalency condition stated
earlier only gives REI once a selection for EI0 is made. Note that the appropriate
value of EI0 to be used is not necessarily the EI0 value for the original blade.
Furthermore, although the nondimensional tip slope will generally be greater than
the nondimensional deflection, the range of root EI available is restricted only
to values that result in a finite, positive value for the tip EI. Let us develop this
further. What the engineer needs is some measure of the ratio of tip stiffness to root
stiffness (REI = EI t /EI 0 ) for a blade design that has two fairly distinct spanwise
portions with distinct values of EI. This configuration is depicted in Fig. 3.19,
wherein the inboard (elevated) value of EI is related to the outboard value, EI1 , by
the factor β. This inboard extend is arbitrary and defined by the factor η1 .
Again, we assume a unit tip load and form expressions for the tip deflection and
slope:
' (
EI1 1 η13 1
wt = η1 − η1 +
2
+ (1 − η1 )3 = W (3.101)
(1)L 3 β 3 2
EI1 η1 η1 1
θ t = 1 − + (1 − η1 )2 = (3.102)
(1)L 2 β 2 2
Then, with a rearranging of Eq. (3.99) the preceding expressions can be inserted
to give the required ratio:
η1 ) * 1
β 1 − η1 + η1 3 + 3 (1 − η1 )
2 3
W
REI = = ) 2 * 1
1
2 +W − 1
2 + β −η1 /2 + η1 /3 + 3 (1 − η1 ) − 2 (1 − η1 )
1 3 3 1 2
(3.103)
Figure 3.20 presents the evaluations of this expression for REI for a range of values
of β and a practical variation of η1 :
Fig. 3.20 Equivalent stiffness ratios for use with the Yntema charts.
Example 3.2
Consider the case of a blade (beam) that has, over the inboard third of its length, a
value of EI equal to 5 and, over the outboard 2/3 of its length, a value of EI equal
to 1. Then, assuming a unit tip load F and unit length, the tip deflections can be
readily calculated from simple beam theory to have the following values:
The equivalency equation, Eq. (3.100), expressing EIt in terms of EI0 , gives
Fig. 3.22 Pictorial representation of the coupling of the rotor in-plane bending with
shaft lateral flexibility.
Figure 3.23 presents the pertinent dynamic characteristics that arise from the
coupling of the rotor dynamics with the lateral stiffness of the rotor shaft. A key
characteristic is that the in-phase mode is unstable in divergence for lateral stiff-
nesses below a certain value. Also in evidence is the sharp increase in natural
frequency for the out-of-phase mode with increasing lateral stiffness while the in-
phase mode approaches an asymptotic value. For reasons relating to ground reso-
nance stability, the rotor is generally designed to be subcritical. That is, the resulting
Fig. 3.23 Typical out-of-phase and in-phase frequency variations with lateral stiffness
for the two-bladed rotor (/ω0 = 1).
natural frequency must be greater than the rotor frequency (i.e., /ω < 1).
The converse of subcriticality is supercriticality wherein /ω > 1. Figure 3.24
presents a map of the various dynamic operational conditions as determined by
nondimensional rotor speed and lateral stiffness.
Fig. 3.24 Map of significant dynamic operational conditions for the two-bladed rotor.
The figure clearly demonstrates that the effective lateral stiffness of the shaft
and/or support structure must be sufficiently robust and the blades sufficiently
stiff to ensure that the coupled frequencies remain subcritical. Note that even for
infinite lateral stiffness supercritical operation will occur for (/ω0 )2 ≥ 15.03.
Also of interest is the boundary for the in-phase coupled frequency reaching a
value of twice rotor speed (ω = 2). Such a condition leaves the rotor susceptible
to resonance excitation from the two-per-rev component of the in-plane airloads
in forward flight and should be avoided.
This coordinate system transformation has the advantage that the force boundary
conditions at the tip of the blade (zero moment and shear) are preserved. With these
definitions and concepts the basic equation of motion for the blade in torsion can
then be addressed. Based on the observation already discussed that, because of the
relatively high aspect ratio of rotor blades, the blade elastic torsional deformation
takes on the character of a “space curve,” attention must be paid to the resulting
nonlinearities. The nonlinearities derive from the action of forces (blade loadings)
and moment arms that are provided by blade bending. The detailed derivation of
the resulting basic elastic description for the blade in torsion is beyond the scope
of this text, and the reader is referred to the cited literature. For present purposes
the basic torsion equation, which contains the essential characteristics resulting
from the space-curve character, is presented without derivation as follows:
+
−[GJθ + Tka2 θ + EB1 (θB )2 θ ] (elastic terms)
+
= [qx5 + y5 qy5 + z5 qz5 ] (“distributed” moment loadings)
where px5 , py5 , and pz5 represent the five-coordinate system components of the
force loading distribution, and qx5 , qy5 , and qz5 similarly represent the components
of the moment loading distribution. The Fy and Fz functions are complicated
integral functions of the loadings and deformations and are not given herein for
clarity. The reader is referred to the works of Bielawa and of Loewy, Rosen,
and Mathew for detailed treatments of this approach. Note that the nonlinear
where
km1 , km2 ≡ mass radii of gyration of the blade section about the major (chord-
wise) neutral axis and about an axis perpendicular to that axis and
through the elastic axis, respectively
(note that the polar moment
of inertia is defined by km = km 2 + k2 )
1 m2
GJ ≡ primary (St. Venant) torsional stiffness
T ≡ local tension
ka ≡ polar radius of gyration of the tensile stress carrying area about
the elastic axis (note that this term is of questionable validity for
isotropic materials with circular sections and built-in twist)
EB1 ≡ incremental torsional stiffening (coiled spring effect) = (bending
modulus of elasticity E) × (section constant defined in Houbolt and
Brooks):
ηLE t2
B1 = tη2 η2 + − ka2 dη (3.113)
ηTE 6
θe (0) = 0 (3.114b)
2) The root restraint is zero. This corresponds to a blade that has lost its pitch
control attachment to the swash plate.
3) The blade cross section has a high aspect ratio and thus is quite thin compared
with its chord. This is expressed mathematically as
2
km 2
2
km 1
−→ km
2
≈ km
2
2
(3.116)
One principal conclusion to be drawn from this result is that the natural freq-
uency of the blade for rigid-body torsion is the rotor frequency itself. The
significance of such a value of frequency is that the 1P (1/rev) control loads that
are required to impart cyclic pitching motion to the blade are significantly reduced
because for 1P motion the dynamic system is being driven at its natural frequency
and hence obtains a dynamic attenuation of the resisting (dynamic) control loads.
Thus, the main control forces to be taken up by the control system are the non-
1P propeller moments (principally the steady loads) and all harmonics of the
aerodynamic damping in pitch:
L% &
dM
Mθ = dx ∼ −θ̇ (3.118)
0 dx aero
Note that this aerodynamic damping is not generally small and constitutes a
principal source of the 1P control loads.
where
#
2 − 2 k 2 cos 2θ)
,
m(ω2 km m α
λ= ≡ (3.121a)
GJeff β
2
km = km2
2
− km1
2
(3.121b)
The A and B coefficients are determined by the deflection and torque conditions
at the two ends of the segment. Eliminating these coefficients gives rise to the
following matrix equation:
1
cos λLn sin λLn θ̄e θ̄
βλ = e (3.123)
−βλ sin λLn cos λLn Q n−1
Q n
This equation is then used to cascade multiply from one segment to the next to form
a matrix relationship relating the boundary conditions at the blade root, {**}0 to
those at the tip {**}N :
θ̄e θ̄e θ̄
= [T1 ] [T2 ] · · · [TN ] = [T1N ] e (3.125)
Q 0 Q N Q N
The boundary conditions at the blade tip are simply that the deflection is unity
and that the torque is zero [QN = 0]. Those at the blade root are determined by
the presence of a finite spring rate Kθ :
ωθ2j = ωNR
2
θ
+ Kθj 2 (3.127)
j
My5 = (EIF cos2 + EIE sin2 )w5 + (EIE − EIF ) sin cos v5 (3.128)
Mz5 = (EIE − EIF ) sin cos w5 + (EIF sin2 + EIE cos2 )v5 (3.129)
where EIE and EIF are the bending stiffnesses in the edgewise (or chordwise) and
flatwise directions, respectively.
Fig. 3.29 Typical blade section with arbitrary bending and torsional deformations.
The internal elastic moments defined by Eqs. (3.128) and (3.129) can then
be combined with the inertial and tension loads to produce the following set of
dynamic equations for bending in the two directions:
d
mv̈5 − [v5 T ] − m2 v5 + (EIE − EIF ) sin cos w5
dx
+ (EIF sin2 + EIE cos2 )v5 = Fy5 (x, t) (3.130)
d
mẅ5 − [w T ] + (EIE − EIF ) sin cos v5
dx 5
+ (EIF cos2 + EIE sin2 )v5 = Fz5 (x, t) (3.131)
This type of formulation gives rise to mode shapes and frequencies that are
generally referred to as coupled modes, inasmuch that each mode has both in-plane
and out-of-plane components:
γv (x)
v5 (x, t)
= q (t) (3.132)
w5 (x, t) γw (x) n n
n
The principal difficulty with this type of coupling scheme is that the use of v5 and
w5 as components of blade modal motion implies that they are being represented
by coupled modes that are necessarily determined at some fixed value of the total
pitch angle . Although this requirement is appropriate for the dynamic analysis
of propeller blades, it is untenable for helicopter rotor blades because the latter
must necessarily undergo substantial variations in (cyclic) pitch even in steady
flight conditions.
Nv
ve = γvm (x)qvm (t) (3.146)
m=1
Nw
we = γwi (x)qwi (t) (3.147)
i=1
Fig. 3.30 Typical variation of blade frequencies with the coupling effects of pitch and
twist rate.
Fig. 3.31 Typical variation of blade mode shape components as a result of the
coupling effects of pitch and twist rate—first flatwise mode (1F) at 300 rpm.
of the mass center and the shear center. Both of these parameters produce lin-
ear coupling of the bending degrees of freedom with the elastic torsion modes.
Furthermore, practical design considerations of rotor blades have driven blade
designs to assume long beam-like structures of high aspect ratios (certainly com-
pared to fixed wings). Furthermore blades have relatively low bending stiffnesses
because so much stiffening is available from the inherent tension field. The further
requirement for minimizing rotor weight combines with these structural charac-
teristics to produce conditions wherein the rotor blade can flex in a manner as to
produce couplings not normally present in fixed wings. This situation has gener-
ated an important source of nonlinear coupling wherein blade flatwise bending
together with edgewise bending can produce significant torsion moments. Brief
examinations of both of these types of torsion coupling forms the remainder of
this chapter.
Linear coupling caused by shear center and mass center offsets. The
variability of the shear center location along the span of the blade inherently defines
an elastic axis that is now a space curve rather than a straight line. Although the
ability to analyze such a structural configuration is well within the capabilities of
current finite element analyses, we seek a more direct (albeit approximate) way
of including these effects. The motivation for such an approach is the need to
construct aeroelastic stability and vibration analyses that are sufficiently compact
that we might understand the physics at work and thereby correct unacceptable
dynamic problems in a rotor design when they occur. The thrust in this text is a
strong proclivity to use modal approaches. We will use such an approach to deal
with blades having a spanwise variable elastic axis with mass center offsets.
It can be reasoned that as far as the torsion kinematics are concerned a space-
curve elastic axis can be the result of either being built-in or caused by elastic
bending deformations. Thus, we can use the material from Sec. 3.7.4 as a starting
point: substitute ySC in place of ve and use the elastic torsion twist rate θe as the
requisite twist rate in the expressions for the deflection correction functions. The
displacements of the mass center will then include the mass center offset yCG in
addition to the already present elastic bending ve . With the retention of only the
linear terms in the bending and torsion variables, and using Eqs. (3.148), (3.149),
and (3.151) as a starting point, the v5 and w5 displacements of the mass center can
first be written as
Then, after separating out the elastic torsion from , the following final
expressions for the mass center displacements are obtained:
v5CG = (ve + yCG ) cos θ − [we + (yCG γθj − wSCj )qθj ] sin θ (3.157)
w5CG = (ve + yCG ) sin θ + [we + (yCG γθj − wSCj )qθj ] cos θ (3.158)
where ve and we are given by the usual representation [i.e., Eqs. (3.146) and
(3.147)] and the following definition is made:
x̄ x̄ x̄1
wSCj = γθj ySC dx̄1 + γθj ySC
dx̄2 dx̄1 (3.159)
0 0 0
The displacements of the mass center as defined by Eqs. (3.157) and (3.158)
are depicted in Fig. 3.32. The elastic bending displacements have been omitted
for clarity. Note that in this development the reference axis is assumed to coincide
with the elastic axis at the blade root. This is not unreasonable given the symmetry
of the pitch bearing and cuff assembly at the root of the blade.
Fig. 3.33 Pictorial representation of the nonlinear torsion moment resulting from the
combinatorial action of edgewise bending and flatwise bending.
linearization that occurs about the steady deflected values of both components of
bending. Minimizing the impact of this coupling typically requires stiff torsion
blade designs. Alternatively, the coupling has been used to provide stability
augmentation for some forms of aeroelastic instability phenomena.
References
Section
3.1 Bisplinghoff, R. L., Ashley, H., and Halfman, R. L., Aeroelasticity,
Addison–Wesley, Cambridge, MA, 1955.
Bramwell, A. R. S., Helicopter Dynamics, Wiley, New York, 1976.
Johnson, W., Helicopter Theory, Princeton Univ. Press, Princeton, NJ,
1980.
Payne, P. R., Helicopter Dynamics and Aerodynamics, Macmillan,
New York, 1959.
Rao, S. S., Mechanical Vibrations, Addison–Wesley, Reading, MA,
1995.
3.2 Friedmann, P. P., and Straub, F., “Application of the Finite Element
Method to Rotary-Wing Aeroelasticity,” Journal of the American
Helicopter Society, Vol. 25, No. 1, 1980, pp. 36–44.
Hurty, W. C., and Rubinstein, M. F., Dynamics of Structures, Prentice–
Hall, Englewood Cliffs, NJ, 1964.
Myklestad, N. O., Vibration Analysis, McGraw–Hill, New York, 1944.
3.3 Bailey, C. D., “Exact and Direct Analytical Solutions to Vibrating Sys-
tems with Discontinuities,” Journal of Sound and Vibration, Vol. 44,
No. 1, 1976, pp. 15–25.
References (continued)
Hodges, D. H., “Vibration and Response of Nonuniform Rotating
Beams with Discontinuities,” Journal of the American Helicopter
Society, Vol. 24, No. 5, 1979, pp. 43–50.
3.4 Payne, P. R., Helicopter Dynamics and Aerodynamics, Macmillan,
New York, 1959.
Yntema, R. T., “Simplified Procedures and Charts for the Rapid Estima-
tion of Bending Frequencies of Rotating Beams,” NACA TN-3459,
June 1955.
3.5 Hodges, D. H., and Dowell, E. H., “Nonlinear Equations of Motion
for the Elastic Bending and Torison of Twisted Nonuniform Rotor
Blades,” NASA TN D-7818, 1974.
Houbolt, J. C., and Brooks, G. W., “Differential Equations of Motion
for Combined Flapwise Bending, Chordwise Bending and Torsion
of Twisted Non-uniform Rotor Blades,” NACA Rept. 1346, 1958.
3.6 Bielawa, R. L., Johnson, S. A., Chi, R. M., and Gangwani, S. T.,
“Aeroelastic Analysis for Propellers,” NASA CR-3729, 1983.
Loewy, R. G., Rosen, A., and Mathew, M. B., “Application of the Prin-
cipal Curvature Transformation to Nonlinear Rotor Blade Analysis,”
Vertica, Vol. 11, No. 1/2, 1987, pp. 263–296.
Problems
3.1 A helicopter rotor blade has been found to have nonrotating first flatwise and
edgewise mode frequencies of 1.6155 and 4.8864 Hz, respectively. Addi-
tionally, it is known that the first flatwise mode frequency (at a nominal rotor
rotation speed of 5 Hz) is 5.6 Hz and that the first edgewise mode has a
resonant 2P frequency at = 2.4903 Hz.
(a) Estimate the variations of the modal frequencies in the rotor speed range
of nominal ±10%.
(b) Are these frequency characteristics acceptable from a structural dynamic
point of view? Substantiate your answer.
3.2 Consider a massless rotating cantilevered beam of length L with a constant
bending stiffness distribution EI and a concentrated mass M attached at its
free end. The beam rotates about an axis passing through its point of cantilever
attachment with a constant rotor speed thereby imparting a constant tension
T to the beam.
(a) Show that its (out-of-plane) natural frequency is given by
tanh λL −1/2
ω = T /M L − (3.160)
λ
where
λ= T /EI (3.161)
3.3 Derive the expressions for displacement and slope transfer matrix elements
for the elastic beam segment given in Sec. 3.3.1 (p. 83), that is,
L
z
z R
=[ ? ]
S
n+1
M n
3.4 Consider a model (articulated) rotor blade that has a total rotor radius of
56.224 in, a flapping hinge of 3 in, and a rotor speed of 69.32 rad/s. The
geometry of this blade and the distributions of the blade’s structural properties
are given, respectively, in Fig. 3.34 and Table 3.3.
Using the Yntema charts, obtain estimates of the first three flatwise and
first edgewise natural frequencies.
3.5 As discussed in Sec. 3.6.2, because of flexibilities in the rotor control system
the blade torsion modes can typically have finite values at the blade root
(x = 0). One way of approximating these root values and the resulting
frequencies and mode shapes from the modal characteristics obtained with
the rigid root boundary condition is to use the Galerkin method. Let the
normal mode shapes and natural frequencies of the blade with the rigid root
boundary condition be denoted by γθn and ωθn , respectively. We wish to
obtain the similar characteristics for the flexible root condition γ̂θn and ω̂θn ,
where it is assumed that the two mode shapes can be related to each other by
the following relationship:
θ (x, t) = [cn γ̂θn (x)] exp(iω̂θn t) (3.162)
n
where
cn γ̂θn (x) = (1)a0n + ajn γθj (x) (3.163)
j
The objective is to calculate the a0n and ajn coefficients for the nth mode.
Mathematically, this objective defines an eigenvalue problem. Note that the
γθj (x) mode shapes are orthogonal to each other but not to the pseudotorsion
mode γθ0 (x)(= 1). These constants can be related to the root torsion flexibil-
ity (spring stiffness constant) Kθ by relating each of the coefficients to the
magnitude of the vibratory root torsion moment:
where
Table 3.3 Tabulation of structural properties for model S-76 rotor blade
3.6 Consider a massless cantilevered beam of length L with a tip mass of weight
W . The beam has a constant high-aspect-ratio rectangular cross section so
that it can be assumed to have infinite stiffness in the direction normal to
the span; thus, it can be characterized by a single (finite) flatwise bending
stiffness EI. Furthermore, the beam has a linear built-in twist and is mounted
so that 1) at the root section, the chord (widthwise, i.e., edgewise dimension
of the beam) has a pitch angle θ0 , which for present purposes can be assumed
to be “small”; and 2) the tip section chord pitch angle is zero [i.e., (0) = θ0 ,
and (L) = 0]. Assuming the beam to be infinitely rigid in torsion, calculate
approximate tip deflections of the beam in the vertical z5 and horizontal
y5 directions because of the weight of the tip mass. Express the component
deflections in terms of the beam length L, the nondimensional factor WL 2 /EI,
and the root pitch angle θ0 .
where pi is the time derivative with respect to the (fixed) inertial coordinate system.
We can then take the cross product of RCk with this equation where RCk is the
vector from the mass center of the solid body to the kth mass particle, noting
that
After taking the cross product and summing over all of the particles, we obtain
pi (RCk × mk pi RCk ) = RCk × F(ext)k (4.3)
k k
= Hc = Mc
where H c is the angular momentum of the system of particles about its mass center.
The equation of dynamic equilibrium then becomes
pi H c = M c (4.4)
127
Therefore,
Hc = mk RCk × (ωib × RCk ) (4.6)
k
and using the formula for the vector triple product yields
Hc = mk [(RCk · RCk ) ωib − (RCk · ωib )RCk ] (4.7)
k
From this expression it can be concluded that the angular momentum and angular
velocity H c and ωib , respectively, are not generally aligned in the same direc-
tion. A more tractable form of the angular momentum vector can be obtained by
expanding out the general expression for H c in terms of the components of the
particle position and the angular velocity. Let
xk
{RCk } = yk (4.8)
zk
and
ωx
measured about
{ωib } = ωy ←− (4.9)
body axes
ωz
These expressions are then substituted into the definition for H c , and the summation
is taken. The result of this summation can then be expressed by the following matrix
equation for the body-fixed components of H c :
Hx Ixx −Ixy −Ixz ωx
Hy = −Iyx Iyy −Iyz ωy (4.10)
Hz −Izx −Izy Izz ωz
where Ixx , Iyy , and Izz are the mass moments of inertia of the rigid body about
arbitrarily defined (orthogonal) x, y, and z axes of the body, respectively. The
off-diagonal terms are the products of inertia, for example,
Ixy = mk xk yk (4.11a)
k
Iyz = mk yk zk (4.11b)
k
Izx = mk zk xk (4.11c)
k
If the principal axes are chosen as the body-fixed coordinate frame, the products
of inertia vanish. Thus, if the resulting principal axis moments of inertia are then
denoted with a single subscript, the angular momentum vector can be expressed
in the following simplified form:
H c = Ix ωx i + Iy ωy j + Iz ωz k (4.12)
This result can then be used with an invocation of the theorem of Coriolis to obtain
the desired more tractable form of the dynamic equilibrium equation:
pb H c + ωib × H c = M c (4.13)
From this vector differential equation the following set of equations (Euler’s equa-
tions for rotational motion of a rigid solid, as expressed in body-fixed coordinates)
can then be written:
These equations are highly nonlinear, and a general closed-form solution does
not exist. Successful solutions usually involve a suitable linearization appropriate
to the problem at hand. The following section, dealing with simplified general
gyroscopic motion, presents development for just such a case where linearization
is possible. In preparation of this development, Fig. 4.2 presents a summary of the
use of vector representations of rotational quantities.
pH ns + ωib × H s = M (4.17)
This equation is then the simplified gyroscope equation. The most signifi-
cant characteristic of this equation is that it is linear in the components of the
angular velocities defining the nonspin angular momentum. The significance of
the three terms of this equation is described as follows: The variable pH ns is the
source of the characteristic dynamics or transient response of the gyro and gives
rise to nutation in two-degrees-of-freedom gyros. The expression ωib × H s is the
source of the precessional characteristics of the gyro. The variable M is the applied
moment defined in terms of body-fixed coordinates (unit vectors imbedded in the
nonspin body).
Simple direction rule: A gyro rotor precesses in such a way so as to align the spin
momentum vector with the applied moment M.
Euler’s equation. Euler’s equations are defined only in an axis system embed-
ded in the rotating solid. Therefore, let the XYZ axis system be attached to the
rotating gyro element, but defined in terms of the rotation rates θ̇x and θ̇y , describing
the motion of the disk as a moving but not spinning solid:
Entering the preceding expressions into the simplified gyro equation yields
from which the same equations can be obtained as those with Euler’s equations.
These equations are general and quite basic. They constitute a fundamental
analytical tool for the inclusion of gyroscopic effects in any given modeling
problem. As demonstrated earlier, these equations are obtainable by more than
one method and thus should appear in the formulation with whatever method
might be used.
These equations are clearly linear coupled equations, and solutions can be obtained
using any of a variety of standard methods. In general, the complete solution
consists of the sum of a homogeneous solution and a particular solution. The form
of the particular solution depends on the specific form of the applied moments
Mx and My . The homogeneous solution depends on the roots of the characteristic
equation, which results from setting the characteristic determinant equal to zero.
Consider first, however, the particular solution that can be obtained from setting
the derivative terms equal to zero and assuming some specific functionality to the
applied moments Mx and My :
1
ωy = Mx (4.28a)
J
−1
ωx = My (4.28b)
J
The first important characteristic of gyroscopic elements can be readily seen from
setting the components of applied moment equal to zero: In the absence of applied
moments the gyroscopic element retains its orientation in space; that is, the compo-
nents of (nonspin) angular velocity are zero. The second important characteristic
of gyroscopic elements is that of precession, wherein the components of non-
spin angular velocity are equal to their respective opposite components of applied
moment in such a way that the angular momentum vector of the gyroscopic element
tends to align itself with the applied moment.
or, recalling that the polar moment of inertia J is ideally twice the diametral
moment of inertia I, gives
λ2 = −42 (4.30)
λ = ± i2 (4.31)
This resulting oscillatory root pair then defines the nutation frequency ωN of the
element:
ωN = 2 (4.32)
H s = Jk (4.33)
where Iˆ and J are, respectively, mass moments of inertia about axes through the
pivot point transverse to and along the spin axis.
Applied (gravity) moment:
Note that, for this configuration, Iˆ = Id + m2 , where Id is the diametral moment
of inertia of the gyro disk. Application of the simplified gyroscope equations and
the linearization of the equations about a constant value of the inclination angle θ0
then yields the following two differential equations of motion:
ˆ
−I(sin θ0 ψ̄¨ + z cos θ0 θ̄)
¨ + Jθ̄˙ = 0 (4.37a)
Deleting the transient terms in the second of these equations and factoring out the
common sin θ0 term from the equation gives a quadratic relationship defining the
equilibrium inclination angle θ0 in terms of the precession angular velocity z :
ˆ 2z cos θ0
z J = mg + I (4.38)
Î mg z
η≡ ; ≡ ; ¯z ≡
(4.39)
J J2
Then the preceding equation can be solved for the two equilibrium precessional
rates possible for the top:
1
¯z =
1 ± 1 − 4η cos θ0 (4.40)
2η cos θ0
Alternatively, the cosine of the inclination angle can be expressed in terms of the
equilibrium precessional rate:
¯ z − )/η
cos θ0 = ( ¯ 2z (4.41)
¯ 2z → 0; hence,
Note that for high spin rates, cos θ0
¯ z = = mg/J2
(4.42)
It can be verified that the solution for λ yields two zero values and an imaginary
pair. The zero pair of roots correspond to the precessional characteristics of the
top, whereas the imaginary pair gives the final desired expression for the (non-
dimensional) nutation frequency ω̄N :
1
ω̄N = 1 − η cos θ0 (4.46)
η
where η is the ratio of moments of inertia, as defined earlier. Note that in the limit
as the pivot to c.g. distance goes to zero, the top becomes a simple gyroscopic
element as described in the preceding subsection. For this case the η parameter
goes to 21 , and the parameter goes to zero. These values then produce a nutation
frequency equal to two times the rotation frequency , as expected.
where
and where βc and βs represent the fore and aft and lateral tilt angles, respectively,
of the rotor disk and are arbitrary functions of time. Indeed, as material in sub-
sequent chapters will show, aeromechanical formulations involving rotor mode
descriptions of both flapping and edgewise motions of the blade will generate
gyroscopic-like, skew-symmetric terms in the equations of motion.
Then assume that the rotor disk as defined earlier is now undergoing sinusoidal
motion, so that the two tilt angles βc and βs are expressed as
Finally, let us assume for simplicity that the amplitudes of these two angles are
equal to each other, but that their relative phase angle φ is different:
From a physical point of view, β̄ represents the amount of tilt that one would
see. These concepts can be further visualized by considering the motion of the
head of an arbitrary constant amplitude vector maintained normal to the disk N.
As a top view (see Fig. 4.6) of the rotor shows, positive values of β̄c and β̄s
produce component projections of this vector, respectively, in the forward x and
lateral (to port) y directions, where the x-y plane represents the undeflected rotor
plane.
If now the first time derivative of βs is related to βc and a value of time is
selected in order to make βc have a purely real positive value, then, depending on
the value of the relative phase angle φ, the two-dimensional motion of the head of
the vector can be determined:
The relationship between βc , βs , and β̇s can then be examined for different assumed
values of φ:
Case 1 (φ = 0): For this case the two-component projection vectors are not
only in phase, but are equal to each other. Thus, the motion of the head of the
vector would be that of a degenerate ellipse (line segment) with semimajor axis
equal to β̄. This type of motion could be described as a back and forth “wobbling”
of the disk:
Case 2 (φ = −π/2): For this case the real-valued components of βs and β̇s
would be zero and positive respectively; hence, the βs component of the N vector
would be moving in the negative y direction. The head of the vector would therefore
describe a circular motion in the same direction as the rotor rotation. This type of
motion is called progressive or forward whirl:
Case 3 (φ = π/2): For this case the real-valued components of βs and β̇s
would be zero and negative respectively, and the motion of the head of the N
vector would be again circular, but in a direction opposite to the rotor rotation.
This type of motion is called regressive whirl:
β̄
βj (t) = Re{exp[i(θj + ωt)] − i exp[i(θj + ωt)]eiφ
2
+ exp[i(−θj + ωt)] + i exp[i(−θj + ωt)]eiφ } (4.53)
β̄
βj (t) = Re{(1 + sin φ − i cos φ) exp[i(θj + ωt)]
2
+ (1 − sin φ + i cos φ) exp[i(−θj + ωt)]} (4.54)
Let us now examine this equation for each of the two types of whirl. First, consider
the case of progressive whirl (φ = −π/2). For this case the preceding equation
simplifies to the following expression:
β̄
βj (t) = Re{2 exp[i(−θj + ωt)]}
2
= β̄ cos[(ω − )t − ψj ] (4.55)
As conceptualized earlier, the rotor disk tilt angles are measured in the fixed or
nonrotating coordinated system. Therefore, the oscillation frequency ω represents
a fixed coordinate system frequency and is what would be seen by the nonrotating
observer. In the following material this frequency will be denoted by a subscript
F. Thus, the whirling motion (either progressive or regressive) of the N vector
or, more specifically, the rotor disk itself would be observed as occurring with a
circular frequency of this value.
Note, however, that the blade oscillating in the rotating coordinate system would
still be governed by its natural frequency, as defined in the rotating frame ωR :
Thus, after the frequencies in these two expressions are equated, the following
frequency relationship emerges:
ωF − = ωR −→ ωF = ωR + (4.57)
Following the same argument for regressive whirl (φ = π/2), a similar frequency
relationship can be deduced:
Note: For this case the fixed coordinate frequency (as expressed by this equation)
can reach zero or even negative values. The interpretation to be given to these cases
is based on the expressions given earlier for relating βc to β̇s . As the fixed coordinate
system frequency approaches zero, the observed whirling motion comes to a halt
and the rotor assumes a steady deflected position. As the calculated frequency goes
negative (as would most likely be the case with increased rotor speed), the whirling
motion would begin again, but in the advancing direction. Thus, for this condition
the rotor would demonstrate two advancing whirl modes; one would be the original
progressive whirl mode, and another, at a significantly lower frequency, would be
what originally was the regressive whirl mode. In the preceding argument the
concept of a “negative frequency” can be used only in the context of a complex-
valued representation for the assumed sinusoidal motion. In reality only positive
values of frequency can be observed: the use of a negative frequency is justified
only for defining the direction of the whirling motion.
It is through the concept of whirl modes that the third identifiable gyros-
copic characteristic of rotor blades, that is, nutation, can be defined. For the case
of rigid blade flapping (in the limit as the offset approaches zero), the (rotating)
blade natural frequency is 1/rev, and the two resulting fixed system whirl mode
frequencies for the rotor are (by the previously given equations) zero for the
regressive mode (precession) and 2/rev for the progressive mode (nutation). Thus,
generalization leads to assigning the progressive mode as the nutation equivalent
characteristic of the rotor.
βj (t) = βc(m) (t) cos m(t + ψj ) + βs(m) (t) sin m(t + ψj ) (4.59)
References
Section
4.1 Scarborough, J. B., The Gyroscope, Theory and Applications, Inter-
science, New York, 1958.
Wrigley, W., Hollister, W. M., and Denhard, W. G., Gyroscopic Theory,
Design and Instrumentation, MIT Press, Cambridge, MA, 1969.
4.2 Wrigley, W., Hollister, W. M., and Denhard, W. G., Gyroscopic Theory,
Design and Instrumentation, MIT Press, Cambridge, MA, 1969.
4.3 Scarborough, J. B., The Gyroscope, Theory and Applications, Inter-
science, New York, 1958.
Wrigley, W., Hollister, W. M., and Denhard, W. G., Gyroscopic Theory,
Design and Instrumentation, MIT Press, Cambridge, MA, 1969.
4.4 Hohenemser, K. H., and Yin, S.-K., “Some Applications of the Method
of Multiblade Coordinates,” Journal of the American Helicopter
Society, Vol. 17, No. 3, 1972, pp. 3–12.
Problems
4.1 An element of good tail rotor design is providing adequate clearance for
teetering (or flapping, as appropriate, depending on the number of blades)
angle responses caused by yawing maneuvers in hovering flight. Consider
the spinning tail rotor to be a rigid disk (gyroscopic element) as shown in
the following:
Mx = M1 α − M2 β̇
My = M1 β + M2 (α̇ + ψ̇)
where M1 and M2 are the effective aerodynamic spring and damping
rates, respectively, as available using rotor aerodynamic (strip) theory or
experimental evaluations.
(b) Find the steady-state teetering angle components ᾱ and β̄ caused by
a constant helicopter yaw rate ψ̇, equal to 100 deg/s, for a tail rotor with
the following specifications:
4.2 Consider the modal differential equation for the first elastic flapwise bending
mode of a main rotor blade (as defined in the rotating coordinate system):
R
Iβ [β̈1 + (ωNR
2
1
+ Kβ1 2 )β1 ] = β1 (t) = 0 γβ1 (x)p(x, t) dx
where β1 is the generalized forcing function for this mode caused by
perturbational load distributions [as arising from either aerodynamic (control
inputs) and/or inertial load considerations].
(a) By representing both β and β1 as Fourier series, each truncated
after only the first harmonic components, [β1c (t), β1s (t), 1c (t), and 1s (t)],
that is,
β(t) = β1c (t) cos t + β1s (t) sin t
β1 (t) = 1c (t) cos t + 1s (t) sin t
Derive two differential equations for the Fourier components of flapping
motion β1c and β1s .
(b) Compare the equations obtained in part 1 with the equations obtained
for the basic gyroscopic element; in particular, explain differences and/or
additional terms obtained.
4.4 Consider the design of a helicopter utilizing an advancing blade concept rotor
system. This rotor system is characterized by having two counter-rotating,
coaxial rotors, which are separated from each other vertically by a distance
Z and which are comprised of “stiff” hingeless rotor blades. The rotor
system must be designed to accommodate hard landings wherein rapid nose-
down pitch rates can occur as a result of a hard initial impact on the tail
landing gear.
Using the results of problem 4.3, determine the minimum safe rotor ver-
tical spacing Z needed to preclude the rotors from striking each other
due to the precession induced by a hard landing pitch rate of −200 deg/s
(nosedown). Assume 1) that the rotors respond in flapping quasistatically,
wherein the nutational dynamics (transients) are disregarded, and 2) that the
aerodynamic damping and stiffness are not factors.
The operational data are given:
1) The upper rotor rotates in the American sense (counterclockwise
when viewed from above) and thereby has a positive right-hand angular
momentum, whereas the lower rotor has a negative angular momentum.
2) The blade properties are as follows:
Rotor radius:
R = 30 ft
Rotor speed:
= 20 rad/s
First blade flapping (rotating) natural frequency:
ωβ1 = 1.6P
First flapping mode generalized inertia:
1
L 3 mγβ21 dx = Iβ1 = 1398 lb-ft-s2
0
First flapping mode coupling inertia:
1
L 3
mγβ1 x dx = Sβ1 = 1677 lb-ft-s2
0
Assumptions:
1) The fuselage has zero compliance so that the wing root is a “ground”
point for the cantilever wing mount.
2) The nacelle is joined to the wing tip by means of a roll spring Ky , which
couples with the wing in bending, and a pitch spring Kθ , which couples with
the wing in torsion.
3) The wing elastomechanical properties can be expressed in terms of
generalized masses and natural frequencies.
4) The mass of the nacelle and rotor are included in the wing generalized
mass.
5) The rotor is rigid and acts as a gyroscopic element, whose “spin”
angular momentum is given by J.
6) The nacelle has mass mN , located a distance h1 (vertically) from the
wing tip attachment point, and a moment of inertia about that c.g., IN , both
in pitch and roll.
7) The rotor has mass mR , located at the hub a distance h2 from the wing
tip attachment point, and a diametral moment of inertia IR .
(a) Derive the four differential equations of motion for the dynamic system
thus defined.
Fig. 5.1 Jeffcott rotor on rigid bearing supports (after Ehrich, 1992).
Mass eccentricity. For this case the elastic axis is considered to be initially
straight but flexible in bending, so as to provide effective springs Kr in the x and y
directions, as shown in Fig. 5.2. Let
G ≡ center of gravity of the disk
b ≡ geometric center of the disk
O ≡ center of rotation
Kr ≡ elastic restoring spring rate of the shaft
Br ≡ equivalent viscous damping (air resistance and bearing damping)
Fig. 5.2 Geometry of a flexible shaft elastically deformed in the transverse direction
as a result of an unbalanced attached disk.
The forced motion dynamic equations for the shaft/disk system are
Initial shaft curvature. For purposes of analysis, we assume the mass center
of the disk to be coincident with the elastic center (i.e., point G is coincident with
point P, with e = 0), but the disk center now has an initial offset from the straight
line point O. For this case the whirl amplification factor wb2 is given by
u 1
wb2 = = (5.4)
s (1 − r )2 + 4ζ 2 r 2
2
Fig. 5.3 Response caused by mass unbalance vs rotor speed for the Jeffcott rotor on
rigid bearing supports.
This function is graphed in Fig. 5.4. A comparison of Fig. 5.3 with Fig. 5.4
shows the following:
1) The two types of unbalance have completely opposite behavior at rotor speeds
both above and below the resonant frequency point, r = /ωn . For rotor speeds
below the resonant frequency, the mass unbalanced rotor has the advantage of
lower response levels, whereas the initial curvature rotor has the advantage at
the higher rotor speeds.
2) At the resonance point the response levels are essentially the same for the two
cases and demonstrate the typical results for all resonant systems.
u σ2
wa = = (5.5)
e [1 − (1 + κ)σ 2 ]2 + 4σ 2 ζs2 (1 − σ 2 κ)2
Fig. 5.4 Response as a result of initial curvature vs rotor speed for the Jeffcott rotor
on rigid bearing supports.
Fig. 5.5 Jeffcott rotor on flexible bearing supports (after Ehrich, 1992).
where σ is a new speed ratio defined by the natural frequency for a rigid rotor
(Kr = ∞, i.e., all of the stiffness accrues from the support bearings):
σ = =√ (5.6)
µ Ks /M
and an appropriately defined new damping ratio η is determined by the support
dampers Bs :
Bs Bs
ζs = = √ (5.7)
2Mµ 2 MKs
A flexibility ratio κ is defined as
Ks
κ= (5.8)
Kr
The whirl amplification function for the center of the disk, wb is then given by
wb = wa (1 + κ)2 + (2σ ζs κ)2 (5.9)
Fig. 5.7 Speed ratio at peak response vs damping—Jeffcott rotor on flexible supports
(after Ehrich, 1992).
designs. Furthermore, the trend has been to operate at higher shaft speeds, making
subcritical designs more difficult.
can be summarized as follows: For the kth lateral mode to which the shaft speed
is supercritical, the motion will be stable provided
ζ2 (ζint + ζext )
= rk < = (5.11)
ωk ζ1 ζint
where
ωk ≡ natural bending frequency of the kth lateral bending mode
ζ1 ≡ damping ratio of the shaft in the lateral bending mode without external
damping
ζ2 ≡ damping ratio of the shaft in the lateral bending mode with external damping
From basic elasticity relationships for a beam, the following influence coefficient
matrix is obtained:
3
δ L /3EI L 2 /2EI F
= 2 (5.13)
φ L /2EI L/EI M
When these two sets of equations are combined, the following equation set for δ
and φ is obtained:
where R is the frequency with the disk effect divided by the frequency without the
disk effect, and
D ≡ Id /mL 2 (5.17)
This result, together with the results for other similar configurations, is pre-
sented in Fig. 5.9. Note that each of the curves is asymptotic to the value that
Fig. 5.9 Change in the natural frequency of a rotating shaft caused by the
pseudogyroscopic effect of a disk.
would be obtained if the disk had remained in its initial plane, thereby providing
a zeroslope restraint to the lateral bending of the shaft.
and
ui+1
v
i+1
{qi+1 } = (5.18)
α
i+1
βi+1
These vectors also define the states at the i and i + 1 ends of the elastic seg-
ment connecting the adjoining disks. A complete elastic description of the segment
makes use of the pictorial representations given in Figs. 5.10a and 5.10b, which
show the transverse deformations in the x − z and y − z planes, respectively:
a) bending in the x − z plane and b) bending in the y − z plane.
The force-displacement for the connecting beam segment is then given by
{Qi } {qi }
= k (i)
(5.19)
{Qi+1 } {qi+1 }
Fig. 5.10 Transverse deformation of a flexible shaft segment (after Ehrich, 1992).
where
12 0 0 6Li
EIi 0 12 −6Li 0
(i)
ki,i = 3 (5.21)
Li (1 + 12εi ) 0 −6Li 4Li (1 + 3εi )
2 0
6Li 0 0 4Li (1 + 3εi )
2
−12 0 0 6Li
T EIi 0 −12 −6Li 0
(i) (i)
ki,i+1 = ki+1,i = 3
Li (1 + 12εi ) 0 6Li 2Li (1 − 6εi )
2 0
−6Li 0 0 2Li (1 − 6εi )
2
(5.22)
Putting all of the pieces together then produces the following matrix equation
for the ith disk:
(i) (i)
[mi ]{q̈i } + [gi ]{q̇i } + ki,i {qi } − ki,i+1 {qi+1 }
(i−1) (i−1)
− ki,i−1 {qi−1 } + ki,i {qi } = 0 (5.25)
where the translational and rotational stiffness coefficients are respectively given by
6EI 6EI
kT = ; kR = (5.26)
(1 + 3ε)L 3 (1 + 3ε)L
Fig. 5.11 Pictorial representation of the Jeffcott rotor with gyroscopic dynamics and
a comprehensive elastic description of the shaft segments (after Ehrich, 1992).
The equation set defines four imaginary paired eigenvalues, that is, four
frequencies:
kT J J 2 kR
ω1 , ω2 = ± and ω3 , ω4 = ± +
M 2Id 2Id Id
Fig. 5.12 Jeffcott rotor whirl speeds and modes and shaft critical speeds (after Ehrich,
1992).
where
For shafts with varying sections, one technique is to form equivalent lengths
of reference diameter shaft (D). These lengths are added together to form a single
constant section length of shaft (with diameter D), and the formula given earlier
is then used. Thus, for basic shaft types (of length L1 ) the equivalent lengths are
given as follows:
L = L1 [D/D1 ]4 (5.28a)
Hollow cylindrical shaft of inner diameter d1 :
L1 D4
L= [1/d13 − 1/D13 ] (5.28c)
3(D1 − d1 )
Cylinder of material G1 (different from the reference material G ):
Inertias:
J/Jj , nj2 (5.31)
Stiffnesses:
K/Kj , nj2 (5.32)
where
Nj ≡ rpm of jth shaft (original system)
N ≡ rpm of all shafts (equivalent system)
See Fig. 5.14 for an example.
L1 = J2 L/(J1 + J2 ) (5.34)
will again be two natural frequencies. The appropriate characteristic equation can
be written using the results of the preceding three-inertia case by taking the limit
as J1 approaches ∞:
J2 J3 J2 J3 J3
ω4 − + + ω2 + 1 = 0 (5.36)
K1 K2 K1 K1 K2
Shafts with significant mass. In some cases the moment of inertia of the
shaft system resides principally in the shaft itself, in which case either an appro-
priate simplified technique must be used, or recourse must be made to the use of
higher degree-of-freedom techniques such as those considered in subsequent sec-
tions. Consider the (three-) inertia system given in Fig. 5.16, which is comprised
of the two concentrated inertias at the ends J1 and J2 , as well as the polar moment
of inertia of the connecting shaft J0 .
Make the following definitions:
G
a2 ≡ (5.37)
ρ
The a parameter can be seen to be the propagation speed of shear waves in the shaft
material. The dynamics of the shaft (with inertia) is then defined by an appropriate
differential equation. The mating of this shaft with discrete inertias at the ends of
the shaft is accomplished using appropriate boundary conditions:
Timoshenko presents the basic classical solution technique to this problem in the
form of the solution βj of the following transcendental equation:
tan βj aβj
(mn) [βj tan βj ] − = m + n; ωj = = J2 ω2 θ(L) (5.40)
βj LGJθ0 (L)
where the solution for jth natural frequency ωj is thus given in terms of βj . Note
that more than one solution exists to the aforementioned transcendental equation.
Because the equation represents the dynamics of a distributed inertia system with-
out discretization, it therefore contains the information for all of the natural modes.
Consequently, the various solutions obtainable represent the modal frequencies of
the system.
Solution using numerical iteration. Apart from the tedium and inconvenience
of iteratively extracting a solution to Eq. (5.41) from tabulated data, the method is
inherently limited in the number of modes it can extract. As plotted in Fig. 5.17, the
functions are given only for abscissa values extending from zero to π . Although
the tan function is periodic with period π , the [tan x/x] and [x tan x] functions are
not. Thus, if the tabulated functions are available only from zero to π , only the first
mode can be extracted with the use of tabulated values. For shaft configurations
that are relatively short, this might be sufficient, but for longer ones, wherein
higher modes would still be significant, an alternate approach is warranted. One
method would be to modify Eq. (5.41) to admit the higher modes and solve the
modified equation in some fashion. To this end, we first rewrite the eigenvalue
problem in a somewhat different form, recognizing that the tan function is repetitive
with a period of π . The solution β can therefore be expressed as a sum of a
periodic part kπ and a principal-value part β̃ (i.e., the total solution is given by
β = kπ + β̃). The basic transcendental equation is then recast as a function of β̃
and put in the standard form for any of a variety of numerical iteration solution
schemes:
tan β̃
f (β̃) = mn(kπ + β̃) tan β̃ − − (m + n) = 0 (5.41)
kπ + β̃
Four or more inertias. For more complicated systems that cannot be sim-
plified down to three inertias or less, recourse must be made to more systematized
methods such as the Holzer–Myklestad , as is introduced in Section 3.2.1. Consider
an idealization of an elastic dynamic torsional shafting system, as approximated by
a series of multiple torsional elements each consisting of a concentrated torsional
inertia interconnected by massless springs, as shown in Fig. 5.18.
The iterative solution for the natural frequencies follows along the same lines
as is described for the blade torsion methodology. Again we seek a transfer matrix
approach that will require statements of equilibrium and continuity at the separate
T1 = J1 ω2 θ1 = J1 ω2 (1) (5.46a)
T2 = J2 ω2 θ2 (5.46b)
T3 = T1 + T2 + J3 ω θ3 2
(5.46c)
θ3 = θ1 − v1 T1 = θ2 − v2 T2 (5.46d)
θ4 = θ3 − v3 T3 (5.46e)
T4 = T3 + J4 ω2 θ4 (5.46f)
where vk = 1/Kk .
These equations can be then combined to yield the following more tractable
forms:
T1 = J1 ω2 θ1 = J1 ω2 (1) (5.47a)
θ2 = ω̄22 (ω̄12 − ω2 )/ω̄12 (ω̄22 − ω2 ) (5.47b)
θ3 = θ1 − v1 T1 = 1 − v1 T1 (5.47c)
T3 = ω 2
θ3 [J3 + 1/v1 (ω̄12 −ω 2
) + 1/v2 (ω̄22 − ω )]
2
(5.47d)
θ4 = θ3 − ν3 T3 (5.47e)
T4 = T3 + J4 ω2 θ4 (= 0, as one of the boundary conditions) (5.47f)
where
ω̄12 = 1/v1 J1 ; ω̄22 = 1/v2 J2 ; ω̄42 = 1/v3 J4 (5.48)
Note that ω̄1 , ω̄2 , and ω̄4 are the frequencies that each of the respective branches
would experience if the branch point were a node (i.e., θ3 had zero torsional
deflection).
1) If ω̄1 = ω̄2 , one natural mode will have a node at the branch point, and the
resulting natural frequency would be given by ω = ω̄1 = ω̄2 .
2) Similar situations can arise between the other two paired combinations of
branches.
3) If ω̄1 = ω̄2 = ω̄4 and J3 = 0, then there are only two distinct nonzero
frequencies (ω = ω̄1 = ω̄2 = ω̄4 ), and those are given by
4) If ω̄1 = ω̄2 = ω̄4 and J3 = 0, then θ3 ≡ 0, and the inertia torques as a result
of J1 , J2 , and J4 must be in balance.
5) The situations depicted in points 3 and 4 define variants of a nodal drive
configuration.
The condition of zero net torque at the branch point is given as a constraint equation
conveniently expressed in matrix form:
θ1
1 0 0
θ −J /J θ1 θ1
2 1 2 −J3 /J2 −J4 /J2
= θ3 = [T ] θ3 (5.51)
θ3 0 1 0
θ4 θ4
θ4 0 0 1
When this constraint equation is combined with the preceding set of dynamic
equations and the second of the resulting equations (second row of the matrix
equation) is discarded, we obtain
J1 0 0 θ̈1
K1 −K1 0 θ1
0
J3 0 θ̈3 + −K̃1 K̃B −K̃3 θ3 = {0} (5.52)
0 0 J4 θ̈4 0 −K3 K3 θ4
where
3
K̃B = Ki + J3 K2 /J2 (5.54)
i=1
where λ = −ω2 .
This matrix equation is nonsymmetric but can still be solved using any of a
variety of standard matrix eigenvalue techniques. Note that a symmetrical form
of the matrix equation can be obtained by instead retaining the second row
after the substitution of the constraint matrix T and then premultiplying by the
transpose of T(= T T = T −1 ). The resulting matrix equation is then 3 × 3 and
symmetric.
4) Formulate and apply the constraint(s) that connect all of the branch roots
together.
5) Use a robust eigenvalue routine to extract the roots of the resulting matrix
eigenvalue formulation.
Figure 5.21 demonstrates how each branch should be idealized into discrete
elements. As the figure implies, the details of the elastic elements can be quite
general; it can be a shaft with distributed inertia and stiffness, or a massless tor-
sional spring. Figure 5.22 gives the mathematical quantities needed to define the
dynamics of the element.
The methodology we use to formulate the dynamic description is a variant of the
Ritz/Hamilton method articulated by Bailey, which takes an energy approach. For
each element we separately formulate the kinetic and potential (strain) energies.
First, however, the torsional deflection over each (kth) element is taken as a power
series:
I
(k)
θk (η) = ai ηi (5.57)
i=0
where η is a the local spanwise coordinate over the length of the element and the
value of I to be used in the summation is selectable to as high a value as is needed
(k)
for convergence of the final eigenvalues. The ai coefficients represent the modal
variables in the final eigenvalue solution. The kinetic and potential energies are
then expressed as
1 t2 lk
T= ρJ0k θ̇k2 (η) dη + Jk θ̇k
(O)
dt (5.58)
2 t1 0
where ρJ0k is the distributed inertia, having the units of (F–S 2 ); Jk , the end
concentrated inertia, has the units of (F–S 2 –L) and (GJ)k , the distributed torsional
inertia, has the units of (F–L 2 ). After assuming sinusoidal motion, the application
of the Ritz–Hamilton method requires that the variation of the difference of these
energies be zero, that is,
δU − δT = 0 (5.60)
For the kth element this variation procedure results in the following expression:
t2 lk
δU − δT = (GJ)k θk (η)δθk (η) dη
t1 0
(O) (I) (O) (I)
+Kk θk − θk δ θk − θk dt
t2 lk
(O) (O)
+ω 2
ρJ0k θ (η)δθk (η) dη + Jk θk δθk dt = 0 (5.61)
t1 0
(k)
All of the variations in Eq. (5.62) are essentially variations on the coefficients ai .
Thus, Eq. (5.62) can then be expressed in a more useful variational form:
(k)
a
0
! " 1 a(k)
δa0(k) , δa1(k) , . . . , δaI(k) (−ω2 ) lk ρJ0k η̄i+j dη̄ + Jk [I] 1
..
0
.
(k)
aI
0
!
" 1 1 a1(k)
(k) (k)
+ 0, δa1 , . . . , δaI (GJ)k η̄i+j+2 dη̄ + Kk [I] .. =0
l 0
.
(k)
aI
(5.62)
Because each of the δai(k) is arbitrary, Eq. (5.62) can be written in the following
matrix form:
(k)
a0
(k)
# $ a1
−ω[M̃i, j ]k + [K̃i, j ]k .. = 0 (5.63)
.
(k)
aN
Note that the stiffness matrix for the kth element [K̃]k has a zero first row and
a zero first column. At this point all of the spanwise properties have been taken
care of, and the remaining chore is to properly string all of the elements together.
The first construction is to string all of the elements in each branch “element to
element”. Two relationships are needed; first, to go from element to element, the
following expressions must hold:
(I) (O)
θk = θk−1
I
∴ ao(k) = an(k−1) 2 ≤ k ≤ Nm (5.64)
i=0
(m)
Secondly, all of the (m) branch root deflections θR must be tied together to a
common trunk point deflection θT .
(m) (1)
θR = [a0 ](m) = θT (5.65)
(1,m) (m)
a0 θR (m)
(1,m)
(1,m)
θR
a1
a1
(1,m)
a
.. ...
1.
. ..
= = [Tm ] (5.66)
a0(2,m)
a0(2,m)
(2,m)
(2,m)
(2,m)
a
a
a
1
1
1
.
.
.. . . .
. .
(1,m)
a0
(1,m)
1 | | |
(m)
θR
a1
1
.. | 0 | 0 | 0
(1,m)
a
.
..
1
. | | |
.
(1,m)
.
.
aI
| | |
− − −
1
(1,m)
− − − − − − − − − − − − − − − − − − − − −
a
a
(2,m)
I
0 1 · · · 1 | 0 · · · 0 | |
− − −
(2,m)
0
a1
(2,m)
0 0 · · · 0 | 1 | 0 | 0
a
1
. . .
.
.. · · · | . . | | .
= 0 0 0
(2,m)
(2,m) 0 0 · · · 0 |
a
1 | |
aI
I
− − −
− − −
− − − − − − − − − − − − − − − − − − − − −
.
.
.
.
0 | · · · | . . | 0
.
.
.
− − −
| | · · · |
− − −
(Nm ,m)
0 0 0
a
(N m ,m)
0
0 | 0 | | 1 1
a
(Nm ,m)
a1
.
.
.
| | | . .
.
(Nm ,m)
.
.
.
| | | 1 a
(Nm ,m) I
aI
(m)
θR
(1,m)
a1
..
.
(1,m)
aI
− − −
a1(2,m)
..
(m)
. θ
= [Tm ] = [Tm ] R(m) (5.67)
aI(2,m)
A
− − −
...
− − −
(Nm ,m)
a
1
.
.
.
(Nm ,m)
aI
Thus, in Eq. (5.67) the mode shape vector has been compacted to include the
branch root deflection θR(m) plus all of the coefficients relating to the higher powers
deflection for the entire branch A(m) . Equation (5.63) can then be combined with
Eq. (5.67) to give the dynamic description for the mth branch:
(m) | |
−ω2 [M̃](m)
1 + [K̃]1 0 ··· 0
| |
− − − − − − − − − − − − − − − − − − − − − − − − − − − − −−
|
−ω2 [M̃]2 + [K̃]2 |
(m) (m)
0 ··· · · ·
| |
− − − − − − − − − − − − − − − − − − − − − − − − − − − − −−
.. ..
| |
−ω2 [M̃]3 + [K̃]3 · · ·
(m) (m)
. .
| |
− − − − − − − − − − − − − − − − − − − − − − − − − − − − −−
| .. | .. ..
0 . . .
| |
(m) (m)
θR θR
× [Tm ] −− = −ω2 [M (m) ] + [K (m) ] [Tm ] −− = {0} (5.68)
(m) (m)
A A
The final mathematical construction is to tie all of the branch root deflections
to the same trunk deflection θT . Also, as formulated by the preceding develop-
ment, the trunk point has not been attached to any inertia. Thus, the dynamics
of the trunk point must be included as a separate equation. The following matrix
operation achieves the required connections of the branch root points and prepares
the equation set for the trunk inertia equation:
θT
(1)
θ
1 0 ···
R
θT
(1)
1 0 ···
A
0 I 0 ··· (1)
(1)
A
θR 1 0 ···
= · · · A (2)
= [U]{X} (5.69)
A 0 I 0
(2)
1 0 ··· (3)
A
(1)
··· 0 · · ·
θR I ..
.. .
A
(3) .
..
.
The final form of the matrix eigenvalue problem is obtained by making the
Eq. (5.69) substitution and then premultiplying by the transposes of the U and
combined Tm matrices:
1 0 ··· T
0 [T1 ] 0 · · ·
..
T [T2 ] · · ·
[U] .
... ..
[T3 ] · · ·
.
.. ..
. .
−ω2 JT | 0 | ··· |
− − − − − − − − − − − − − − − − − − − − − − −
| −ω2 [M (1) ] | |
|
0 +[K (1) ] | 0 | ···
− − − − − − − − − − − − − − − − − − − − − − −
| | −ω2 [M (2) ] |
×
...
| 0 | +[K (2) ] | 0 ···
− − − − − − − − − − − − − − − − − − − − − − −
| | | −ω2 [M (3) ]
| .
. | |
. 0 +[K (3) ]
− − − − − − − − − − − − − − − − − − − − − − −
| | | ..
.
1 0 ···
0 [T1 ] 0 · · ·
..
[T2 ] · · ·
×
.
. [U]{X} = {0}
(5.70)
.. ..
. [T3 ] · · ·
.. ..
. .
Equation (5.71) is now in the final desired matrix eigenvalue form. The com-
bined mass and stiffness matrices comprising Eq. (5.72), [M] and [K], respectively,
are both symmetrical, and hence the equation is amenable to a variety of standard
matrix eigenvalue solution routines.
achieves its frequency characteristics using an elastic spring and is thus effective
at only one frequency of excitation, actually contributing to amplified responses
above and below this frequency. On the other hand, the torsional pendulum
absorber will attenuate responses at a frequency that is a constant factor of the
shaft rotation frequency regardless of that frequency!
Let the drive system have a torsional natural frequency of n, and let it be
excited at that frequency by a torque of amplitude T0 . Consider a pendular mass
m attached to the drive system as shown in Fig. 5.23. The equation of motion for
the pendulum response θ is
where θn is the torsional response of the shaft to be attenuated. Let the responses
be sinusoidal:
then
θ̄n 2 [R/L − n2 ]
= (5.74)
θ̄nA [n2 2 (R + L)/L]
A
T0 = mn2 2 (R + L)Lθ n (5.75)
drive train rotational degrees of freedom. For this reason, the additional inertia
provided by the absorber must be included with the base element, as shown in
Fig. 5.25.
Using Eqs. (5.72) and (5.75), we can form the appropriate attachment, torque
TnA , and the equation for the absorber:
One difficulty with this formulation is that it offers the solution the opportunity
to go singular for frequency values equal to [R/L]1/2 . One remedy for this is
to introduce a small amount of damping in the absorber equation of motion and
perform the Holzer–Myklestad method in complex arithmetic.
A second method for including the absorber dynamics in a general drive train
dynamic analysis is to use the matrix synthesis method outlined in Sec. 5.2.5 for a
typical helicopter. The basic methodology outlined in that section can be applied
to this case by including an additional equation, treating the base inertia as a trunk
point, and including a branch on either side of the absorber base inertia. Thus, the
2 cos β
= (5.79)
1 1 − sin2 β cos2 θ
where
1) The input and output angular velocities through the joint are not in a constant
ratio. For small shaft angles (β 1), this ratio contains a significant second
harmonic (2P) in the angular frequency of the driver end of the joint.
2) A velocity ratio of unity can be obtained at any angle β using two Hooke’s
joints with an intermediate shaft, as long as the input and final output shafts are
parallel and the pins on the ends of the intermediate shaft are oriented parallel.
An example of the use of a Hooke’s joint and the usage of Eq. (5.75) is the
calculation of the 2P torque in a tail rotor drive shaft, which is driving a (rigid) tail
rotor connected by means of a Hooke’s joint (gimbal) (see Fig. 5.27). Assumptions
that can be made to define the dynamics are as follows:
1) The flapping angle β is small.
2) The perturbational torsional deflections inboard and outboard of the gimbal
are θ1 and θ2 , respectively, and are also small.
Using these assumptions, we can write the velocity ratio equation between the
input and output ends of the shaft:
θ̇2 +
= cos β(1 + sin2 β cos2 t) (5.80)
θ̇1 +
Differentiation yields
where the underlined terms are small relative to the terms with which they are
added. Therefore,
where
GJs
ω2 = (5.85)
JR
The solution to this linear equation presents no difficulties, and we can write the
particular solution as
JR β 2 ω2 2
|T1 | = (5.89)
|42 − ω2 |
References
Section
5.1 Bishop, R. E. D., “The Vibration of Rotating Shafts,” Journal of
Mechanical Engineering Science, Vol. 1, No. 1, 1959, pp. 50–65.
Den Hartog, J. P., Mechanical Vibrations, McGraw–Hill, New York,
1940.
References (continued)
Ehrich, F. F., Handbook of Rotordynamics, McGraw–Hill, New York,
1992.
Mindlin, R. D., and Deresíewicz, H., “Timoshenko’s Shear Coefficient
for Flexural Vibrations of Beams,” Proceedings of Second National
Congress Applied Mechanics, ASME, New York, 1954.
Tse, F. S., Morse, I. E., and Hinkle, R. T., Mechanical Vibrations, Allyn
& Bacon, Boston, 1964.
Tuplin, W. A., Torsional Vibration, Pitman, London, 1966.
Young, W. C., Roark’s Formulas for Stress and Strain, McGraw–Hill,
New York, 1989.
5.2 Myklestad, N. O., Vibration Analysis, McGraw–Hill, New York, 1944.
Timoshenko, S., Vibration Problems in Engineering, Van Nostrand,
New York, 1937.
Tuplin, W. A., Torsional Vibration, Pitman, London, 1966.
5.3 Doughtie, V. L., and James, W. H., Elements of Mechanism, Wiley, New
York, 1954.
Schwamb, P., Merrill, A. L., and James, W. H., Elements of Mecha-
nisms, Wiley, New York, 1938.
Problems
5.1 Derive the expression for the frequency ratio R given in Section 5.1.4.
5.2 Derive the transcendental equation statement for the eigenvalue problem
defining the natural frequency characteristics of the simple three-inertia
problem wherein the shaft connecting two outer inertias also has significant
inertia.
5.3 Deduce an equation for determining the torsional natural frequencies for a
shaft with two end inertias and significant torsional inertia wherein one of
the end inertias approaches infinity. (This corresponds to a case of a built-in
condition at one end of the shaft with significant inertia and a finite inertia
at the other end.)
5.4 Set up the matrix partitioning and additional equation for the case of two
trunks (sheaves) connected by a stiffness KC as shown in Fig. 5.20.
5.5 Consider the tail rotor drive shafting of a certain helicopter wherein the (hol-
low) shaft is made of aluminum, is 12 ft long, and has an outside diameter
of 3 in. and a wall thickness of 0.1 in. This shaft drives a tail rotor that turns
at the same speed as the shaft and has a polar moment of inertia of 14 lb.-
s2 -in. For each of the two cases of 1) inboard attachment inertia approach-
ing infinity (rigid attachment to the remainder of the drive system, i.e., main
rotor, transmission, engine, etc.), and 2) zero inboard attached inertia (rep-
resentative of a failed coupling), calculate the first three torsional natural
frequencies. Additional useful information is given here: weight density of
A1, w = 0.101 lb/in.3 ; shear modulus of A1, G = 3.9 × 106 lb/in.2 .
elastic restoring forces the applied forces
= (6.1)
and moments and moments
189
Example 6.1
Given a simple spring-mass system, as shown in the following sketch, knowledge
of the dynamic load at the point P, FP , is required.
Solution
The differential equation for the single-degree-of-freedom system is, by now,
familiarly given by
These two alternate methods for formulating the load at point P define the basis for
the mode-deflection and force-integration methods, respectively, for calculating
dynamic loads. The procedure for calculating the dynamic load can then be stated
as follows:
1) Solve for the response z(t).
2) Calculate the load FP at the point P using either of the two expressions given
earlier.
Consider a step function for the applied load F(t):
0; t < 0
F(t) = (6.4)
1; t ≥ 0
The load at point P, FP (t), is then calculated using the two methods:
Mode-displacement method:
1
FP (t) = kz = k (1 − cos ωt) = 1 − cos ωt (6.6)
k
Thus, for the single-degree-of-freedom system the two methods are exactly
equivalent.
Example 6.2
Find the shear and bending moment at the midspan of the nonuniform cantilevered
beam shown in following sketch:
Solution
1) Differential equation of motion is
mz̈ + (EIz ) = F(x, t) (6.8)
the solution of which can be expressed using natural modes:
∞
z(x, t) = γn (x)qn (t) (6.9)
n=1
(6.13)
b) Shear S:
∞
∞
L
S ,t = −[EIw ] = − [EI(x)γn (x)] qn (t) = Sn qn (t)
2
n=1 x=L/2 n=1
(6.14)
By this approach the modal shear and moments are thus directly available.
4) Loads by the force integration method are discussed here: For the given
example beam the force integration method can be implemented to predict the
internal shear by integrating the total loading, both the external loading F(x, t), as
well as the motion-induced loading, that is, the loading produced by the beam in
responding to the external loading:
L
L
S ,t = [F(x, t) − mz̈] dx
2 L/2
L ∞
L
= F(x, t) dx − Gn q̈n (t) (6.20)
l/2 2
n=1
pseudostatic shear
where the Gn modal coefficients are related to the already defined Mbn modal shear
coefficients:
L
Gn (x) = m(x1 )γn (x1 ) dx1 = Sn (x)/ωn2 (6.21)
x
This representation of the internal shear thus consists of two distinct parts: 1) the
part that would be calculated if the external applied loading F(x, t) were applied
as a static loading and the beam were not allowed to respond structurally; and
2) a part that is caused by the inertial loading induced in the beam by virtue of
its dynamic response to the applied loading. It is for this reason that the force-
integration method is often referred to as the mode-acceleration method. In a like
manner the force-integration method can be used to calculate the bending moment
in the beam:
L
L L
Mb ,t = x− [F(x, t) − mz̈] dx
2 L/2 2
L L
∞
L
= x− F(x, t) dx − Hn q̈n (t) (6.22)
L/2 2 2
n=1
pseudostatic moment
where the Hn coefficients are defined in a manner similar to that for the Gn
coefficients:
L
Hn (x) = (x1 − x) m(x1 )γn (x1 ) dx1 = Mbn (x)/ωn2 (6.23)
x
Example 6.3
The following classic example taken from Bisplinghoff et al. concerns the wing-
root bending moments and shears of an idealized fuselage-wing combination.
Consider an unrestrained uniform wing attached to a central fuselage mass with a
disturbance force loading having the form of a half sine wave. For this simplified
dynamic structure we require the transient distributions of the maximum shear and
bending moment in the unrestrained wing (see Fig. 6.1).
Solution
1) Time-history response is as follows: The vertical response of the slender
beam can be represented by a superposition of natural mode shapes of the freely
Fig. 6.1 Uniform wing attached to fuselage mass subject to half sine wave external
force distribution (after Bisplinghoff et al.).
The equations of motion for the generalized coordinates q0 and qi are as follows:
mφi2 ( y) dy = M = MF + m ; i = 0, 1, 2, . . . , n (6.28)
0
(See Bisplinghoff et al. for detailed development and formulas for the evaluations
of Ci .) The time-history solution for w(ξ , t) then becomes the problem of solving
for the generalized coordinates qi (t):
1
t − sin t ; 0 ≤ t ≤ π/
M
q0 (t) = (6.30)
t
; t > π/
M
Ci
(sin t −
0 ≤ t ≤ π/
sin ωi t);
M(ωi − )
2 2 ωi
Ci
qi (t) = sin t + sin (t − π/2) (6.31)
M(ωi2 − 2 )
− [sin ω i t + sin ω i (t − π/2)] ; t > π/
ωi
Moment:
∞
(i)
Mb (ξ , t) = Mb (ξ ) qi (t) (6.33)
i=1
where
1
S (i) (ξ ) = m ωi2 φ(λ) dλ (6.34)
ξ
1
Mb(i) (ξ ) = S (i) (λ) dλ (6.35)
ξ
Moment:
∞ M (i) (ξ )
b
Mb (ξ , t) = (Mb )pseudo − q̈i (t) (6.37)
static i=1
ωi2
where
Spseudo = (M − m )(1 − ξ )q̈0 (t) (6.38)
static
(Mb )pseudo = ( /2)(M − m )(1 − ξ )2 q̈0 (t) (6.39)
static
Note that the pseudostatic shear and moment Spseudostatic and (Mb )pseudostatic ,
respectively, are obtained by treating the beam (wing) as a rigid body.
Fig. 6.2 Shear and bending moment distributions for displacements of the vibratory
modes (R = 1) (after Bisplinghoff et al.).
Fig. 6.3 Maximum shear distribution (TI /T1 = 0.5, R = 1) (after Bisplinghoff et al.).
Figure 6.3 shows the spanwise variation of maximum shear and the instants
of time that the maximum shears occur, as determined by the two methods. It is
apparent that the mode-acceleration (force-integration) method gives much bet-
ter convergence than the mode-displacement method, particularly in the vicinity
of the root, where the higher modes contribute quite substantially to the shear
when the mode-displacement method is used. In fact, it can be seen that, even
with five modes, satisfactory convergence of the shear is not obtained by the
Fig. 6.4 Maximum bending moment distribution (TI /T1 = 0.5, R = 1) (after
Bisplinghoff et al.).
1) Bending moment Mb :
∞
Mb (x, t) = EIz = EI(x)γn (x)qn (t) (6.40)
n=1
2) Shear S:
∞
S(x, t) = −[EIz ] = − [EI(x)γn (x)] qn (t) (6.41)
n=1
The only new element here is the problem of finding appropriate expressions for
the load coefficients for rotating beams when the second- and third-order spanwise
derivative forms of the coefficients are not used in the evaluations. Consider the
application of the semianalytical (integral) method (as given earlier) for evaluating
the coefficients. In particular, consider the inertial loads acting on the beam when
it is deflected in one of its flatwise natural modes, as depicted in the free-body
diagram given in Fig. 6.5. Upon equilibrating the inboard internal forces and
moments with those arising from the loads acting on the beam over its outboard
portion, one can then write the following expressions:
L
Mbn (x) = ωn2 (x1 − x)m(x1 )γn (x1 ) dx1
x
L
− 2 [γn (x1 ) − γn (x)]m(x1 )x1 dx1 (6.42)
x
Thus, the modal bending moment coefficients can be written in the following
abbreviated form:
where H̄n (x) is now the contribution caused by centrifugal forces. In a similar
manner the modal shear coefficients can be written as follows:
The Gn (x) and Hn (x) coefficients are as defined in the preceding section, and the
new Ḡn (x) and H̄n (x) coefficients are summarized as follows:
L γn (x)
Ḡn (x) = γn (x) m(x1 )x1 dx1 = T (x) (6.46)
x 2
tension
L L
H̄n (x) = m(x1 )γn (x1 )x1 dx1 − γn (x) m(x1 )x1 dx1 (6.47)
x x
= T (x)/2
As in the case without rotation, the results of a Myklestad technique for obtaining
the mode shapes can also be used to give values for Sn (x) and Mbn (x) directly.
as that developed in the preceding sections. Again, let us consider the uncoupled
flapwise bending of the blade caused by out-of-plane loads:
L
Mb (x, t) = (x1 − x)[F(x1 , t) − mz̈1 ] dx1
x
L
− 2 (z1 − z)mx1 dx1 (6.48)
x
After breaking out the parts of the load distribution as a result of flapwise motion
and the part of the out-of-plane deflection caused by steady flapping (or, in the case
of hingeless rotor blades, built-in coning), the expression for out-of-plane bending
can be stated as follows:
L L
Mb (x, t) = (x1 − x)F(x1 , t) dx − β0 2 m x1 (x1 − x) dx1
x x
pseudostatic flapwise bending moment
∞
− [Hn (x)q̈n + 2 H̄n (x)qn ] (6.49)
n=1
Note that the bending moment relief caused by steady coning β0 becomes readily
apparent with this formulation. In a similar manner we can write a force-integration
expression for the flapwise shear:
L L
S(x, t) = [F(x1 , t) − mz̈1 ] dx1 − 2 z m x1 dx1 (6.50)
x x
And, again, after separating out the part of the loading caused by modal accele-
ration, the final requisite form of the flapwise shear can be written as
L L
S(x, t) = F(x1 , t) dx1 − β0 2 mx1 dx1
x x
pseudostatic flapwise shear
∞
− [Gn (x)q̈n + 2 Ḡn (x)qn ] (6.51)
n=1
We now complete this section with a consideration of the two loads analysis
methods as applied to the in-plane bending moment and shear of a rotating beam
(i.e., rotor blade).
Note that here the bending moment relief comes from e0 , the equivalent of a
steady-state lead-lag angle. In a similar manner we can write a force-integration
expression for the in-plane shear:
L L
S(x, t) = [F(x1 , t) − m( ÿ1 + 2 y1 )] dx1 − 2 y mx1 dx1 (6.57)
x x
And, again, after separating out the part of the loading as a result of modal
acceleration, the final requisite form of the in-plane shear can be written as
L L
S(x, t) = F(x1 , t) dx1 − e0 2 m dx1
x x
pseudostatic in-plane shear
∞
− {Gn (x)q̈n − 2 [Gn (x) − Ḡn (x)]qn } (6.58)
n=1
The in-plane offset distance e0 , which provides in-plane load relief, is defined in
Fig. 6.6. Note that, in the case of rotor blades with in-plane articulation, the offset
distance e0 is simply related to the hinge offset distance e and the equilibrium lag
angle ζ0 :
e0 = e sin ζ0 (6.59)
In all of the preceding development, the portions of the transverse load distribution
caused by the inertial acceleration and the centrifugal force loadings were explicitly
broken out in order to implement the two basic types of internal load calculation.
In the next section the whole issue of loads caused by motion is more generally
developed.
The principal observation to be made is that the M n terms now couple the
(previously) uncoupled natural modal equations. The system can no longer be
represented by a series of single-degree-of-freedom forced equations. Assuming
for simplicity that M n are functions of only the generalized velocities q̇i , then the
system of equations is expressible in the following matrix format:
M1 0 q̈1
ξ11 ξ12 · · · ξ1N q̇1
M2 q̈2 ξ21 ξ22 · · · ξ2N q̇2
. + .
.. . . .. .. ..
. . . . ··· . .
0 MN q̈N ξN1 ξN2 · · · ξNN q̇N
D
M1 ω12 0
q1
1 (t)
M ω2
D
2 2 q2 2 (t)
+ ..
. =
. .. (6.65)
. .
.
D
0 MN ωN2 qN N (t)
Before we discuss ways to solve such a matrix equation, let us use the preceding
ideas on some examples.
Example 6.4
Consider the two-dimensional airfoil section of an earlier homework problem that
is characterized by the presence of inertia coupling. Let the airfoil now experience
a freestream velocity of air of 100 mph. Find the equations of motion describing
the response of the airfoil section to a step gust velocity wG of 10 ft/s.
Solution
The differential equations of motion, assuming generalized forces, are given by
(a) (a)
where the generalized forces fz and fθ are now assumed to be the result of
a disturbance gust penetration and motion dependency. Furthermore, let these
loads be defined using a simplified quasi-steady aerodynamic formulation wherein
the lift acts at a chordwise point xa in front of the elastic axis. (Note that this
quasi-steady formulation is deemed simplified because we are neglecting, for illus-
trative purposes only, the effects of pitch rate θ.) Then, with this simplification in
mind, we obtain the required generalized forces:
1 2 1 2 dcI 1
fz(a) = − ρV ccI = − ρV c θ + (wG + ż) (6.68)
2 2 dα V
(a) 1 1 dcI 1
fθ = ρV 2 ccI xa = ρV 2 cxa θ + (wG + ż) (6.69)
2 2 dα V
A solution for the time-history response of this system can be obtained using a
variant of the Galerkin technique wherein the physical variables z and θ are first
represented by a combination of the two natural modes as obtained from the zero
It can then be verified that the various coefficients in this equation are given by
where
$
1 (
ζi = cii Mi kii ; ωi = kii /Mi 1 − ζi2 ; i = 1, 2 (6.77)
2
In this example the calculation of the internal loads could be accomplished
using either of the two basic methods. The mode-deflection method would
have one advantage in the simplicity afforded by only calculating the modal
responses and then simply multiplying them by appropriate constants. The force-
integration method in this case (and generally for most aeroelastic problems)
requires the consistent calculation of all of the external loadings—both inertial and
aerodynamic—for an accurate solution. In this example this requirement involves
the use of the time histories of the modal displacements and velocities, as well as
the accelerations.
Example 6.5
Examine the in-plane response and resulting in-plane bending characteristics of
an articulated rotor with a lead-lag damper (see Fig. 6.7).
This example demonstrates how motion-dependent loads can enter the dynamic
description through the boundary conditions, in this case as a result of the loads
Fig. 6.7 Pictorial representation of the in-plane loading and root restraint of an
articulated rotor blade.
imparted by the lead-lag damper. To this end, we denote the in-plane elastic
deformation by the variable v(x, t), and make the following simplifying assump-
tions:
1) The in-plane loads result solely from inertia considerations. The airloads,
represented by fy (x, t), do not depend on the in-plane responses, that is, the
transverse load distribution is given by
2) At the blade root the bending moment is related to the time derivative of the
slope and the damper rate of the lead-lag damper:
∂ 2v ∂ 2v
EI 2
= −Kζ (6.79)
∂x ∂x∂t
Solution
The basic differential equation of motion is that of the rotating beam for
in-plane motion (as given in Chapter 3) with the addition of the disturbance
transverse loading fyD :
J
v(x, t) = γj (x) qj (t) (6.81)
j=0
The appropriate mode shapes are those for the “pinned” boundary condition at
the root. Note that, in this case the index j = 0 corresponds to the rigid-body
rotational mode [γ0 (x) = x] about the hinge point. We then invoke a standard use
of the Galerkin technique wherein we multiply the equation by the kth mode shape
and integrate over the blade. Consider the integration of the elastic stiffness term,
denoted for this purpose by Ksi , using integration by parts:
L
Ksi = γi (x) [EIv ] dx (6.82)
0
where
Thus, once again, the motion-induced loads at the root caused by the lead-lag
damper result in a coupling between the natural vibrations modes. The resulting
(matrix) differential equation is again of the form described earlier:
A solution can then be obtained for the response vector {q(t)} by either a rigorous
solution with the modal variables qi (t) being coupled by the fully populated C
matrix, or by an approximate one wherein the off-diagonal C matrix coupling terms
are neglected (bij = 0, where i = j), and then qi (t) are solved for independently.
Once the responses are obtained the loads in the blade can then be solved for
using either of the two basic load calculation methods. However, one problem
with the use of the mode deflection method is that, because pinned root boundary
condition modes were assumed, there is no mechanism present that can give a
nonzero bending moment at the blade root, consistent with the bending moment
imparted by the lead-lag damper. However, this difficulty is overcome with the
force-integration method.
One note of caution with the force-integration method follows: Because the
method implies a degree of balancing of the loads, the statement of the loads
calculation must be consistent with that for the response calculation. For instance,
if an approximate solution scheme is used, then the force-integration method will
give completely erroneous results. However, the mode-deflection method is not
typically prone to this type of inaccuracy.
The DMF is, in general, defined for only single-degree-of-freedom systems for a
variety of “standard” forcing functions. Some of these standard forcing functions
are considered in this section.
1) Step function (undamped spring-mass system):
Basic equation:
The solution for the response (subject to zero initial conditions) is given simply as
x = 1 − cos ωn t (6.89)
which is graphed in Fig. 6.8. For this case the DMF is clearly equal to 2.
2) Step function (damped spring-mass system):
Basic equation:
In a similar manner the solution for this case is graphed in Fig. 6.9. For this case
the DMF is given by
DMF = 1 + exp(−π ζ / 1 − ζ 2 ) (6.93)
where, by providing for a phase angle φkn , we can select qkn to be a real-valued
quantity. With this solution form the amplitude can be written as
Ckn
qkn = (6.96)
{[1 − (n/νk )2 ]2 + (nσk / νk )2 }1/2
The pseudostatic response is obtained by neglecting the acceleration and velocity
terms:
where (DMF)kn is as given earlier. With this formulation the following should be
noted:
1) (DMF)kn will be approximately unity (1.0) for modes whose natural frequ-
encies (vk ) are well above the harmonics of excitation and will be ≈0 for
modes whose natural frequencies are well below these harmonics. This follows
directly from the well-established frequency-response characteristics of a single-
degree-of-freedom dynamic system.
2) The formula tends to be conservative because it assumes that the maximum
displacements of each mode occur simultaneously.
These concepts can then be illustrated by considering the flapping excitation
of a rotor blade. Figure 6.10 shows the velocity components contributing to the
airloading at a typical blade section (whose flapping motion ż is again defined in
the downward direction).
Using a simple two-dimensional strip theory formulation of the aerodynamic
loads on this section, an approximate loading distribution can be written. Of prime
importance is the role of the induced velocity vi appearing in this formulation.
Induced velocity. The induced velocity is the component of the air velocity
impinging on the airfoil blade section that arises from the presence of the lifting
rotor itself. Generally, the induced velocity, as encountered by the lifting air-
foil section, is highly variable both in time and space (i.e., spanwise position).
The zero-frequency component of the induced velocity can be attributed to the
response of the air mass to the action of the rotor in producing a thrust. In a
similar manner time-variable components of the induced velocity arise from the
harmonic responses of the blade sections. These responses cause motion-induced
Fig. 6.10 Velocity diagram for airloads at a typical rotor blade section.
harmonic loadings on the blade and must therefore also be accompanied by har-
monic changes of momentum in the air mass. At this point, such a description
is a necessarily overly simplified explanation of a complicated phenomenon that
encompasses a variety of other important fluid dynamic effects. But it is sufficient
to establish the case that the induced velocity vi can be considered to be comprised
of a steady part vi(0) and a multiharmonic part that is also a spanwise variable
(H)
vi . This latter component provides the major harmonic excitation of the blade
dynamics.
Upon taking the geometric blade pitch angle to be comprised of a steady part
θ0 and the (1P and higher) harmonic part θ1 , we can then write the (upwardly
directed) lifting airload distribution. [Here, we have assumed again that the effects
of pitch rate are negligible (for illustrative purposes) and that we have essentially
a hovering flight condition.] The resulting simplified expression is given by
1 ż − vi
fza (x, t) ≈ ρac(x) θ +
2
2 x
1
= ρac2 [x 2 θ0 − xRλ] ← (steady loads)
2
1 (H)
+ ρac[x 2 θ1 + xż − xvi ] (6.100)
2
Mk q̈k + vk2 2 Mk qk = M
k + k
D
(6.101)
where
1
1
M
k = − ρacR4 γk (η)γj (η)η dη q̇j = −σkj q̇j (6.102)
2 0
1 1
k = − ρac R (A2k θ0 − A1λ ) − ρac R A2k θ1
D 2 4 2 4
2 2
1
1 (H)
− ρacR 4
γk (η)vi (η, t)η dη (6.103)
2 0
where, in the first part, A1k and A2k are integration constants involving the mode
shape γk and the nondimensional spanwise variable η(= x/R), as given in the
following:
1
A1k = γk (η)η dη (6.104a)
0
1
A2k = γk (η)η2 dη (6.104b)
0
Although the effect of steady coning β0 was not expressly included in the preceding
development (in order to focus on the effects of the aerodynamics), we can see that
the first part of D
k represents the steady blade loading that this coning relieves.
Root shears. On any single blade the shears at the root will be harmonic;
because the blades are assumed to be identical, they are all described by the same
Fourier series but with a different phase angle. Thus, for the jth blade the root
where, again, k = 0, 1, 2, . . . .
Root bending moments. Because root bending moments, like root shears,
are vector quantities, they are transferred through the hub to the fuselage by the
same trigonometric resolution equations as are the root shears. Thus, by noting the
similarity of the appropriate vectors, the nonrotating hub moments can be obtained
using the following substitutions:
Sy → −MbF
Sz → MbI
Hx → Mhx (6.112)
Hy → M hy
Hz → −Q (negative of the rotor torque)
where the b and h subscripts refer to blade and hub, respectively, and where ( )F
and ( )I refer to flapwise and in plane, respectively.
The following summary can be made:
1) The hub acts as a filter for steady harmonic loads, that is, it admits only the
harmonics that are integral multiples of the number of blades b.
2) Although vertical admitted blade loads have the same harmonics as those of
the hub (k × b), the lateral admitted blade harmonics are +1 and −1 those of the
hub (again, k × b). Thus, the steady and kb harmonic vertical blade loads and the
first (kb + 1) and (kb − 1) harmonic in-plane blade loads are the only ones passed
(as zeroth and kb harmonic hub loads) by the hub.
ωF = ωW + kb (6.113)
The simplest case would be where k = 0 and the rotor disk would appear to
be undergoing a whirling motion at a frequency of ωW , as is discussed in
Sec. 4.4. For the cases where k ≥ 1, the rotor would appear to be undergoing
a “traveling-wave” motion either in the progressive or regressive direction, or
a combination of the two directions. The oscillatory loads would correspond-
ingly have similar interpretations. These nonintegral harmonics, as was stated
earlier, are not generally excited or driven by the main rotor airloads because such
airloads are essentially integral harmonic in character. Such nonintegral loads
(and motions) must therefore come from some other source of fuselage excitation
and, in order to have significant response levels, must necessarily involve rotor
blade motions that are relatively devoid of (aerodynamic) damping. The follow-
ing section treats the area of non- (main) rotor sources of fuselage excitation;
the subject of what conditions lead to the rotor blades losing aerodynamic damp-
ing falls under the category of blade aeroelastic stability and is covered in later
chapters.
sensed as local rattling of the canopy. This type of vibration is typically controlled
by providing adequate stiffness to the top of the canopy and sufficient mast height
to the rotor to give clearance.
The second source of integral harmonic excitation arising from the aerodyna-
mics of the main rotor is the system of trailing vortices constituting the far wake
of the rotor. These vortices impact on the two fuselage empennage aerodynamic
surfaces (i.e., the horizontal stabilizer and the vertical pylon fin). Such excitations
are quite dependent on the geometry and proximity of the wake structure relative
to these aerodynamic surfaces. As shown in Fig. 6.13, depending on the flight
condition, the trailing blade-tip vortex can pass within close proximity to the
horizontal stabilizer twice per rotor reduction before it is convected away from the
helicopter.
The analysis of the pressure excitations on the horizontal surface caused by the
near passage of the trailing vortices is a substantial exercise in fluid dynamics, and
the details of a solution are beyond the intent of this text. However, some of the
main characteristics of the solution can be identified. When vortices pass within
close proximity to a lifting surface such as a wing, nonpotential flow characteristics
of each element, such as the boundary layer of the wing and the finite core of the
vortex, can each interact with each other as well as with the potential flow charac-
teristics of each. Furthermore, a requirement of such a flow problem is knowledge
of the trajectories of the vortices, which constitutes a separate (substantial) poten-
tial flow problem in itself. Generally, this type of aerodynamic problem falls in
the category of wing-vortex interaction and is not yet fully solved. Because of the
ability of the trailing vortices to pass within close proximity to the surface twice per
Fig. 6.13 Division of the rotor wake into upper and lower parts, µ = 0.4, only tip
vortices shown for ψ = 30 deg (after Gangwani).
rotor revolution and because of the highly impulsive nature of the pressure pulses,
these excitations are integral harmonic with substantial higher harmonic content
(i.e., ωE = kb, k = 1, 2, 3, . . .). The excitations caused by this main rotor wake
impingement on the empennage surfaces are aggravated by the susceptibility of the
fuselage vibrations to excitations at this relatively flexible portion of the fuselage.
the interactions occurring between the rotor and the fuselage. The importance of
this rotor-fuselage interaction is manifested in the fact that oscillatory loads being
transferred from the rotor to the fuselage via the hub can often be greatly amplified
from what they would be if the fuselage were considered to be infinitely rigid. And
yet, such “hub-fixed” harmonic rotor hub loads calculations have been accepted
methodology for many years with scant attention paid to the coupling mechanism.
Indeed, several important mechanical and aeromechanical instabilities result from
rotor-fuselage interactions including ground resonance and air resonance, wherein
the rotor can drive the fuselage in divergent oscillations. Furthermore, as could
be anticipated, the greater control power of hingeless and bearingless rotor con-
figurations tends to make interaction effects more pronounced than they would be
with an articulated rotor.
As developed in this section, the first major task in assembling the principal
elements needed for accurate fuselage vibration calculation is to gain accurate
knowledge of the vibration characteristics of the rotor and fuselage taken sepa-
rately. That is, we must know the natural frequencies and mode shapes of both the
rotor blades and the fuselage, as well as the (aerodynamic) forcing of the blades
and the fuselage. The second task is to gain commensurately accurate knowledge
of the dynamics between the rotor and the fuselage. In an interactional situation, by
definition, the fuselage and rotor are each developing and, to a degree, amplifying
oscillatory loads in the other at some common frequency, typically the blade pas-
sage frequency. Figure 6.14 depicts an idealization of the loading of the fuselage
by the rotor airloads while the rotor and fuselage are undergoing an interactional
coupling. In this depiction the applied (non-motion-induced) airloads acting on
the rotor blades are denoted { fapp1 } while the total integrated airloads acting on
Fig. 6.14 Schematic of the aerodynamically excited fuselage and the rotor-fuselage
interaction.
the fuselage at the hub are denoted { fapp2 }. Note that in this figure the interaction
between the rotor and the fuselage is represented by the internal loads at the hub
{ fh } and the motions of the hub {qh }. The interactional process can then be further
conceptualized by the block diagram given in Fig. 6.15.
where [K], [M], and [C(ω)] are the N × N stiffness, mass, and damping matri-
ces, respectively. The matrix terms on the left-hand side of the equation define
a response-to-loads relationship and can be termed the displacement impedance
matrix [Z] or
[Z]{q} = { f } (6.116)
{ fH }F = { f M }F + { fapp2 } = −{ f M }R + { fapp2 }
= −[ZR ]{qH } + { fapp2 } (6.119)
Hub responses:
Then, upon eliminating {qH } and { fH }F from these expressions separately and in
turn, the following two sets of relationships are obtained:
Hub loads:
−1
{ fH }F = [I] + [ZR ][YF ] { fapp2 } (6.121)
Hub responses:
−1
{qH } = [I] + [YF ][ZR ] [YF ]{ fapp2 } (6.122)
Thus, vibratory response of the fuselage can then be calculated once the three basic
quantities are known:
Natural frequencies:
ω1 , ω2 , . . . , ωN
Generalized masses:
M̃1 , M̃2 , . . . , M̃N
Fig. 6.16 Comparison of measured and analytic responses of fuselage c.g. as a result
of tail excitations.
Fig. 6.16 Comparison of measured and analytic responses of fuselage c.g. as a result
of tail excitations (continued).
where the generalized masses are defined as
K
(k) (k)
M̃i ≡ mk φ i × φ i (6.123)
k=1
Note that k is the index over all of the fuselage mass elements as defined in the
FEM analysis; K is therefore typically a large number.
or
where [ZF ( p)] is the generalized fuselage impedance matrix, as a function of the
Laplace operator p. For our purposes, at present, the number of modes selected
must be equal to six. Correspondingly, let the modal matrix, containing only the
components of hub motion, be as denoted by ( ˆ ), as defined by the following
expression:
where
x Q1
y
Q2
z
Q
3
{qH } = = [] = [] {Q} (6.127)
θx
Q4
θy
Q5
(6×6)
θz Q6
and where {Q} is a vector of generalized coordinates defining the modal activity.
By definition of the generalized mass, a diagonal (uncoupled) modal mass matrix
can be formed:
Then, by assuming that the [] matrix is well behaved (i.e., is not near a singular
condition), the following expression can be written:
Similarly,
It can be verified that these [M], [C], and [K] matrices will then yield a six-mode
eigensolution characterized by natural frequencies, mode shapes, and damping
values that are very near to those six selected modes of the original elastic fuselage
structure. The general (Laplace transform domain) fuselage impedance can then
be written as
[ZF ( p)] = [M] p2 + [C] p + [K] (6.132)
from which the (frequency-domain) fuselage mobility matrix can then be obtained:
−1
[YF (ω)] = (iω)2 [M] + iω[C] + [K]
−1
= [K] − ω2 [M] + iω[C] (6.133)
Two comments should be made regarding this method for calculating fuselage
mobility:
1) The method does require eigenvalue results for exactly six vibrational modes,
and the mode shapes must form a well-conditioned modal matrix.
2) The method yields a potentially convenient representation of the fuselage
dynamics for use in analyses wherein the coupling of the rotor with a rigid-
body (free-flying) fuselage is already available. This methodology thus extends an
otherwise rigid-body analysis into the modal response regime.
Note that because no inverses are indicated in this expression, [] need not be
square. Therefore, any number of natural modes can be used (i = 1, 2, 3, . . . , N
where N is arbitrary).
or
+
qR
[Z11 ( p) · Z12 ( p)] −·− = {FR1 } (6.138)
qH
where the first subscript on the Zij submatrices refers to the equation “group.” In
this case, i = 1 refers to the rotor modal equations. The second subscript j refers to
the degree-of-freedom group. Again, the subscript j = 1 refers to the rotor modal
degrees of freedom, whereas the subscript j = 2 refers to the hub (fuselage) degree-
of-freedom group. Next, consider the equations of motion defining the hub or
fuselage degrees of freedom. (Note that, although the fuselage degrees of freedom
can be expressed in either modal or explicit finite element nodal descriptions, the
following formulation is most appropriate to a modal description for the fuselage):
or
+
qR
[Z21 ( p) · Z22 ( p)] −·− = {FH1 } (6.140)
qH
These equations can then be combined to yield the total matrix equations of motion
for the rotor-fuselage coupled dynamics:
, -+ +
Z11 ( p) · Z12 ( p) qR FR1
−·−·− − ·− ·−·−·−−·− −·− = −·− (6.141)
Z21 ( p) · Z22 ( p) qH FH1
This equation is then a general statement of the fully coupled rotor-fuselage interac-
tion problem and can be solved using standard inverse Laplace-transform methods.
In each of the equation group sets, the first of the applied loads is identified as
being of aerodynamic origin. Although this generalization is fairly accurate for
the rotor, there can typically be additional sources of nonrotor excitation for the
fuselage, as arising from the drive system and/or the empennage, for example.
Thus, {FH 1 } consists of the sum of the aerodynamic excitation components at the
(h) (nh)
hub {FH 1 }, and all of the remaining (nonhub) components {FH1 }:
(h) (nh)
{FH1 } = {FH1 } + {FH1 } (6.142)
Now, restricting our attention to only the excitations at the hub, let us consider the
fuselage modal equations of motion in the preceding rearranged form:
whereupon, after substitution for {qR } (using straightforward matrix algebra tech-
niques), we obtain the following equation for the fuselage response variables:
[Z22 ( p)] − [Z21 ( p)][Z11 ( p)]−1 [Z12 ( p)] {qH }
= {FH1 } − [Z21 ( p)][Z11 ( p)]−1 {FR1 } (6.144)
Next let us separate out those contributions to [Z22 ] because of the fuselage and
the rigid-body rotor masses, respectively:
[Z22 ] = [Z22
F
] + [Z22
R
] (6.145)
The integrated aerodynamic loads and (hub-fixed) blade aeroelastic induced loads
can be written as
[ZR ] = [Z22
R
] − [Z21 ] [Z11 ]−1 [Z12 ] (6.150)
where [Z22R ] denotes the rotor acting as a rigid body, [Z ] denotes rotor (mode) to
21
hub motion coupling, [Z11 ]−1 denotes hub-fixed rotor modal response dynamics,
and [Z12 ] denotes hub motion to rotor (mode) coupling. Although the preceding
development serves to define the concepts involved, the actual calculation of rotor
impedance can be accomplished using a comprehensive rotor aeroelastic analy-
sis for which the dynamic effects of prescribed hub motion are included. Using
the aforementioned development as a guide, we can formally define the rotor
impedance matrix as follows:
, -
∂
[ZR ] = (hub loads as seen by the fuselage)
∂qHj
, -
∂
= [[Z22 ( p)]{qHj } + [Z21 ( p)]{qR }]
∂qHj
, , --
∂qR
= [Z22 ( p)] + [Z21 ( p)] (6.151)
∂qHj
where
1
{Z̃Rj } = {FH2 (qR + qR , q̇R + q̇R , q̈R + q̈R )j − FH2 (qR , q̇R , q̈R )}
qHj
(6.154)
Here, {FH2 } is the vector already defined of hub excitation terms arising from blade
aeroelastic responses. The vector {qR } denotes the rotor (or blade) degrees of free-
dom (aeroelastic responses) resulting from the fixed-hub condition ({qH } = {0}).
The vector {qR + qR }j denotes the modified rotor responses resulting from the
sinusoidal motion of the jth hub degree of freedom, with all other hub degrees of
freedom set to zero, that is, {qR + qR }j results from
0
..
.
{qH } = q̄Hj eiωt (6.155)
..
.
0
References
Section
6.1 Bisplinghoff, R. L., Ashley, H., and Halfman, R. L., Aeroelasticity,
Addison–Wesley, Cambridge, MA, 1955.
6.2 Bramwell, A. R. S., Helicopter Dynamics, Wiley, New York, 1976.
Payne, P. R., Helicopter Dynamics and Aerodynamics, Macmillan,
New York, 1959.
References (continued)
6.3 Gangwani, S. T., “Calculation of Rotor Wake Induced Empennage Air-
loads,” Journal of the American Helicopter Society, Vol. 28, No. 2,
1983, pp. 37–46.
6.4 Dowell, E. H., Curtiss, H. C., Scanlan, R. H., and Sisto, F., A
Modern Course in Aeroelasticity, Sijthoff & Noordhoff, Rockville,
MD, 1978.
Payne, P. R., Helicopter Dynamics and Aerodynamics, Macmillan,
New York, 1959.
Cansdale, R., Gaukroger, D. L., and Skingle, C. W., “A Technique for
Measuring Impedance of a Spinning Model Rotor,” Royal Aircraft
Establishment TR-71092, London, 1971.
Problems
6.1 Consider a simplified representation of a beam that is free at one end and
constrained by a pin connection and a rotary damper at the other. Let the free
end of the beam be excited by a force of amplitude F and step function time
variation:
(a) Write the equations of motion for the two degrees of freedom, z1 and z2 .
(b) Solve for the response vector [z1 (t), z2 (t)] both exactly and using a
Galerkin (modal) approach. Assume unit values for m, c, k, and .
(c) Considering the rotary damping to be applied exactly at the pin con-
nection (in the limit), calculate the bending moment just outboard of the
damper attachment point three ways:
1) using the exact solution and the (damper rate) × (angular velocity) at
the pivot point;
2) using the mode displacement method together with the modal response
solution; and
3) using the force integration method again together with the modal
response solution.
(d) Discuss how answers using the preceding three methods would
compare if the excitation function were an impulse function.
6.2 Find the DMF for a damped single-degree-of-freedom system for the
following forcing function:
Sketch the variation of the DMF with the parameter (T ωn ) for damping ratios
of 0 and 0.5.
In the previous chapter the basic tools are identified (if not as yet completely
formulated) for analyzing the helicopter/rotorcraft for vibration. The basic vibra-
tion analysis problem boils down to knowing the following items with sufficient
accuracy:
1) the aerodynamic environments of both the rotor and the fuselage;
2) the aerodynamic and aeroelastic response characteristics of the (rotating)
rotor in the form of integrated hub loads;
3) the structural characteristics of the (nonrotating) fuselage, in the form of
accurate frequency response functions ( fuselage mobilities); and
4) the rotor-fuselage interactions, in the form of rotor impedances.
Once each of these separate items can be predicted accurately, coupling tools
exist for combining them to give reliable vibration predictions. Such predictive
capability, if available sufficiently early in the design process, should lead to the
quantum reduction in vibration (with minimum weight) that is so highly prized.
However, the required knowledge to achieve this state of affairs is not yet at
hand, and aircraft design, fabrication, and certification must continue without the
benefit of this foreknowledge. Thus, methods that have necessarily been devised
for modifying the aircraft design “after the fact” will continue to be used to reduce
vibration to some tolerable level. This chapter describes major methods devised
and presently either used outright or seriously being considered for controlling
vibration.
(mode shape) detuning. For any given dynamic problem this process requires some
knowledge of how the resonant magnification varies with the set of design para-
meters appropriate to the problem. This variation of the resonance characteristics
constitutes trend information, which must somehow become known to guide the
modification process. The information can, of course, be obtained by calculat-
ing an in-depth ensemble of results for wide variations in the design parameters.
These results typically take the form of graphs showing the explicit variations of
the characteristics of interest for all feasible values of the design parameters. An
alternate method of dealing with trend information is to compute the derivatives
of the pertinent resonance characteristics directly either analytically, if possible,
or by direct numerical calculation of the partial derivatives constituting this trend
information. The following sections deal with methods for computing the trend
information.
where the response vector {Z} is given by a modal decomposition using the first
J modes:
J
t
{Z} = {φ(x, pm )} eiωj ; where ωj = ωj ( pm ) (7.2)
j=1
Note that because the eigenvalue problem is defined in terms of the modifica-
tion parameters pm (m = 1, 2, . . . , M), both the modal matrix [] (a J × J matrix
whose columns are the mode shapes) and the J natural frequencies ωj are implicit
functions of the design parameters. The solution of the preceding eigenvalue prob-
lem for a given set of modification parameters pm = p̃m yields a set of natural
frequencies and mode shapes:
(i)
which can be partitioned with one arbitrary degree of freedom φk separated out
as follows:
| | 1 (i)
Z11 |· Z1k |· Z13 −·−
−·−·− ·−·−·−·−·−·−
| |
Zk1 · Zkk · Zk3 φk = {0} (7.6)
−·−·− |· −·−·−·−·−·−
|
| |
−·−
Z31 · Z3k · Z33
| | 3
(i)
v1
−·−
∂(i)
= 0 + ci {(i) } = {v (i) } + ci {(i) } (7.7)
∂pm
−·−
v3
(i)
v1
−·−
v3
Then the solution for this vector is obtained from the following simultaneous
equations:
| | |
Z11 |· Z13 (i) Z
11pm · Z 1kpm · Z13pm
−·−·− · −·−·− {v̂ } = −·−·− |· −·−·−·−·−·−· | {(i) } (7.8)
| | |
Z31 · Z33 Z31pm · Z3kpm · Z33pm
| | |
where [(i)C ]T [i.e., = {(i) }C ] is the complex conjugate of {(i) } and [S] is some
symmetric matrix (often taken to be the mass matrix). For this case the ci constant
is determined by
The derivatives of the [Zik ] partitionings, with respect to the parameters pm , require
not only the derivatives of the dynamic matrices [M] and [K] but those of the
eigenvalues as well, as per the prior development. Note also that the preceding
expressions for the eigenvalue and eigenvector derivatives are to be calculated
from the results of the eigenvalue solution for the particular set of modification
parameters pm = p̂m .
Fig. 7.1 Polar diagram of predicted 3/rev rotating lateral shear for an articulated
blade in high-speed flight.
such methodology is reserved for a subsequent chapter. One major hurdle to using
highly mathematical optimization schemes is that the results have to be “sold” to
the designer who has the responsibility for transforming the rotor design to the
real world.
An emerging technology is the use of appended devices based on special mate-
rials such as piezoelectric crystals and shape-memory alloys that are typically
activated electrically. The voltage levels are much higher than that typically used
for solid-state avionics. The advantage of using such devices, however, is that their
activation can be be made as harmonically variable as desired and can thereby sig-
nificantly impact on reducing rotor vibration inputs. The disadvantages are that
1) they require comparatively sophisticated implementation in going across the
nonrotating to rotating coordinate system boundary, and 2) a high degree of relia-
bility has to be maintained, which is difficult using sophisticated high-voltage-level
electronics.
Fig. 7.2 Schematic of the dynamic interaction between an existing structure and a
dynamic subsystem modification.
The basic approach to deriving the analytical testing equation is first to define the
interface load vector and introduce it into the mobility equation for the fuselage
(substructure A). When this load is so introduced, the new response is then the
required modified response:
a
qF YFF YFI f
= (7.14)
qI YIF YII fI
Note that, because the impedance matrix partitions were measured on the fuselage
alone, the deflection vector is identified as looking into substructure A [i.e., {∗ }a ].
The lower partition of the deflection vector can then be related to the deflection
vector looking into substructure B:
Then the previous equation can be rewritten to isolate the interface load vector:
a
qF YFF YFI f YFF YFI f YFI
= = + { fI }
qI YIF YII fI YIF YII 0 YII
a
qF0 YFI
= + { fI } (7.16)
qI0 YII
The second partitioning of this equation can then be combined with the preceding
result to give
From this equation the required expression for the interface load vector can be
found:
The derivation is then completed by using this expression to eliminate the interface
load vector from the preceding equation. The preceding expression is for a structure
that is being excited at some point(s) separate from either the critical response
points or the interface points. Thus, in the case of a rotorcraft vibration problem
the “old” responses at the points F and I are those resulting from excitation located
at some other point H, which might correspond to one of the rotor hubs. An
important assumption inherent in the preceding formulation is that the excitation
at point H does not change because of the modification at the interface point.
This is not always the case with rotorcraft problems, and the analytical testing
methodology has been expanded to treat this more general case.
Analytic testing predictions are then made for the nine motion coordinates shown
in Fig. 7.3.
The matrix of mobilities for these nine coordinates (the j index) relative to
forcing at locations Z200L and Z200R (the r index) is given by Eq. (7.28).
This matrix has the units of load factor per Newton for sinusoidal motion
generated by the blade outboard of the pendab attachment location. The operation
of the pendab is most meaningfully described by means of the load attenuation
characteristics with variation of the pendab frequency as typically depicted in
Fig. 7.6.
Note that optimum pendab performance is usually achieved at a tuned frequency
ωP , which is somewhat less than the excitation frequency, in this case the blade
passage frequency b. Note also that for tuned frequencies above the excitation
frequency, the pendab actually amplifies the hub loads.
The advantages of using pendabs are as follows:
1) Pendabs are relatively simple, both to analyze and design and to install.
2) Pendabs are effective in substantially reducing (if not eliminating outright)
the hub shears at the blade root.
3) For reasonable blade-root shear attenuation, without excessive pendab
response, the pendab should be made as massive as possible. However, the addi-
tion of substantial nonstructural mass to the rotorcraft design impacts negatively
on the payload productivity.
4) The presence of damping in the pendab hinge is inevitable; however, this
damping is detrimental to the operation of the pendab.
5) Pendabs typically reduce only vertical blade-root shears at one nondimen-
sional frequency. A separate pendab must therefore be installed for each harmonic
of blade-root shear to be attenuated.
6) Because pendabs must be mounted directly to the blade, they change ori-
entation with blade pitch angle. With this orientation change coupling (usually
adverse) invariably occurs with the blade in-plane motion.
A principal disadvantage is that they create aerodynamic drag and are thereby
somewhat detrimental to performance.
Fig. 7.7 Installation of rotor bifilar absorbers on the S-76 helicopter. (Reproduced
courtesy of the Sikorsky Aircraft Corporation.)
Fig. 7.8 Basic geometry of a rotor hub appended bifilar dynamic absorber.
The basic hardware components of an HHC system are 1) some form of instru-
mentation (e.g., accelerometers) to measure the vibratory load factors g at some
selected locations in the airframe; 2) an onboard dedicated computer (controller)
to interpret the values of the measured vibratory accelerations and to calculate the
appropriate values of the six components of higher harmonic pitch control angles;
and 3) a means, usually a set of hydraulic actuators, for implementing the values
of HHC angles determined by the computer. The calculation of the appropriate
(optimal) values of HHC within the computer can be achieved by a number of
existing optimal control methodologies. The calculation of the gradient matrices
for establishing the sensitivities of the components of vibratory g to each of the
six components of HHC pitch angles is typically achieved using some form of
parameter identification (e.g., Kalman filtering). These features are schematically
represented in Fig. 7.10.
Controller operation. The controller has two primary functions that must be
implemented in the computer at a sufficiently high sampling rate (typically several
times per second) to ensure an accurate and rapid response to a changing vibration
environment. First it must perform a parameter estimation to establish the vector
of responses and the gradient matrix, which are needed for the use of the optimal
control methodology. Specifically, the following equation defines the quantities
needed for implementing the optimal control theory:
Fig. 7.10 Schematic representation of a higher harmonic control system for vibration
reduction.
coefficients relating changes in {Zi } to the vector of HHC control angles {u}. Both
{Z0i } and [T ] are typically evaluated using Kalman-filtering techniques.
The second primary function that the controller must perform is that of calculat-
ing the optimal controls that minimize some cost function based on the measured
components of vibration. It must be stressed that the vibration environment of the
rotorcraft is subject to constant change caused by changes in the flight condition
and/or the initiation of maneuvers. The usual implementation of the HHC system
therefore involves the use of an adaptive control algorithm to achieve the optimal
control settings. The vibration level(s) are sampled at several times per second in
order to produce accurate updates of both the zero control vibration level {Z0i }
and the gradient matrix [T ]. These updates are then continually used to determine
the values of the HHC pitch angles needed for an optimal or minimal value of the
vibration-related cost function.
The advantages claimed for HHC are as follows:
1) The method is effective. Experimental implementations have verified the
ability to produce substantial reductions in vibration. The advantage of an active
vibration control system is that it does not require some residual value of vibration
in order to drive it.
2) The controller is generic in that it need not be sized to any one specific
aircraft.
3) The use of HHC has been shown to have negligible or even beneficial effects
with regard to both blade vibratory stresses and aerodynamic performance.
However, the balancing disadvantages with HHC are as follows:
1) One pays a weight penalty with the HHC system. The extra weight associated
with the computer, hydraulics, and associated servovalves, and the additional
airframe strengthening is not negligible and could be a decisive factor in some
designs for which payload productivity is a primary consideration.
2) Compared with the passive systems and devices available for the attenuation
of fuselage vibrations, the HHC system is quite complex and, therefore, subject to
considerations of reliability and maintainability.
3) Despite the increasingly more rapid computational speeds available with
microcomputers, the calculations are not instantaneous. Consequently, for some
high-frequency band the HHC system would not track random or otherwise rapidly
changing vibration levels well.
With the increasing compactness and speed of computation and the commen-
surate decrease in weight of microcomputers HHC approaches should become
more competitive with respect to weight, reliability, and maintainability. Wood
et al. report on successful flight tests of an OH-6A helicopter implemented with
a prototype HHC system that showed significant reduction in vibrations without
undue penalties in blade loads or aircraft performance. Production installations
of HHC systems will hinge on the vibration benefits obtainable vs the additional
system weight and complexity.
nW = Kδsteady = ngM
(7.25)
δsteady = gn/ωn2
Thus, the need to limit the steady deflection places a constraint on the natural
frequency that can be designed into any device that relies on frequency tuning for
vibration attenuation. With that caveat in place, we now consider a basic device that
again relies on frequency tuning but has properties that are especially attractive.
The results of Fig. 7.13 present a mixed blessing however. The most effective
configuration for creating antiresonance is clearly one with very little or no damp-
ing, ζa [=ca /(ca )crit ]. In this condition, however, instead of one high response
condition where the forcing frequency is equal to the natural frequency, we now
have two resonant responses straddling the original resonance condition. The addi-
tion of damping does reduce the two straddling resonances but at the price of
diminishing the effectiveness at the antiresonance point.
The figure also shows an interesting and significant property of the system.
The two points labeled A and B are known as the fixed points and are actually
independent of the level of damping. The frequency location values of these points
are determined solely by the ratio of individual frequencies f (=ωa /ω0 ) and the
mass ratio. Rao presents formulations showing that the abscissas (i.e., frequencies)
of these points ωA /ω0 and ωB /ω0 , respectively, are then given by the solutions rA
and rB of the following quadratic equation:
1 + f 2 + µf 2 2f 2
r 4 − 2r 2 + =0 (7.28)
2+µ 2+µ
Tuned absorber. Of even greater importance than the frequency locations are
the transmissibilities (ordinates) of the two fixed points. As shown in Fig. 7.13,
where the frequency ratio is unity, these transmissibilities are quite dissimi-
lar. However, a tuning condition exists wherein the transmissibility values of
both points can be made equal. This condition defines a unique configuration
denoted the TVA (tuned vibration absorber) and a unique relationship between
1
f = (7.29)
1+µ
For the tuned condition represented by Eq. (7.28), the transmissibility value for
both the A and B points is given by
2
TA&B = 1 + (7.30)
µ
Optimum damping. Examination of Fig. 7.13 further shows that at each of the
two fixed points a proper selection of damping can be made such that the transmis-
sibilities can be minimized to the values given by Eq. (7.30). Thus, for the tuned
absorber optimum damping values can be found such that the transmissibilities are
bounded by those values defined for the tuned condition. Although the bounded
values cannot generally be met for each of the fixed points simultaneously for the
same value of damping, the average damping value proves to be quite practical for
design purposes:
3µ
(ζa )optimal = (7.31)
8(1 + µ)3
Fig. 7.14 Optimally damped tuned dynamic absorber characteristics for mass
ratio of 0.2.
Fig. 7.15 Optimum damping ratios as functions of mass ratio for the tuned dynamic
absorber.
The figure shows that the differences between damping optimized separately
for points A and B are not much different from each other or from the average
value given by Eq. (7.28). Figure 7.15 presents the variations with mass ratio of
the optimum damping ratios, as separately optimized for points A and B, and for
the average.
The use of this equation is of most use in a design that is both tuned with
respect to f and incorporates optimal damping. Figure 7.16 presents for a range
of frequency ratios straddling the resonance point the amplitude response ratio of
that of the absorber mass to that of the parent mass for these conditions.
The use of the results of Fig. 7.16 requires first obtaining the tuned response
of the parent mass |y1 |, which requires only the use of Eq. (7.30) for the tuned
condition.
Fig. 7.16 Variation of the amplitude response ratio for the absorber mass with mass
ratio and selected excitation frequency ratios—the optimally damped, tuned condition.
Similar to the results for the dynamic absorber, the transmissibility at the fixed
point can be determined:
2+µ
TB = (7.35)
µ
The transmissibility defined by Eq. (7.35) thus defines the upper bounded value
obtainable with the optimum damping given by Eq. (7.34). Figure 7.18 presents
a comparison of the transmissibilities obtainable with optimum damping for the
use of an auxiliary mass either as a tuned dynamic absorber and as a damper-only
supported mass.
The figure clearly shows the more advantageous transmissibilities available
with the dynamic absorber vis-à-vis the damper-only attached mass. The figure
does show, however, that with a sufficiently large auxiliary mass comparable
transmissibilities can be achieved with the damper-only configuration.
where I is the mass moment of inertia of the arm containing the m2 element about
its mass center. Figure 7.21 contrasts the transmissibilities of the basic simple
isolator and of the DAVI for the same installation for which a static load of 100 lbf
produces a static deflection of 0.11 in. Note that for the set of parameters used in
this comparison the spring stiffness K is the same for both isolators. Consequently,
the resonant frequency for the DAVI is always lower than that for the simple
isolator. There then exists a range of frequencies, starting from some point between
the two resonant frequencies to a point above the DAVI antiresonant frequency,
wherein the DAVI will have a lower transmissibility than that of the simple isolator.
However, for sufficiently high frequencies the simple isolator will have superior
performance. Finally, for some configurations the tuned antiresonance condition
of the DAVI can be so narrowly frequency-banded as to be unusable, especially for
those cases wherein the excitation frequency is subject to relatively wide fluctua-
tions. A graphic demonstration of the ability of the DAVI to suppress vibration is
given in Figs. 7.22 and 7.23.
Note in these figures that the DAVI units (at the four corners of the platform)
exhibit significant motion in the arm parts. Although the lower base of the platform,
as well as the DAVI arms, register as blurs in the photographs, the upper (isolated)
platform is quite smooth.
IRIS rotor isolation system. The improved rotor isolation system (IRIS) is
a passive system developed by the Boeing Helicopter Company. This system is
shown schematically in Fig. 7.24 and is seen to be somewhat a combination of the
nodamatic and the DAVI systems. It would appear to possess the antiresonance
characteristics of the DAVI without the need for low damping bearings, but it would
also appear to share the relatively high weight penalty disadvantage associated with
the nodamatic system.
Fig. 7.23 DAVI pilot seat isolation laboratory test, ±1.0 g at 10 Hz. (Reproduced
courtesy of Kaman Aerospace Corporation.)
Focused pylon. Whereas most of the devices discussed earlier are directed
at eliminating vibratory shears, the focused pylon device is directed at eliminating
vibratory moments. As shown in Fig. 7.25, the device consists principally of a
means of supporting an inertia with nonparallel linkages so that its motion is
constrained to occur about the point at which the support linkages are “focused.”
Usually, the focal point is taken to be the center of gravity of this inertia so that the
inertia will be forced to act as a “rotational” isolator. As with the simple passive
(linear motion) isolator, the focused pylon will function best with a minimum of
damping in the system.
References
Section
7.1 Nelson, R. B., “Simplified Calculation of Eigenvector Derivatives,”
AIAA Journal, Vol. 14, No. 9, 1976, pp. 1201–1205.
7.3 Flannelly, W. G., Fabunmi, J. A., and Nagy, E. J., “Analytical Testing,”
NASA CR-3429, 1981.
Jetmundsen, B., Bielawa, R. L., and Flannelly, W. G., “General-
ized Frequency Domain Substructure Synthesis,” Journal of the
American Helicopter Society, Vol. 33, No. 1, 1988, pp. 55–64.
Loewy, R. G., “Helicopter Vibrations: A Technological Perspective
(The AHS Alexander A. Nikolsky Honorary Lecture),” Journal of
the American Helicopter Society, Vol. 29, No. 4, 1984, pp. 4–30.
7.4 Desjardin, R. A., and Hooper, W. E., “Antiresonant Rotor Isolation for
Vibration Reduction,” Journal of the American Helicopter Society,
Vol. 25, No. 3, 1980, pp. 46–55.
Flannelly, W. G., “The Dynamic Antiresonant Vibration Isolator,”
Proceedings of the 22nd Annual National Forum of the American
Helicopter Society, May 1966.
References (continued)
Paul, W. F., “Development and Evaluation of the Main Rotor Bifilar
Absorber,” Proceedings of the 25th Annual National Forum of the
American Helicopter Society, May 1969.
Taylor, R. B., and Teare, P. A., “Helicopter Vibration Reduction with
Pendulum Absorbers,” Journal of the American Helicopter Society,
Vol. 20, No. 3, 1975, pp. 9–17.
Halwes, D. R., “LIVE—Liquid Inertia Vibration Eliminator,” Proceed-
ings of the 36th Annual National Forum of the American Helicopter
Society, Paper 80-22, May 1980.
Reed, F. E., “Dynamic Vibration Absorbers and Auxiliary Mass
Dampers,” Shock and Vibration Handbook, 3rd ed., edited by
C. M. Harris, McGraw–Hill, New York, 1988, Chap. 6.
Rao, S. S., Mechanical Vibrations, Addison–Wesley, Reading, MA,
1995.
Wood, E. R., Powers, R. W., Cline, J. H., and Hammond, C. E., “On
Developing and Flight Testing a Higher Harmonic Control System,”
Journal of the American Helicopter Society, Vol. 30, No. 1, 1985,
pp. 3–20.
Problems
7.1 For a certain dynamic multiple-degrees-of-freedom (zero-damped) system
the mass and stiffness matrices are given to be the following functions of a
parameter x:
10 + 2x x−1 3
x−1 x2 + 5 2
[M] =
√
3 2 10 5x
10x 3 x−5
[K] = 3 4x 2 x 2 − 21
x−5 x 2 − 21 2.4x 3
Natural frequencies:
Mode shapes:
1.0000 −0.5287 −0.2436
(1) (2) (3)
[φ , φ , φ ] = 0.2393 1.0000 −0.0360
0.0412 −0.0206 1.0000
7.2 Verify (using numerical means) the derivatives found in Problem 7.1. [Note
that access to a reliable eigenvalue solution (computational software) is
required for this problem.]
7.3 Using the basic analytical testing equation, we might wish to specify a
new (required) response vector {qFR }, thereby defining a specific change
in response; the problem is then to solve for the mobility (or inversely
impedance) at the selected interface location that will achieve the required
response change. Show that a solution for this required interface mobility
change is expressible as
where
and
Right pseudoinverse:
!−1
[B]# = [B]T [B][B]T (7.40a)
Left pseudoinverse:
!−1
[C]# = [C]T [C] [C]T (7.40b)
7.4 Derive (linear) simplified equations of motion governing the pendular motion
(βP ) of a rotor-appended pendular vibration absorber (pendab) attached to a
rotor blade whose flatwise bending flexibility is described by a single flatwise
bending mode generalized coordinate (qw1 ). The following assumptions can
be made:
be a point mass with an effective mass of 1.86 lb-s2 /ft. It is proposed to use
a 30-lb battery as a dynamic absorber to reduce the panel vibrations.
(a) What is the best vibration level in the panel that can be achieved with
this configuration?
(b) What attachment spring and damper rates are required to tune the
absorber mass (the battery) to achieve this level of attenuation?
(c) What is the maximum steady deflection of the absorber for a maximum
load factor n of 2.5?
(d) What is the maximum vibratory amplitude the absorber will
experience?
As the previous chapters have stressed, the ability to analyze rotorcraft for
vibration requires a good knowledge of the dynamic characteristics of the fuselage.
These characteristics generally include, as a minimum, the natural frequencies of
the airframe in the general frequency range of the blade passage frequency. Of
increasing importance are the mode shapes, the structural damping descriptors,
and the various mobility matrices [i.e., frequency-response functions (FRFs)] in
the frequency range of the fundamental blade passage frequency. Until such time
as the dynamicist has the tools (and the skill for using them) for accurately calcu-
lating these dynamic characteristics, some form of vibration testing will remain as
part of the development of new rotorcraft configurations. An adjunct to the direct
vibration testing of the airframe and airframe components is the testing for compo-
nent fatigue life. Just as in the case of natural frequency and mode prediction, the
prediction of component fatigue life is an inexact science that requires some form
of experimental verification. These two related types of vibration testing are sim-
ilar to the extent that they deal with structures operating in an environment that is
almost wholly in the frequency domain; therefore, some of the techniques overlap.
following sections of this chapter define and describe the various basic parts of
any shake test program.
Figure 8.1 presents the simplest form of shake testing: single-point excitation.
The application of force at only one point at a time does not, however, rule out
measuring responses at several locations. For rotary-wing aircraft, the use of single-
point excitation is quite appropriate in most cases. Single excitation at the rotor
hub closely simulates actual rotorcraft environment in that the rotor is the principal
source of excitation of the fuselage. This type of testing has the principal advantage
that it is a minimum effort approach. Single-point excitation does have somewhat
of a disadvantage when it comes to measuring modal characteristics in that this
type of testing inherently measures response amplitudes rather than mode shapes.
There are two principal reasons for this discrepancy. First, the excitation of the
structure at a given point will excite any mode that has a natural frequency close
to the excitation frequency, and so there will always be an admixture of modal
responses throughout the range of test frequencies. The second reason is that the
test structure will always have a degree of internal damping. This damping is
the mechanism whereby the structure dissipates the vibrational energy that the
shaker is imparting to it. This dissipation will distort the measurement of modal
characteristics because the points nearest to the excitation can be expected to
absorb the most energy. Thus, if the measurement of mode shapes is the desired
objective, there will be a limit to the accuracy that single-point testing can provide.
The alternative to single-point excitation is that of multipoint excitation wherein
a multiplicity of shakers are simultaneously driven to make the structure respond
in the desired modes. The full description of multipoint excitation systems is
beyond the scope of this text, but a description of the basic approach is given in a
subsequent section.
provide. In this manner the shake-tested fuselage will respond with resonances
and antiresonances reasonably close to those that the free-flying aircraft would
manifest.
Fig. 8.3 Alternate operational modes of fuselage suspension for shake testing using
an air-bag configuration. (Reproduced courtesy of Bell Helicopter Textron, Inc.)
Fig. 8.4 Typical helicopter fuselage shake test setup using an air-spring suspension
system. (Reproduced courtesy of McDonnell Douglas Helicopter Company.)
The principal excitation quantities are the five components of load at the hub
[the three components (x, y, and z) of vibrational force, and the roll and pitch
components of vibrational moment] and at selected fuselage locations that are
either locations of secondary sources of vibration (e.g., the tail rotor hub) and/or
are to be considered to be interface locations for potential vibration suppression
modifications. Thus, the linkage between the actuator and the hub should contain
three principal elements: 1) where appropriate, an elastic attachment element such
as a coil spring (or torque tube for moments) to convert relatively large actuator
strokes into controlled forces (or moments); 2) a load cell fixture located close to the
airframe to measure the applied load; and 3) a drive rod or appropriate appliance to
connect the elements together. The actuator system should be configured for high
performance by the selection of a sufficiently large, high performance servovalve
for the hydraulic systems. The shakers should be capable of adequate excitation at
frequencies somewhat in excess of twice the maximum frequency of interest in the
fuselage structure. For main rotor hub excitations this frequency is approximately
2 × 25 or 50 Hz. For tail-rotor excitation purposes this frequency can run about
four times greater.
8.1.2 Instrumentation
Basic measurement devices. The two types of basic measuring devices
needed for shake testing are load cells and accelerometers. The load cells are
part of the excitation equipment as discussed earlier and are used not only as
part of a mobility (transfer function) measurement, but also to ensure that suffi-
cient energy is imparted into the structure at all frequencies tested. With each of
these devices (accelerometers and load cells), measurement is made using either
of two physical principles. One principle is typically to use a flexible member
that is instrumented for elastic deflection: Either a conventional strain-gauge or a
piezoresistive device is used, which deflects in response to an applied or inertial
load. The second principle is to use the strain of a suitable crystalline material
directly (piezoelectric devices). Strain-gauged units have the advantages of being
relatively inexpensive and “tried and true.” They also have the advantage of being
able to measure both steady and vibratory loads or accelerations. Piezoelectric and
Advanced measuring devices. Whereas the required load cells are rel-
atively few in number and are relegated for use in the excitation system, the
number of accelerometers used in a shake test can be quite high (often more
than 100). The number of accelerometers is dictated by considerations of being
able to identify responses at key locations in the structure such as the pilot seats,
engines, and hubs; being able to describe fundamental modes; and being able to
identify local modes relating to specific components such as wings and pylons.
Another reason for the large number of required accelerometers is the fact that
most accelerometers typically measure only linear motion, and suitable rotational
motion accelerometers are relatively new and hence somewhat expensive. Thus,
rotational measurements are usually obtained indirectly using multiple accelerom-
eters appropriately positioned near each other. Another advancement currently
being made in vibrational measurements is the laser/fiber-optic vibrometer. Such
vibrometers operate by illuminating the vibrating structure with laser light and
using optical phenomena (Doppler frequencies and interference fringes) to mea-
sure vibrational amplitudes of both displacements and velocities. Their advantages
are that they can measure with significantly improved sensitivities and accuracies
relative to accelerometers, they do not need repeated calibration, and they can
measure the vibratory responses at any location on the structure that can be illu-
minated with the laser beams. Additionally, they can measure either in an absolute
motion mode or in a differential motion mode (using two optical sources). The
disadvantages of the new devices are that they are presently expensive, they are
not as rugged as comparable accelerometer-based measurement systems, and they
cannot directly measure load.
Acceleration and force will usually constitute the two quantities measured, as
shown in Fig. 8.1. Because the principal output from the shake test will be the
determination of frequency response functions [i.e., motion (acceleration)/force],
the voltage from the accelerometer will ultimately be divided by the voltage of
the force transducer. One practical way to ensure the proper calibration of the
transducers is to use a calibration mass, as shown in Fig. 8.5.
The idea is to excite the mass with the shaker and measure the voltages from
the two transducers. Because the mass is known (by weighing), the following
expression can be formed:
ẍ −1 Sẍ Vẍ Vẍ
=M = =S (8.3)
f Sf Vf Vf
If the voltages are divided before the calibration factors are employed, then the
preceding equation provides a way of evaluating the all-over calibration factor
S. If the calibration factors are used before the division, then either of the sepa-
rate calibration factors can be calibrated relative to the other using either of the
following expressions:
Vẍ
Sẍ = MSf (8.4)
Vf
or
Vf Sẍ
Sf = (8.5)
Vẍ M
where, in both of these equations, everything to the right of the equal sign is either
known or assumed to be correct.
Data acquisition. Data of the type that are typically acquired during shake
tests can be stored on magnetic tape (either in FM, direct analog, or digital forms)
and/or some similar form of high-density data storage device. State-of-the-art
data acquisition systems typically provide for computer control of the test oper-
ation, data acquisition, and often some sort of online data reduction. A variety
of approaches is presently employed and is to a large extent driven by the ever-
increasing availability (and affordability) of computer-based equipment on the
market. Some of the major features of such equipment are described in the next
section.
Fig. 8.7 Typical block diagram of a dual channel dynamic signal analyzer, spectrum
averaging mode.
FFTs are digitally calculated with essentially the same algorithm and are thus
available with the same speed and accuracy characteristics.) This is important with
regard to all of the digital signaling processing required for modern frequency anal-
ysis. High-speed implementations of FFTs on microprocessors with architecture
specialized for digital signal processing (DSP), and FFT computations in partic-
ular, are now fairly standard, and later we shall see that this algorithm forms the
heart of the analysis and processing capability of these analyzers.
4) Means are provided for producing a variety of specialized excitation signals
for driving the shakers. Thus, in addition to a straightforward sweep of stationary
frequencies with incrementally indexed frequencies, a variety of alternate non-
stationary excitations are provided; these are discussed in greater detail in a later
section.
5) Another feature present in many models is built-in “firmware” relating to
modal analysis, wherein typical modal parameters such as natural frequency, mode
shape, and damping are calculated (e.g., see Sec. 2.4).
Figure 8.7 summarizes the basic functions of the spectrum analyzer.
1
N
SAA ( fk ) = A( fk )n × A∗ ( fk )n (8.6)
N
n=1
1
N
SBB ( fk ) = B( fk )n × B∗ ( fk )n (8.7)
N
n=1
1
N
SAB ( fk ) = A( fk )n × B∗ ( fk )n (8.8)
N
n=1
1
N
SBA ( fk ) = B( fk )n × A∗ ( fk )n (8.9)
N
n=1
(1) SAB ( f )
HAB ( f ) = (8.11)
SAA ( f )
and
(2) SBB ( f )
HAB ( f ) = (8.12)
SBA ( f )
Significant differences between these two versions of the frequency-response
(1)
function relate to the presence of noise in the system. It can be shown that HAB
tends to minimize noise in the output of the process. In the case of vibration testing,
this takes the form of sensor and amplifier noise. Conversely, it can also be shown
(2)
that HAB tends to minimize noise in the input of the process. Such noise can take
place when the system is vibrating near resonances and/or when the force levels
are low relative to background excitations. The principal importance of these two
estimates for the frequency-response function is that they provide a measure of
the linearity of the measured response. This measure is the coherence function
described in the next subsection.
Fig. 8.8 Stanford Research Systems Model SR785 Dynamic Signal Analyzer.
(Reproduced courtesy of Stanford Research Systems.)
Fig. 8.10 Brüel and Kjaer Model 3560-L Pulse Lite Pocket Analyzer. (Repro-
duced courtesy of Brüel and Kjaer Instruments, Inc.)
Fig. 8.11 Graphical determination of generalized hub masses and stiffnesses for
low-frequency modes involving lateral hub motion.
the slope of the resulting line is proportional to (1/Keff ), and the effective mass is
given by Keff × (1/ω2 ) for (δM = 0).
Discrete (fixed) frequency excitation. This form of testing is the simplest and
represents the mainstay method for many years. The method consists very simply
of exciting the structure at a discrete frequency until a steady oscillatory response
has developed, taking a number of data records, and then proceeding to the next
frequency, etc. One technique that is used to increase the accuracy of the data reduc-
tion is to acquire data at a rate to ensure that the sampling rate exactly divides the
current oscillatory period by some integer (typically some power of two: e.g., 16).
With this kind of data sampling, inaccuracies caused by leakage effects are mini-
mized. The advantages of this type of testing are 1) it provides a direct control of the
accuracy [the presence of higher harmonics of response (caused by nonlinearities
and/or transients) can be monitored for subsidence], and 2) it is relatively sim-
ple and easily implemented. The disadvantages are that it can be time consuming
(and, commensurately, costly) and it does not allow for the use of the statistical
data checking algorithms commonly available with the more complex excitations.
Burst chirp and burst random. Both of these excitation methods entail the
supplying of the signal only during a specified percentage of the time record and
not for the remainder of the record; usually there is no signal at the start of the
record and at the end. During the nonzero record, the signal is either swept sine or
random.
The more advanced methods of testing are clearly the “wave of the future,” as
noted in present design standards (e.g., ADS-27, which defines appropriate test
procedures for future military helicopter development programs). Additionally,
as new concepts of vibration become formulated, new methods of testing will
emerge. One example of such a new concept is that of chaos, which can be thought
of as being just a more complicated form of orderly motion subject to its own
appropriate mathematical descriptors. When such descriptors are well developed,
they will most certainly be implemented with dedicated algorithms in the frequency
analyzers of the future.
Mathematical basis. The basis for this type of testing results from a con-
sideration of the general multiple-degree-of-freedom equation set defining the
dynamics of any structure:
[M] Ẍ + [C(ω)] Ẋ + [K] {X} = {F} eiωt = {Fs } sin ωt (8.15)
For sinusoidal motion the velocities will be in quadrature (90 deg out of phase)
with the displacements and accelerations. We seek a condition wherein all of the
components of the displacement vector {X} have the same phase and are then out
of phase with the accelerations. This condition defines a quadrature condition with
the velocities:
{X} = {Xc } cos ωt (8.16)
Ẋ = ω {Xc } sin ωt (8.17)
The requirement for these conditions is that the excitation force vector {F} be
exactly equal to and in phase with the damping terms:
ω [C(ω)] {Xc } = {Fs } (8.18)
At these conditions the response vector then defines the undamped natural mode
shapes, and the excitation frequency is the undamped natural frequency:
[K] − ω2 [M] {Xc } = {0} (8.19)
Fig. 8.12 Construction for evaluating the loss factor (after Zaveri).
the maximum spacing of frequency on the the arc. Using this construction, the
loss factor (taken to be twice the damping ratio) is then given by
ωA − ωB
η = 2ζ = (8.22)
ω0
The second method is based on the experimental data having been taken at
equal frequency intervals (ω and the angle α is taken to be that angle that defines
a maximum arc length). The natural frequency ω0 would be midway between ω1
and ω2 . The loss factor would then be given by
2ω 2(ω2 − ω1 )
η = 2ζ = = (8.23)
ω0 tan(α/2) ω0 tan(α/2)
Loading spectrum. The intensity of the loads applied to the structural ele-
ment is defined by the stress (or load) time histories appropriate to the projected
service life of the element. However, to save time, because the part must be tested
to the service life, the time histories should be applied to the structural element at
an accelerated timescale; this is justified because it is the number of cycles that is
important for fatigue, and not time, per se.
Number of cycles to be tested. Although the chosen service life can define
a spectrum of load levels, each with a corresponding number of maximum cycles,
the structural elements are usually tested at N times the number of cycles to ensure
that the parts that survive the cyclic loadings are not lucky scatter points on the
S-N curves. This number N is defined as the scatter factor. The magnitude of
this number depends on the requirements (usually) of the governmental agency
involved, for example, the Federal Aviation Administration, for commercial cer-
tification, and upon the amount of accumulated data available on the material of
which the part is manufactured. Some typical values of scatter currently in use are
shown in Table 8.1.
Material N
Steel 4
Aluminum 5–6
Titanium 8
Magnesium 8–10
References
Section
8.1 Broch, J. T., Mechanical Vibration and Shock Measurements, Larsen
& Son, Denmark (for Brüel and Kjaer), 1984.
Fabunmi, J. A., “Developments in Helicopter Ground Vibration Test-
ing,” Journal of the American Helicopter Society, Vol. 31, No. 3,
1986, pp. 54–59.
Huelsman, L. P., Active and Passive Analog Filter Design, An
Introduction, McGraw–Hill, New York, 1993.
Jones, R., Flannelly, W. G., Nagy, E. J., and Fabunmi, J. A., “Exper-
imental Verification of Force Determination and Ground Flying on
a Full-Scale Helicopter,” U. S. Army Applied Technology Lab.,
USAAVRADCOM-TR-81-D-11, Fort Eustis, VA, May 1981.
Nagy, E. J., “Improved Methods in Ground Vibration Testing,” Journal
of the American Helicopter Society, Vol. 28, No. 2, 1983, pp. 24–29.
Randall, R. B., Frequency Analysis, Larsen & Son, Denmark (for Brüel
and Kjaer), 1984.
“Standard Guide for Dynamic Testing of Vulcanized Rubber and
Rubber-Like Materials Using Vibratory Methods,”ASTM D5992-
96, 1996 (Reapproved 2001).
8.2 Kennedy, C. C., and Pancu, C. D. P., “Use of Vectors in Vibration
Measurement and Analysis,” Journal of the Aeronautical Sciences,
Vol. 14, No. 11, 1947, pp. 603–625.
Zaveri, K., Modal Analysis of Large Structures—Multiple Exciter
Systems, Naerum Offset, Denmark (for Brüel and Kjaer), 1985.
There are several ways of investigating instabilities. The fact that rotorcraft
dynamic issues must invariably be described by multiple-degrees-of-freedom
systems of equations drives our attention to mathematical techniques that can
deal with them. All too often these techniques, while giving accurate indications
of stability levels, also lead to a loss of understanding of the physics involved. At
the heart of any instability issue is the condition wherein there is an energy source
and one or more motion-induced force(s) that are in phase with velocities and grow
with response level. Perhaps the clearest example of this principle is the action of a
person on a swing. To initiate and increase the amplitude of the oscillatory motion,
the “person” must pull on the swing ropes at the bottom of the swing amplitude to
generate a gravity moment in phase with the angular velocity of the swing. This
approach to issues of instability forms the basis of examining the instabilities of
shafting (to be examined in a subsequent chapter), as well as forming the basis of
the method of force-phasing matrices, to be discussed in a subsequent section.
Other approaches to investigating instabilities require dealing with multiple-
degrees-of freedom equation sets and the use of sophisticated eigenvalue calcula-
tions. Within this broad class of methods are the methods associated with periodic
coefficients and with the marriage of frequency-domain data of one subsystem
with the analytic description of another. Both of these issues have application to
all aircraft aeroelasticity analysis, but show significantly greater application to
rotorcraft than to other types of aircraft.
In this chapter we restrict our attention to looking at stability analysis associ-
ated with linear systems, which precludes investigating a wide class of limit-cycle
unstable responses. The linear instability problem is perhaps most crucial, how-
ever, because the inception of unstable motion entails infinitesimal motion, which
inherently defines a linear process. In the real world the examination of test data
to analyze a vibratory response problem can become quite difficult because of
the distinction between resonant responses and unstable ones. Also, in extreme
catastrophic cases, there might not be any intact hardware or data left to examine!
297
In the case of most aeroelastic and aeromechanical systems, the applied load
typically contains components that arise from aerodynamic loadings. Furthermore,
these loadings contain components that arise from a variety of external excitations
(i.e., because of a control angle input or a gust), and one or more that arise from
the response of the structure itself. These loads are classified as motion-induced
loads:
For most aeroelastic systems the motion-induced loads contain terms that are
proportional to both rate and deflection, the presence of which yields differential
equations of the following general and by now familiar form:
If, in the preceding development, we take the applied load to consist of only that
caused by an external excitation sinusoidally varying at a frequency of ωn , while
omitting that caused by the motion-induced loads, then the resulting differential
equation has the following form:
Stability of the system is ensured when all of the characteristic roots of the
characteristic equation have negative real parts (all σ are negative).
Complex characteristic root plane. For systems that are described by linear
differential equations whose (constant) coefficients are all real-valued, the charac-
teristic roots must necessarily occur as either real roots or as complex conjugate
pairs:
λj = λR (9.11a)
or
λj·j+1 = σ ± iω (9.11b)
For the latter case of complex conjugate characteristic roots, the differential equa-
tion defining the motion of the corresponding characteristic amplitude can be
written in the following alternate form:
q̈ + 2ζ ωn q̇ + ωn2 q = 0 (9.12)
which is equivalent to the following characteristic equation:
λ2 + 2ζ ωn λ + ωn2 = 0 (9.13)
and, upon solving for λ,
λ = −ζ ωn ± i (1 − ζ 2 ) ωn (9.14)
Thus,
σ = −ζ ωn ; ω= (1 − ζ 2 ) ωn (9.15)
where ωn is denoted the undamped natural frequency for the mode. These
quantities are graphically described in Fig. 9.1. The substitution of an exponential
solution [Eq. (9.8)] renders the time derivative to be powers of the character-
istic root λ. For this reason the exponential solution shares some of the same
characteristics of the Laplace transform (see Section 2.4). The characteristic
root plane, as herein defined, can alternatively be referred to as the Laplace
transform plane.
expressed as
where one of the quadratic factors is indicated in Eq. (9.16). The Lin iteration
scheme, as described by Hildebrand, presents a method for sequentially factoring
out the quadratic factors.
Other methods. A variety of additional methods are given in both Press et al.
and Hildebrand that demonstrate differing convergence characteristics.
Hence, each of the components of the complex valued characteristic equation must
separately be equal to zero also:
f (λ) = a0 λ5 + a1 λ4 + · · · + a5 = 0 (9.19)
Then
(ω) = a1 ω4 − a3 ω2 + a5 = 0
(9.20)
(ω) = a0 ω5 − a2 ω3 + a4 ω
or
= (a0 ω4 − a2 ω2 + a4 )ω = 0 (9.21)
The assumption is that λ = iω will be fulfilled only when (ω) and (ω) both equal
zero for the same value of ω. In general, they will not both equal zero because
λ = iω only at certain combinations of the system parameters. It is precisely these
combinations of parameters that are desired. To this end, a reasonable approach
is to solve for ω using each of the component equations and to vary the pertinent
system parameter until the two values of ω are equal. This procedure is illustrated
in Fig. 9.2.
Routh array. Both variants of the criterion are amply described in the litera-
ture, but the Routh array finds more application to the dynamic systems of interest
(λn ) a0 1 a2
* a4 . . .
↓ ↓
↓
(λn−1 ) a1 a3 a5 . . .
(a1 a2 − a0 a3 ) (a1 a4 − a0 a5 )
(λn−2 ) =A =B
a1 a1
(Aa3 − Ba1 )
(λn−3 ) ···
A
· ·
· ·
(λ0 ) ·
The array is continued until no nonzero terms remain. Although not necessary to
the use of the array, the associated power of the eigenvalue λ is indicated for each
row at the left of the array. The Routh array is then used in the following manner:
If all of the terms in the first column are of one sign, the equation has no roots with
positive real parts; the number of roots with positive real parts equals the number
of changes in sign.
a0 λ3 + a1 λ2 + a2 λ + a3 = 0 (9.23)
For stability,
Quartic:
a0 λ4 + a1 λ3 + a2 λ2 + a3 λ + a4 = 0 (9.25)
For stability,
can be used to form the auxiliary equation, which is the polynomial comprised
of the factorization of the characteristic equation containing the roots that are
symmetrically located about the origin. Generally, the auxiliary equation is of
even-numbered degree and takes for the power of its highest-degree root term the
power of the eigenvalue that would be written to the left of the array, as exemplified
in the representation just given. This auxiliary equation would then have only even
powers of the eigenvalue, a condition that would be necessary in order to have
the roots symmetrically placed about the origin. In some instances this auxiliary
equation is of small enough a degree that it can be solved to determine these
symmetrically placed roots.
Modifications for very lightly damped systems. For many cases wherein the
Routh array is used to obtain stability boundaries, variations of certain parameters
(usually damping) are to be made that include, in the limit, zero values. In such
instances either of two difficulties can develop. On one hand, the array can degen-
erate to the case discussed earlier wherein one row of the array is zero, in which
case the array construction terminates prematurely. On the other hand, one of the
calculated coefficients in the first column can take on a zero value thereby render-
ing the sign change criterion as undefined. For these two cases recourse can be
made to the use of a small fictitious parameter , which is then taken to be zero in
the limit (after the construction of the array).
As we have seen, the case of one of the rows of the array becoming zero
amounts to a condition wherein some of the roots are symmetrically located about
the origin. If, however, the roots are (artificially) displaced to the “right” in the
complex plane a distance , then the symmetry condition would be disrupted, and
the Routh criterion could then be applied. Thus, if in the characteristic equation
[F (λ) = 0] the following substitution is made
where λ̄ is now the “displaced” eigenvalue from the line σ = , then the new
characteristic equation in λ̄ will typically be “complete” in that the previously
zero-valued row will be finite. The coefficients will generally be proportional to
some power of . If the Routh array is then constructed as usual, the sign change
criterion can then be validly invoked with the provision that the parameter be
taken to zero in the limit. In such a way a sign change can be detected even though
the coefficients might be vanishingly small.
In the case of one of the coefficients in the first column becoming zero, this
coefficient can be replaced by , and the array construction can be continued as
usual. To be sure, some of the ensuing construction can involve divisions by . If
now in the limit, as goes to zero, the criterion is invoked, then some of the terms
can become vanishingly small and others become singular, but the sign change
criterion, as already defined, can still be applied with validity despite the size of
any of the terms.
used for making flutter calculations of fixed wings (e.g., the so-called V -g method).
Such methods have the following general characteristics:
1) Sinusoidal motion is assumed: λ = iω. One advantage of working in the fre-
quency domain is that the unsteady aerodynamic formulations currently available
for use in rotary-wing aeroelastic calculations are in large measure adaptations
of fixed-wing formulations and are generally most accurately formulated in the
frequency domain.
2) Artificial damping is introduced into the system by making the usual simpli-
fied structural damping representation assumption but leaving the actual damping
level g and the frequency of oscillation ω as the eigenvalue components of the
problem. Generally, the equations can be rearranged in order to reduce the (flutter)
eigenvalue to a single complex quantity:
Fig. 9.3 Structural damping and frequency required for neutrally stable motion vs
air speed.
mẍ + c1 ẋ + kx = y = c2 ẋ (9.30)
where c = c1 − c2 .
For this form of the equation, the appropriate transfer functions are as follows:
x 1
KG ≡ = (9.31)
y k + c1 iω − mω2
y
H ≡ = c2 iω (9.32)
x
Thus, the combined open-loop transfer function is given by
c2 iω
KGH = (9.33)
k + c1 iω − mω2
As shown in Fig. 9.5, the locus of this product forms a loop that encloses or does
not enclose the +1 point on the real axis depending on whether c2 is respectively
greater or less than c1 .
From the preceding equations we know that when c1 is equal to c2 the total
damping c is zero and the system is neutrally stable. This condition corresponds
to the case wherein the locus passes exactly through the +1 point.
Fig. 9.5 Loci of open-loop transfer functions KGH, showing stability characteristics.
which, again, can be evaluated to form the required locus. As developed more fully
in a subsequent section, this methodology can be generalized and applied quite
effectively to the general multiple-degrees-of-freedom problem.
where [M], [C], and [K] denote, respectively, the inertia, damping, and stiffness
matrices. This form, however convenient for defining the physics of structural
dynamics, is not in a standard (canonical) form as typically formulated for any of
a wide variety of standard matrix eigenvalue extraction routines. For this purpose
we first rewrite the preceding equation in the following “semicanonical” form:
M 0 −C −K λX
λ − =0 (9.38)
0 I I 0 X
With some matrix eigenvalue routines this semicanonical form is sufficient. With
others the full canonical form is required:
where
Note that the [A], [B], and [D] matrices are not generally symmetric, and the
resulting eigenvalues (roots) and eigenvectors (modes) are typically complex. For
the cases wherein these matrices are real-valued, the resulting complex roots and
modes will occur as complex conjugates.
K
{x} = {φ (k) }eλkt (9.42)
k=1
where λk denotes the kth eigenvalue and {φ (k) } denotes the corresponding kth
eigenvector. In general, both λk and {φ (k) } are complex-valued. The eigenvalues
λ (= σ ± iω), which give stability level and natural frequency information, are
obtained from the solution to the matrix eigenvalue problem.
Upon inserting the preceding general solution into the eigenvalue equations,
each row of the resulting equation set represents the equilibrium of forces acting
on a corresponding degree of freedom. Each equation can be written as the sum of
the mass, damper, and spring forces of the nth diagonal element degree of freedom
along with the remainder of the terms lumped together as a combined excitation
force fn :
For the usual, nonpathological case mnn , cnn , and knn are all positive numbers;
that is, each autogenous mass (i.e., when uncoupled from the others) is gener-
ally a stable spring-mass-damper system. Because the eigenvalue λk is generally
complex, the preceding equation can be interpreted as the sum of four complex
quantities or vectors in the complex plane that must, furthermore, be in equi-
librium. Assuming that for any complex pair the eigenvector with the positive
imaginary part is used throughout, we can state that the argument of the eigen-
vector θk is the angle by which the inertia force vector is rotated (counterclockwise)
relative to the damper force and the damper force is rotated relative to the stiff-
ness force. For unstable motion the real part of λk (i.e., σk ) is a positive number,
and, hence, θk will be less than π/2 (i.e., 90 deg). If a point in time is taken
when the velocity of the nth degree of freedom is pure real and positive (i.e., zero
dispacement), then the autogenous damper force i.e., −λk cnn φn(k) must corre-
spondingly have a negative real part. The same argument applies to the autogenous
inertia force.
Furthermore, if it is realized that the four vectors represent forces that are in
equilibrium and governed by the constraint already developed on θk , then the real
parts of the autogenous spring, damper, and inertia forces will all be imaginary
(the spring force) or negative; hence, the remaining lumped sum of all of the off-
diagonal terms must always have a positive real part. Figure 9.6 demonstrates this
argument; the four force vectors are shown in the complex plane for an unsta-
ble oscillatory mode [(ωk ) = σk > 0; θk < π/2] and pure imaginary stiffness.
Because the fn vector is a sum of all of the off-diagonal terms for the nth degree
of freedom, any of the individual terms within fn that has a positive real part must
be deemed a driver for that degree of freedom. Conversely, any of the individ-
ual terms within fn that has a negative real part can be deemed a quencher for
that degree of freedom. This fact thus provides the dynamicist with the comple-
mentary information regarding which parameters might be increased to stabilize
an instability.
The aforementioned interpretations of unstable motion can be quantitatively
implemented by forming herein defined force phasing matrices. For any unstable
Fig. 9.6 Force vector diagram for nth degree of freedom, kth mode.
Note that this formulation renders the diagonal terms in the stiffness force phasing
matrix to near zero values. In the case of an aperiodic instability (i.e., divergence),
the argument of θk will be zero. For the drivers to have a real part, the diagonal
spring force should be phased to be pure negative real. This would be accomplished
by substituting −1 in place of i in the denominators of Eqs. (9.44–9.46). Although
not strictly required for the methodology, the division by |kii | in the preceding
equations serves the useful normalization of the matrices relative to the diagonal
stiffness terms.
Use of force phasing matrices: energy flow paths. The interpretation and
usage of the force phasing matrices can be summarized as follows:
1) Identify the most active degrees of freedom from the eigenvector information
for the unstable mode in question.
2) Look for relatively large positive (+) values in the force phasing matrices
involving the most active degrees of freedom as identified from the eigenvector.
Such elements are the drivers for the unstable motion.
3) Of the drivers identified look for those that involve degrees of freedom that
mutually drive each other. Such drivers we denote critical drivers. As illustrated
in Fig. 9.7, such critical drivers would occur in the simplest form as off-diagonal
terms involving two distinct degrees of freedom, say, the nth and mth.
4) Thus, critical drivers would show up as relatively large (+) values in both
the ( )mn and ( )nm elements of one or more of the three force phasing matrices.
The interaction through these terms is defined herein as the energy flow path.
5) It is possible for damping terms on the diagonal to be zero, but unlikely
that the diagonal stiffness terms would be zero. The diagonal stiffness terms could
easily be negative, however. In such a case the [PK(k) ] force phasing matrix would
still return a near zero value for the diagonal terms. Thus, it would be appropriate
to examine the actual stiffness matrix to identify such a condition.
λ4 + (g12
2
+ k11 + k22 )λ2 + k11 k22 = 0 (9.53)
Solving for λ2 shows that both values of this quantity can be made negative for
sufficiently large g12 , thereby causing λ to have pure imaginary values (i.e., neutral
stability). Furthermore, it can be shown that this reasoning can be extended to those
cases wherein either one or the other of the k11 or k22 values is negative and the
other is positive, and/or the total system is composed of pairs of degrees of freedom
that can be so reduced to diagonal form.
Basic K-T-C result for positive-definite matrices. For the cases wherein
both the [M] and [C1 ] matrices are positive definite and where the [K] matrix
has no zero eigenvalues, the K-T-C theorem states the following: For a nonzero
(positive-definite) damping matrix [C1 ], the number of unstable roots of the total
system is equal to the number of unstable roots of the truncated system. That is,
the addition of damping cannot stabilize an unstable system and will destabilize a
system neutrally stabilized by gyroscopic moments.
The basis of the theorem proof is that the presence of damping denies the system
from having purely imaginary roots (λ = ±iω). Therefore, any roots that reside
in the right half-plane with zero damping and gyroscopics must remain in that
half-plane with the addition of damping and gyroscopics.
where the [M] matrix is generally, but not always, constant and the damping and
stiffness matrices are periodic:
[C(t + T )] = [C(t)]
T = 1/ (9.55)
[K(t + T )] = [K(t)]
where {X̃(ψ)} = {X̃(ψ + 2π)}. When this representation of the trim responses
(assuming that they are somehow obtained) are inserted into the equations of
motion (including the nonlinearities), the following matrix equation of motion for
the perturbational responses is obtained:
∗∗ ∗ ∗ ∗
[M] {δ X } + [C̃(ψ, X̃, X̃)] {δ X } + [K̃(ψ, X̃, X̃)] {δX} = 0 (9.62)
In the preceding equation the [C̃] and [K̃] matrices are periodic, and the equa-
tion then represents a standard form of a set of linear equations with periodic
coefficients.
Trim solutions. The principal reason for making trim calculations is to obtain
the proper values of the control angles and rotor attitude angles that produce
the required values of the integrated hub loads (thrust, propulsive force, pitch-
ing moment, and rolling moment), that is, performance trim. A secondary reason
for obtaining trim calculations is to obtain the equilibrium responses {X̃(ψ)}, that
is, response trim. Typically, all trim calculations must be made iteratively because
both the trim force and moment relationships and the equations of motion are non-
linear; some form of numerical solution of the equations of motion is usually the
only practical way of getting accurate solutions. These two trim calculations (per-
formance and response) are sometimes referred to as major and minor iterations,
respectively. Within any one (major) iteration can be several minor iterations. With
each major iteration the calculated integrated hub loads are compared with those
required (or specified), and then variations of the controls are made (using gradient
methods) in order to drive the errors in these quantities to specified tolerances.
The problem of obtaining response trim solutions {X̃(ψ)} is a nontrivial one and,
indeed, one that can be approached from a number of ways. One straightforward
approach to the problem is to integrate numerically the full set of (nonlinear) equa-
tions of motion (with or without periodic coefficients) until the solution converges
to periodicity. This approach, although quite simple, has two serious shortcomings.
First, for lightly damped systems this process can take a long time to converge.
Second, because the entire equation set is being used, the solutions thus obtained
inherently contain not only the desired trim solutions but the perturbation solutions
as well. Thus, if a solution is indeed obtained in this manner, then it must nec-
essarily be stable because the perturbations will have died out. Conversely, if the
solution refuses to die out or actually diverges, then a trim solution is not obtainable
by this method, and the perturbational solution must be deemed unstable. Conse-
quently, this approach is merely doing what the Floquet theory was intended to do
(i.e., check for stability), but much more expensively.
One particularly intriguing approach is that of Peters and Izadpanah. This
approach still uses some form of numerical integration to solve the (entire) non-
linear equations of motions for the forced response case [i.e., retaining the full
∗
description for {F(ψ, X̃, X̃)}]. Solutions are obtained repeatedly, but each time
only over one rotor revolution, assuming a specified set of response initial con-
ditions at the start of each revolution. From these separate responses a partial
derivative matrix can be formed so that a truly harmonic solution for the responses
can be calculated independently of the response stability. This approach has the
dual advantages that it will not be overpowered by any inherently unstable pertur-
bational solution, and that a transition matrix, useful in itself for stability analysis,
falls out as a byproduct. Although the method could be used as a superior basis for
forming a minor iteration, the two iteration loops can be combined into a single
iteration scheme. Thus, in essence, the method treats the aeroelastic responses at
the end of each rotor revolution as additional trim quantities and treats the initial
conditions of these responses as trim variables (along with the control and rotor
attitude angles).
equation. Thus, we rewrite the basic matrix equation in the following form:
+N
∗
[A] {Y } + [Bn einψ ] {Y } = 0 (9.63)
n=−N
Fourier series ≡ [B(ψ)]
For [B(ψ)] to be real valued, [Bn ] and [B−n ] must be complex conjugates.
Floquet theory gives as a general solution to this equation the following
expression:
Note that the characteristic exponents in Floquet theory are distinguished from
the characteristic roots of systems with constant coefficients by the presence of the
multiples of 2π in the imaginary or frequency part of the characteristic expo-
nent. This multifrequency content can be interpreted as follows: for transient
motion defined by any one of the characteristic exponents η, the response would
be observed as having a basic frequency content ω (=ω ), with a multiplicity
of secondary frequencies defined by the two frequencies ω and |ω − |, each
± all integral harmonics of the frequency , defined by the period of the time-
variable coefficients, (= 1/T ). This phenomenon is referred to as an aliasing
of the frequencies. It is a real phenomenon that can be observed from attempts
to frequency analyze the responses of dynamic systems that are characterized by
periodic coefficients in their system matrices.
Although the presence of the frequency aliasing makes the interpretation of the
imaginary parts of the characteristic exponents difficult, the real of the charac-
teristic exponent σ is completely and unambiguously a measure of the stability
of the system. Just as in the case of systems with constant coefficients, stabil-
ity is ensured when the real parts of all of the characteristic exponents σk have
negative values.
Fig. 9.9 Idealized characteristic aerodynamic zones of the rotor disk in forward flight.
(k)
ψ2
2
[Ck ] = [C(ψ)] dψ (9.69)
π ψ1
(k)
(k)
ψ2
2
[Kk ] = [K(ψ)] dψ (9.70)
π ψ1
(k)
The principles presented in Section 9.3 can then be introduced to yield the following
general solution of the preceding matrix differential equation:
where the [k ] matrix of eigenvectors is n × 2n, where n is the dimension of {X}
and the number of degrees of freedom in the aeromechanical/aeroelastic formu-
lation. Both the columns of [k ] and the eigenvalues [λ(k) ] generally occur either
as complex conjugate pairs or as real-valued pairs. The column of undetermined
coefficients {C (k) } for each of the four zones is determined by directly equating
the displacements and velocities between zones 1 and 2, 2 and 3, and 3 and 4. The
solution is then completed by requiring that the displacements and velocities across
zones 4 and 1 (to complete the solution around the azimuth) be proportional with a
coefficient of proportionality . This coefficient of proportionality is, in fact, the
basic eigenvalue for this type of problem and is referred to as the characteristic
multiplier. These ideas are illustrated by the following equations, which define the
boundary conditions applicable to the solutions for each zone boundary:
π π
X (1) = X (2)
4 4
and
∗ (1) π ∗ (2) π
X = X (9.73a)
4 4
3π 3π
X (2) = X (3)
4 4
and
∗ (2) 3π ∗ (3) 3π
X = X (9.73b)
4 4
5π 5π
X (3) = X (4)
4 4
and
∗ 5π ∗ 5π
X (3) = X (4) (9.73c)
4 4
7π −π
X (4)
= X (1)
4 4
and
∗ (4) 7π ∗ −π
X = X (1) (9.73d)
4 4
The solution scheme is thus seen to comprise two levels of eigenvalue solution. First
there is the eigenvalue solutions within each zone to determine the eigenvectors
[k ] and the eigenvalues [λ(k) ], and second, the eigensolution for the characteristic
multipliers j , resulting from combining the respective undetermined for each
zone through the aforementioned boundary conditions.
Once this matrix is found, the stability is ascertained by finding the eigenvalues
of [(T , 0)], which are again denoted the characteristic multipliers j . These
eigenvalues are related to the characteristic exponents that we seek by the following
relationship:
j = exp(ηj T ) (9.75)
Thus,
1
ηj = σj ± iωj = n(j )
T
1 jI
= n|j | ± i tan−1 ± 2nπ (9.76)
T jR
which shows these quantities to have the same characteristic aliasing behavior as
that discussed earlier.
Two basic methods exist for calculating the transition matrix, both of which rely
on numerical integration of the equations of motion: the n-pass and single-pass
methods.
The n-pass method. In this method (n) time-history solutions are obtained
for the given periodic coefficient differential equations of motion. For each of
the (n) solution passes, one of each of the (n) initial conditions (including both
displacements and velocities) is systematically set to unity and all of the remainder
set to zero. By definition, the resulting response vectors obtained at the end of one
rotor revolution (period) are the respective columns of the transition matrix. Thus,
if there are N degrees of freedom in the dynamic system, then 2N passes must be
performed to obtain the transition matrix in this manner.
Single-pass method. In the former n-pass case the initial conditions were
incorporated explicitly in a one-at-a-time sequential manner. In the single-pass
method the initial conditions are incorporated all at once in a more implicit manner.
Although there are alternate methods for achieving a single-pass solution, one
basic requirement is that it leave the initial conditions in a completely general
form, that is, that the method be independent of startup conditions. Typically,
Runge–Kutta methods are ideally suited to this application, and one version that
is well suited to the single-pass method is the fourth-order Runge–Kutta method
with Gill coefficients briefly described in the following material. For this method
a canonical (first-order) form of the differential equations of motion is required.
Define the state vector {Y } as follows:
∗
{Y } = [X, X]T (9.78a)
With the single-pass method a numerical integration scheme must be used that
leaves the initial conditions as explicitly stated quantities rather than using them
where the matrix [K(ψi )] can be formed from the development given earlier. The
manner in which this matrix is used is as follows: first, it is formed for each of the
time steps (i = 1, 2, . . .) to form a sequence of the following matrices: [K(ψ0 )],
[K(ψ1 )], [K(ψ2 )], . . . . These are then cascade multiplied to form the required
transition matrix:
N
[(T , 0)] = [K(ψN−i )] (9.82)
i=1
The principal advantage of the single-pass method is that, for large-order dynamic
systems, it is computationally faster. Its principal disadvantage is that it locks
the solution algorithm into using a specific, fixed integration step size, and
methods that adjust the step size to achieve specified accuracy cannot be easily
used.
Equation (9.85) mandates that the responses at the end of the opened loop X 2
shall be proportional to those at the start of the loop X 1 . The undetermined constant
of proportionality at this point is allowed to be arbitrary and complex-valued.
λ = σ ± iω = −ζ ωn ± i 1 − ζ 2 ωn
The critical damping ratio is a convenient index that “scales” the level of sta-
bility and by extension instability. As presented so far, using only Eq. (9.86), the
use of the method of characteristic loci yields only information as to whether the
total system is stable or unstable. For many applications a more definitive answer,
as to how stable or unstable the system is, is required. For conditions near the
point of instability, it is possible to show equivalence between the two methods.
Essentially, the objective here is to recoup, at least approximately, the eigenvalue
results (i.e., λ = σ ± iω) from the characteristic loci results (i.e., = R ± iI ).
In a typical application of this methodology, the characteristic loci stability infor-
mation would be available only as numerical values of magnitude and phase of
for only discrete frequencies near the positive real-axis crossing. As described
earlier, numerical techniques can be used to obtain the actual real-axis cross-
ing values of ω and . However, in order to recoup, at least approximately, the
equivalent eigenvalue (characteristic exponent) results, a straightforward applica-
tion of analytic continuation can be performed about the real-axis crossing point
= 1 + ε + i0. This formulation begins by writing Eq. (9.86) in the following
more general form:
(λ1 ) {X} − [G2 (λ1 )] [G1 (λ1 )] {X} = {0} (9.87)
where, for the usual eigenvalue solution, λ1 = λ = σ ± iω, and (λ1 ) = +1.
However, the characteristic locus solution is known only for λ on the imaginary
axis: λ∗ = iω∗ , and (λ∗) = R + iI , where ω∗ is the frequency at the real-
axis crossing point (calculable using the techniques discussed in the preceding
subsection). The objective is to closely approximate the characteristic exponent
σ knowing the behavior of as a function of ω near the critical point. Note
first that, in the frequency domain, (iω*) will have the value of 1 + ε (where ε
is a calculable, but reasonably “small” quantity near the neutral stability bound-
ary). In the Laplace-variable domain, however, (λ) has the value of exactly 1
at the value of λ that we seek. Assuming that (λ) is an analytic function in
the neighborhood of (1 + i0), we can, therefore, expand (λ) in a Taylor series
taking the derivatives in the frequency (imaginary) axis direction and evaluating
them at the point λ = λ∗ = iω∗ . At the real-axis crossing point, we represent
the characteristic loci (λ∗ ) as (1 + ε). The analytic continuation equation then
becomes
d ∗ 1 2 d2 ∗
(λ) = 1 = 1 + ε + λ − λ∗ λ + λ − λ∗ λ + · · · (9.88)
dλ 2 dλ2
We note that, assuming continuity, the derivatives can be taken in any direction:
d dR dI dI dR
= +i = −i (9.89a)
dλ d(iω) d(iω) dω dω
2 2
d d d d R d2 I
2
= =− 2
−i (9.89b)
dλ d(iω) d(iω) dω dω2
and
λ − λ∗ ≡ σ + iω (9.90)
where ω = ω − ω∗ . Because (ω) would be known numerically only at a few
distinct points [corresponding to discrete values of ω near the (1 + ε + i0) point] in
the complex plane, a numerical curve-fit calculation must be made to evaluate the
derivatives. One efficient method of fitting the points to analytic approximations
to the real and imaginary parts of (ω) is the use of cubic-spline interpolation.
Because spline fit interpolation schemes fit the data points, as well as the second
derivatives, they also provide an excellent method for approximating first and
second derivatives. The use of cubic-spline fits would thus limit the ability to use
the analytic continuation defined by Eq. (9.88) beyond the second derivative term.
The most accurate estimate of the damping and frequency σ and ω would therefore
be limited to what can be achieved by retaining only the second-order derivative
term in Eq. (9.88):
1 d2 2 d
2
λ − λ∗ + λ − λ∗ + ε ∼
=0 (9.91)
2 d(iω) d (iω)
Solution of Eq. (9.91) is perhaps most practically accomplished using the
binomial theorem with complex arithmetic. It is to be expected that the accu-
racy of Eq. (9.91) to give an analytic continuation would be most precise only
when ε is small (i.e., is close to 1 + i0). For cases where the zero-phase
crossing is at a value significantly greater than unity, the inverse of is more
useful. In this case the analytic continuation occurs about the point −1 (iω∗ ),
where
ε
−1 iω∗ = 1 − ε̂ ; ε̂ = (9.92)
1+ε
The appropriate equation for solving for (λ − λ*), using the inverse of , then
becomes
1 d2 −1 d −1
2
λ−λ ∗ 2
+ λ − λ∗ − ε̂ ∼=0 (9.93)
2 d (iω) d (iω)
Equation (9.93) is quite useful in that, for any unstable characteristic loci (i.e.,
∗ = 1 + ε), the deviation of −1 (iω∗ ) from the point 1 + i0 is always less
than unity, thereby affording a more accurate analytic continuation. Consequently,
Eq. (9.91) would more appropriately be used for stable characteristic loci and
Eq. (9.93) for the unstable ones.
as the angle the locus makes with the real axis when the locus has unit value.
Either of these concepts can be meaningfully used to define a stability margin for
the multiple-degrees-of-freedom frequency-domain method.
Savant also defines an M criterion relating the nearness of the locus to the
singular point (in this case the +1 point). For present purposes, the M quantity
would most consistently be defined as the ratio of the distance from the origin to
the locus, to the distance of the locus from the +1 point:
(ω)
M = (9.94)
1 − (ω)
where the sign is taken to be positive for a stable system and negative for an
unstable one.
References
Section
9.2 Bisplinghoff, R. L., Ashley, H., and Halfman, R. L., Aeroelasticity,
Addison–Wesley, Cambridge, MA, 1955.
Dowell, E. H., Clark, R. C., Cox, D., Curtiss, H. C. Jr., Edwards, J. W.,
Hall, K. C., Peters, D. A., Scanlan, R., Simiu, E., Sisto, F., and
Strganac, T. W., A Modern Course in Aeroelasticity, Fourth Edition,
Kluwer Academic Publishers, Norwell, MA, 2004.
Franklin, J. N., “On the Numerical Solution of Characteristic Equa-
tions in Flutter Analysis,” Journal of the Association for Computing
Machinery, Vol. 5, No. 1, 1958, pp. 45–52.
Fung, Y. C., An Introduction to the Theory of Aeroelasticity, Wiley,
New York, 1955.
Hildebrand, F. B., Methods of Applied Mathematics, Prentice–Hall,
New York, 1952.
Savant, C. J., Jr., Basic Feedback Control System Design, McGraw–
Hill, New York, 1958.
Scanlan, R. H., and Rosenbaum, R., Introduction to the Study of
Aircraft Vibration and Flutter, Macmillan, New York, 1951.
9.3 Hurty, W. C., and Rubinstein, M. F., Dynamics of Structures, Prentice–
Hall, Englewood Cliffs, NJ, 1964.
MacFarlane, A. G. J., and Postlethwaite, I., “The General Nyquist Sta-
bility Criterion and Multivariable Root Loci,” International Journal
of Control, Vol. 25, 1977, pp. 81–127.
Zajac, E. E., “The Kelvin-Tait-Chetaev Theorem and Extensions,”
Journal of the Aeronautical Sciences, Vol. 11, No. 2, 1964,
pp. 46–49.
9.4 Horvay, G., andYuan, S. W., “Stability of Rotor Blade Flapping Motion
When the Hinges are Tilted: Generalization of the ‘Rectangular
Ripple’ Method of Solution,” Journal of the Aeronautical Sciences,
Vol. 14, No. 10, 1947, pp. 583–593.
Friedmann, P. P., Hammond, C. E., and Woo, T.-H., “Efficient
Numerical Treatment of Periodic Systems with Application to Sta-
bility Problems,” International Journal for Numerical Methods in
Engineering, Vol. 11, No. 7, 1977, pp. 1117–1136.
References (continued)
Peters, D. A., and Izadpanah, A., “Helicopter Trim by Periodic Shoot-
ing with Newton-Raphson Iteration,” Proceedings of the 37th Annual
National Forum of the American Helicopter Society, Paper 81-18,
1981.
Saaty, T. L., and Bram, J., Nonlinear Mathematics, McGraw–Hill,
New York, 1964.
Ralston, A., and Wilf, H. S. (eds.), Mathematical Methods for Digital
Computers, Wiley, New York, 1960.
9.5 Bielawa, R. L., “Frequency-Domain Methodology for Practical
Ground Resonance Analysis and Test,” Proceedings of the 28th
European Rotorcraft Forum, Paper 69, 2002.
Bielawa, R. L., “Practical Analysis and Testing of Rotorcraft Insta-
bilities Using a Frequency-Domain Based Methodology,” Proceed-
ings of HeliJapan 2002, AHS International Meeting on Advanced
Rotorcraft Technology and Life Saving Activities, Paper T222-3,
2002.
Press, W. H., Flannery, B. P., Teukolsky, S. A., and Vetterling, W. T.,
Numerical Recipies in FORTRAN, the Art of Scientific Computing,
2nd ed., Cambridge Univ. Press, New York, 1992.
Savant, C. J., Jr., Basic Feedback Control System Design, McGraw–
Hill, New York, 1958.
Tuma, J. J., Handbook of Numerical Calculations in Engineering,
McGraw–Hill, New York, 1989.
Problems
9.1 A two-degree-of-freedom system, which is defined in part by two parameters,
p1 and p2 , respectively, has the following dynamic equations of motion:
Determine (and sketch) the map of the two parameters, p1 and p2 , for which
the dynamic system defined by the preceding equations is stable.
9.2 Using the Routh array method, determine the stability (including numbers of
unstable roots, if present) of each of the dynamic systems whose characteristic
a) λ4 + 198λ2 + 10,201 = 0
b) λ4 + 20λ3 − 2,000λ − 10,000 = 0
c) λ5 + 3λ4 + 3λ3 + 9λ2 + 2λ + 5 = 0
9.3 Recast the Ripple method in a standard canonical matrix eigenvalue form
using the linear solution statements for each zone together with the zone
boundary conditions. Note: let
{X (k) (ψ)} = [k ] [\exp(λk ψ)\]{C (k) } = [Gk (ψ)]{C (k) } (9.96)
Similarly, let
∗
{X (k) (ψ)} = [Hk (ψ)]{C (k) } (9.97)
systems, which, although they do have some form of elastic shaft support, are
nonetheless relatively quite rigid. The key feature of the rotor is that it has
anisotropic inertias and/or stiffnesses either in the fixed coordinate system or
in the rotating shaft. Because the rotor description is relatively simple, the analysis
is applicable to the stability of unsymmetrical shafting.
1 (ka + kb ) + (ka − kb ) cos 2φ −(ka − kb ) sin 2φ
=−
2 −(ka − kb ) sin 2φ (ka + kb ) − (ka − kb ) cos 2φ
θξ
× (10.2)
θη
Using these expressions, together with the form of Euler’s equations given earlier,
we obtain the following basic equation set:
Iξ 0 θ̈ξ 0 1 θ̇ξ 2 (Iζ − Iη ) 0
+ (Iζ − Iξ − Iη ) +
0 Iη θ̈η −1 0 θ̇η 0 (Iζ − Iη )
θξ 1 (ka + kb ) + (ka − kb ) cos 2φ −(ka − kb ) sin 2φ
× +
θη 2 −(ka − kb ) sin 2φ (ka + kb ) − (ka − kb ) cos 2φ
θξ
× =0 (10.3)
θη
Iξ − Iη
i ≡ , −1 < i < +1 (10.4)
Iξ + Iη
Stiffness inequality:
ka − kb
s ≡ , 0≤ s ≤ +1 (10.5)
ka + k b
Iζ
A≡ , | i| < A < 1 (10.6)
Iξ + I η
(ka + kb )
ω0 = (10.7)
(Iξ + Iη )
It can then be seen that the stability equations and the resulting stability characteris-
tics depend on the values of only five nondimensional parameters: (/ω0 ), φ, i , s ,
and A.
If the time derivative is nondimensionalized, d/dt = d/dτ , then the following
nondimensional form of the equations of motion results:
∗∗ ∗
(1 + i ) 0 θξ 0 1 θ ξ
+ 2(A − 1)
0 (1 − i ) ∗∗ −1 0 ∗
θη θη
(2A − 1 + i ) 0
+
0 (2A − 1 − i )
(1 + s cos 2φ) − s sin 2φ θξ
+K =0 (10.8)
− s sin 2φ (1 − s cos 2φ) θη
where
α =1− 2
i (10.10)
β = A2 − 2
i + (1 − A)2 + (ω0 / )2 (1 − i s cos 2φ) (10.11)
γ = (ω0 / ) (1 − 4
s ) − 2(ω0 / ) (1 − 2A + i s cos 2 φ)
2 2
+ (1 − 2A)2 − 2
i (10.12)
It can be shown that, for ranges of the parameters involved, the first three conditions
are automatically satisfied. The condition γ > 0 is then the governing stability
criterion; this implies that the nature of the unstable motion is that of an aperiodic
divergence. The principal parameter subject to variation is the rotor speed . With
variations in this parameter, the rotor can have three different modes of behavior,
depending in part on the value of the rotor speed relative to two critical speeds
1 and 2 :
I. Stable for all speeds.
II. Stable at low speeds, but unstable for all speeds above the first of the critical
speed 1 .
III. Stable at low and high speeds, but unstable within a certain speed range
1 < < 2 .
The critical speeds 1 and 2 are obtained from the condition γ = 0 :
Thus, a condition that a critical speed exists at all is that the square of the rotor
speed, as determined from the preceding equation, be a positive number. These
results are summarized in Table 10.1. Figures 10.2–10.5 summarize the results of
using the stability criteria defined by inequalities presented by Eq. (10.13).
Fig. 10.2 Stability diagram for rotors with stiffness inequality and gyroscope coupling
but symmetrical inertia properties (after Crandall and Brosens).
Fig. 10.3 Stability diagram for rotors with inertia inequality and gyroscopic coupling
but symmetrical elastic properties (after Crandall and Brosens).
Fig. 10.4 Stability diagram for rotors with equal elastic and inertia inequalities;
maximum inertia axis aligned with minimum stiffness axis (after Crandall and
Brosens).
Fig. 10.5 Stability diagram for rotors with equal elastic and inertia inequalities;
maximum inertia axis aligned with maximum stiffness axis (after Crandall and
Brosens).
2 (1 − i ) < A < 21 (1 + i )
1
(10.15)
which is exactly the unstable triangular region shown in the limit as (/ω0 ) → ∞.
Figures 10.4 and 10.5 show the effect on stability of the angle φ. In general, it is
seen that the case that has the greatest spread between the two natural frequencies
is the least unstable.
dL 1 1
= ρcU 2 c = ρcU 2 (aα) (10.18)
dr 2 2
dD 1
= ρcU 2 cd (10.19)
dr 2
Fig. 10.6 Vector diagram showing component velocities at a typical blade section.
and the perpendicular and tangential components of the velocity UP and UT are
respectively given by
UP = − Rλ − ż − Rµ cos ψz (10.21)
UT = R(x + µ sin ψ) + ẏ (10.22)
where µ and λ are, respectively, the advance ratio (nondimensional forward-
flight speed) and the inflow ratio. The in-plane and out-of-plane components of
the airloads can then be respectively related to the lift and drag distributions as
follows:
dFy dL dD
≈φ − (10.23)
dr dr dr
dFz dL
≈ (10.24)
dr dr
The inflow ratio λ in the expression just given for UP can be expressed in its
most general form as a function of both span x(=r/R) and azimuth ξ . Some typical
ways of expressing the inflow are given in the following.
Uniform inflow. The simplest assumption that can be made about the inflow
is that it is both constant and uniform over the rotor disk:
λ = λ0 (10.25)
angle of attack can be used statically when the UP component of local inflow,
including pitch rate effects, is defined at the three-quarter chordwise location.
Thus, as shown in Fig. 10.7, two important aerodynamically relevant chord-
wise locations are defined: the quarter-chord point (i.e., the aerodynamic center),
wherein the airloads are considered to act, and the three-quarter-chord point,
wherein the angle of attack (including pitch rate effects) is defined.
Thus, the quasi-steady angle of attack αQS is again defined in terms of the
geometric pitch and inflow angles:
−1 UP
αQS = θ + tan (10.28)
|UT |
where, for rotations about an arbitrary point, a distance ySC from the quarter-chord
point (measured forward toward the leading edge), UP is then defined as follows:
This definition of the angle of attack then gives the appropriate impact of pitch
rate on lift and drag (as defined by the lift and drag coefficients). Furthermore, this
reasoning can be extended to the case of the moment coefficient as well. Thus, for
pitching motion about the chordwise point ySC the aerodynamic coefficients for
unsteady motion would be approximated as follows:
c c
cd = cd (10.30)
cmc/4 unsteady cmc/4 static, α = α
QS
Fig. 10.7 Pitching motion relevant geometry of blade section operating quasi-
statically.
Pitch rate effects on section moment. Apart from the effects of redefining
the angle of attack quasi-statically (for evaluating the steady moment coefficient),
an additional direct effect of pitch rate on the airfoil section moment must be
included. This effect takes the form of a direct pitch damping, which exists even
when the moment coefficient at the quarter-chord point cmc/4 is zero. Again, ref-
erence to classical two-dimensional unsteady airfoil theory gives the following
approximation for the total section pitching-moment distribution, as taken about
the rotation point ySC , defined in Fig. 10.7:
dMxSC 1
= ρcU 2 [cmc/4 (αQS ) + yAC c (αQS )]
dr 2
c3
− πρ U(1 − 2yAC /c) θ̇ (10.31)
8
direct pitch damping
This second term is not generally small, and, except for higher-frequency unsteady
effects, this term accounts for the principal source of damping available to the blade
in pitching and elastic torsional responses.
For some cases, especially those for which a full-blown nonlinear analysis is
needed, the airfoil section properties are expressable in terms of the tabulated
values of the airfoil coefficients. Such tables typically express the three airfoil
coefficients as discrete functions of angle of attack α and Mach number M. For
these cases the quasi-steady approximation then takes the following form:
c c (α, M)
cd ≈ cd (α, M) (10.33)
cmc/4 unsteady cmc/4 (α, M) static
Note that the two component velocities UP and UT are as given earlier (i.e., with
the inclusion of the pitch rate dependency).
x ≤ −µ sin ψ; −→ π ≤ ψ ≤ 2π (10.35)
dL
= 1/2 ρcc U 2 = 1/2 ρacαU 2
dr
= 1/2 ρacU 2 [sgn(U T )θ + φ]
≈ 1/2 ρac[UT2 sgn(UT )θ + |UT |UP ]
= 1/2 ρac sgn(UT )[UT2 θ + UT UP ] (10.36)
Similarly, while allowing for a more general form of cd , we can rewrite the drag
load distribution using the sgn(UT ) function:
dD
= 1/2 ρccd sgn(U T )UT2 (10.37)
dr
sgn(UT ) = 1 − ϒ (10.38)
where the ϒ function is as shown in Fig. 10.8. This function can be explicitly
stated mathematically as follows:
0; ψ < π + sin−1 (x/µ)
ϒ= ψ > 2π − sin−1 (x/µ) (10.39)
2; π + sin−1 (x/µ) ≤ ψ ≤ 2π − sin−1 (x/µ)
This function can then be made more tractable using Fourier analysis. Upon
noting that the function is periodic, we can write an appropriate Fourier series
representation for it:
∞
sgn(UT ) = (1 + I0 ) + [Ins sin nψ + Inc cos nψ] (10.40)
n=1
where
2π 2π−sin−1 (x/µ)
1 (−2)
I0 = (−ϒ) dψ = dψ
2π 0 2π π+sin−1 (x/µ)
2 x
= −1 + sin−1 (10.41)
π µ
Note that all Ins equal zero for even values of n and all Inc equal zero for odd values
of n. Thus,
4 µ2 − x 2
I1s = ; I2s = 0; etc. (10.44)
π µ
and
4 x 2
I2c = µ − x2 ; I3c = 0; etc. (10.45)
π µ2
The resulting Fourier series has the distinct advantage that, because the coefficients
are only spanwise dependent, the inclusion of this form of reversed flow allows
for the numerical evaluation of various (spanwise) integrals of the airloads. Such
a situation would arise in a Galerkin procedure, for example, wherein generalized
excitations are required that involve integrals involving modal shape functions,
which are typically discretely evaluated and are thus available only at a finite
number of spanwise locations.
However, before the perturbations are taken, let us first introduce refinements to
the drag coefficient and drag load distribution representations. Beyond the simple
representation given earlier, let the drag coefficient be additionally represented
by an nth-degree truncated power series variation in angle of attack. Moreover,
allowing for negative angles of attack and a desire to maintain analyticity, we
restrict our approximation to only even powers of angle of attack:
cd = cd0 + n
cdn αQS (10.47)
n=2k
The utility for using perturbational airloads is based on the premise that per-
turbations of the three components (i.e., lift, drag, and pitching moment) can be
expressed as perturbations in the four elemental kinematic descriptors of the air-
foil’s dynamic state, as given in Eqs. (10.46), namely, δUT , δUP , δθ , and δ (dθ/dt).
For any given aeroelastic problem, these basic perturbational descriptors can then
dFz dL
δ ≈δ
dr dr
1
= ρac(sgn UT0 )UT0 UT0 δθ + δUP + [2θ + (sgn UT0 )φ]δUT
2
(10.48)
dFy 1 cd
n−1
δ = ρac sgn(UT0 )UT0 UT0 sgn(UT0 )φ − n n αQS δθ
dr 2 a
n=2k
cd
n n−1
+ θ + 2 sgn(UT0 )φ − n αQS δUP
a
n=2k
cd cd
n n−1
+ φθ − 2 + φ n αQS δUT (10.49)
a a
n=2k
Pitching-moment distribution:
dMxSC 1
δ = ρcmc/4 c sgn(UT0 )UT0 δUT
dr 2
c dFz
+ yAC − [1 − sgn(UT0 )] δ
2 dr
π 3
− ρc θ̇0 δUT + sgn(UT0 )UT0 δ θ̇ (10.50)
8
which equals the mass moment of inertia about the teeter hinge and where km1
and km2 are mass radii of inertia of the blade section about the major and minor
2
principal axes, respectively. (Thus, for most “thin” sections, mkm mkm2 .)
2 1
The aerodynamic related integrals are then defined as follows:
γ
Cβ̇ = ; [γ ≡ blade Lock number = ρacR4 /Ib ] (10.55a)
8
γ
Cθ = − (10.55b)
8
γ c 1 yA C 1
Cθ̇ = − − − β0 θ0 (10.55c)
2 R 4 3c 2
!
γ 1 2 cd " c yA C
Dβ̇ = − β 0 θ0 − λ 1 + 0 − (10.55d)
2 4 3 a R 3c
γ cd0 1 1
− + λθ0 − λ2 ; rotor type 1
2 4a 3 2
Dβ = (10.55e)
0; rotor type 2
! "
γ1 c 2 1 yA C 1 ya
Dθ̇ = − −
22 R 4 3c 2 c
c 2 yA C 3 cd0 1
+ β0 θ0 − λ + β02 + θ0 λ (10.55f)
R 3 c 8 4a 3
γ 2 c yA C
Dθ = β0 θ0 − β0 λ − (10.55g)
2 3 R 3c
Divergence. The conditions for divergence stability are that the constant term
in the characteristic equation be positive. Thus, the divergence stability bound-
ary is given by setting that term to zero. This is accomplished by the following
determinantal equation:
# #
# 1 (E3 + Cθ ) #
# #=0 (10.57)
#(E + D ) [(E2 − E1 ) − β0 + Dθ + K̄θ ]#
2
3 β
or
Application of this inequality leads to a quadratic equation for the required control
stiffness K̄θ . Solution of this equation gives this stiffness as a function of the
remaining parameters.
Fig. 10.18 Perspective view of blade defining the in-plane and out-of-plane degrees
of freedom.
Fig. 10.19 Section view of the blade deflection components at an arbitrary radial
station.
αg = Lζ ζH + Lβ βH (10.64)
6) Because the blades sustain root bending moments, the blade bending deflec-
tions ve and we are defined by the blade (cantilevered) normal modes in the flatwise
and bending directions, as developed in Chapter 3 and as coupled because of the
pitch angle θ0 , using Eqs. (3.145) and (3.146).
Lead-lag:
S̄1 K ζH S̄12
p̃ + cζ ωζ p̃ + ē0 +
2
2
ζH − 2β0 p̃βH − sin θ0 p̃2 qw
S̄2 S2 S̄2
S̄46 S̄12 S̄46 S̄16
+ cos θ0 p̃ qv − 2β0
2
cos θ0 p̃qw + sin θ0 p̃qv − ē0 sin θ0 qw
S̄2 S̄2 S̄2 S̄2
R
S̄48 1 dFy γ 1 dF̄y
+ ē0 cos θ0 qv = r δ dr = x δ dx
S̄2 S2 2 e0 dr 2 ē0 dx
(10.66)
Blade bending:
Flatwise bending:
Edgewise bending:
where
d( ) 1 d( )
p̃ ≡ = (∗ ) = ; γ ≡ (Lock number) = ρacR4 /Ib (10.69)
dψ dt
and where cβ , cζ , cw , and cv represent structural damping coefficients for flapping,
lead-lag, flatwise bending, and edgewise bending, respectively.
The perturbational airloads are nondimensionalized according to
dF 1 dF̄
δ = ρac (R)2 δ (10.70)
dr 2 dx
and the matrix relationship between the basic perturbational descriptors and the
system variable perturbations is given by
p̃βH
δ ŪT 0 x −γw sin θ0 γv cos θ0
p̃ζH
δ ŪP = −x 0 −γw cos θ0 −γv sin θ0
0 0 0 0 p̃qw
δθ p̃qv
βH
0 0 00 0
ζH
+ −µ cos ψ 0 −γw µ cos ψ cos θ0 −γv µ cos ψ sin θ0
Lβ Lζ 0 0 qw
qv
(10.71)
The use of such an implicit formulation allows for a more rigorous modeling
of the inflow. For purposes of generating useful results defining the stability char-
acteristics, the blade element momentum theory is used. This theory is standard
in rotor aerodynamic texts (e.g., Leishman) and treats the inflow by considering
the momentum and mass conservation laws on annuli of the rotor disk. This for-
mulation produces an inflow model with spanwise variation with a completely
arbitrary spanwise distribution of pitch angle. For the case of zero twist, the inflow
distribution is given by
−1 σ clα 32
λ(x) = UP = 1+ θ0 x − 1 (10.72)
R 0 16 σ clα
The resulting aerodynamic terms in the equations of motion result from spanwise
integrations and are thus implicit; they do reflect a more comprehensive modeling
of the perturbational airloads, however. This type of modeling with the attendant
spanwise integrations is actually easily implemented in a computer code.
The appropriate nondimensional natural frequencies for the blade in bending
are taken to be those for β0 = θ0 = 0:
ωw
ω̄w = ≡ first flatwise natural frequency (10.73a)
ωv
ω̄v = ≡ first edgewise natural frequency (10.73b)
cl = cα α = 0.11α
cd = cd0 + cd2 α + cd4 α = 0.006 + (6.4421 × 10−5 ) α 2 + (9.4142 × 10−8 ) α 4
2 4
where the angle of attack α is in degrees. Also, the blade coning is calculated using
simple blade element theory:
γ 1 dF̄z [x, θ0 , λ(x, θ0 )]
β0 = x dx (10.74)
2ω̄β2 ē0 dx
S̄1 KβH
ω̄β2 = 1 + ē0 + (10.75)
S̄2 S2 2
and, as defined in Appendix D, S1 and S2 are, respectively, the blade first and
second mass moments. For a uniform mass distribution m0 , Eq. (10.75) can be
simplified to
3 ē0 3KβH
ω̄β2 = 1 + + (10.76)
2 (1 − ē0 ) m0 R3 (1 − ē0 )3 2
For completeness the nondimensional lead-lag frequency is given by
Note that the barred (i.e., nondimensional) forms of all of the Si coefficients are
defined in Appendix D.
Figures 10.20a and 10.20b and 10.21 present the principal stability characteris-
tics for a nondimensionalized rotor configuration. Figure 10.20a is representative
of a completely articulated rotor with zero hub spring rates. Figure 10.20b shows
the similar characteristics as Fig. 10.20a, but for a case wherein the hub spring
rates KβH and KζH are finite, thereby raising the blade natural frequencies.
Fig. 10.21 Effect of Lock number on the pitch-lag stability boundaries (soft in-plane
hingeless rotor).
by simple variations of the lead-lag hub spring rate KζH . For hingeless rotors it
would be expected that there is generally less movement of the pitch horn/push
rod with lead-lag motion than what would be available to the fully articulated
rotor. Also for a hingeless rotor there is generally no lead-lag damper, and so the
damping parameter Cζ must be taken to be zero.
The figure clearly shows that the character of the stability essentially inverts
at the point wherein the lead-lag frequency is equal to the flapping frequency.
Fig. 10.22 Variation of pitch-lag stability boundaries with lead-lag frequency and
pitch angle—simplified hingeless rotor.
Consistent with the results presented in Figs. 20a and 20b, the effect of increased
collective angle, with the corresponding increase in coning angle, is destabilizing.
Lead-lag system:
# ##
# S̄46 2 S̄48
# (ω̄ζ2 − ω̄2 ) − ω̄ + ē0 #
# #
# S̄2 S̄2 #
# #=0 (10.83)
# #
# − S̄46 ω̄2 + ē0 S̄48 S̄49 2
(ω̄v − ω̄2 ) #
# S̄ S̄ S̄2 #
2 2
where ωβ and ωζ are given by Eqs. (10.76) and (10.75), respectively.We first
rearrange the characteristic equation resulting from Eq. (10.82) to isolate the hub
flapping spring KβH :
Or, after some rearranging, invoking Eq. (10.81a), and relating the hub spring to
the blade spring using the coupling factor, we obtain an equation for the equivalent
blade spring KβB :
(1 − Rβ ) S̄1 [S12 (1 − ω̄2 ) + e0 S16 ]2
KβB − S2 ω̄ − 1 − ē0 −
2 2
=0
Rβ S̄2 S2 [−S10 ω̄2 + KCFw + R2 KβB ]
(10.85)
where
1 dγ 2
w
KCFw = R T (x) dx (10.86)
ē0 dr
and the coupled frequency ω for this case is now the effective coupled lead-lag
frequency (ωζ 0 )C .
Hover condition. With the work of Ormiston and Hodges as a guide, varia-
tions can be made with both the effective coupled flapping and lead-lag frequencies,
as developed in Eqs. (10.83) and (10.85), to establish mappings of the resulting
stability boundaries. The selected method of modeling of the rotor blade (i.e., with
elastic bending modes supported by discreet springs) is intended to model only the
resulting dynamic responses that are roughly in phase. Mathematically, as part of
the eigenvalue solution out-of-phase modes are also predicted. These modes are
found to be of very high frequency and can be considered to be “phantom” modes
and not realistic. To suppress any unstable conditions for these modes, a small
amount of equivalent structural damping is introduced. In the results to follow, all
four of the degrees of freedom were given values of damping amounting to 0.1%
of critical (i.e., cβ = cζ = cw = cv = 0.002).
Figures 10.24 and 10.25 present stability boundaries as a function of blade
coupled frequencies and blade pitch angle, for the hingeless rotor as calculated
with the fully elastic formulations. Figure 10.24 presents results for low values of
the structural coupling factors (Rβ and Rζ = 0.02), commensurate with a relatively
stiff outer blade and soft hub attachment. On the other hand, the results of Fig. 10.25
are for a configuration of moderate structural coupling (Rβ and Rζ = 0.5).
Fig. 10.24 Variations of flap-lag stability boundaries with coupled frequency and
pitch angle for a hingeless rotor blade with low structural coupling.
Fig. 10.25 Variations of flap-lag stability boundaries with coupled frequency and
pitch angle for a hingeless rotor blade with moderate structural coupling.
The trends shown in these figures are consistent with the findings of Ormiston
and Hodges:
1) The instability is limited mostly to subcritical rotors (i.e., those with in-plane
frequencies greater than rotor speed).
2) The instability is most intense for those conditions wherein the in-plane
frequency is close to that for the flapping motion.
3) The effect of increased structural coupling is to translate the instability
boundaries to higher values of in-plane frequency.
Fig. 10.26 Effects of trim conditions and inflow modeling on blade flap-lag stabil-
ity in forward flight: ωβ = 1.15, ωζ = 1.4, R = 0, γ = 5, σ = 0.05, cd0 = 0.01, and
CT /σ = 0.2 (after Peters).
Fig. 10.27 Force balance for a forward whirling or whipping shaft (after Ehrich).
Clearly, two rotations are in evidence, the rotor rotational rate , as well as a
whirling frequency ±ω, the sign depending on the direction of whirl. Most of the
forces acting on the rotor are either in phase or out of phase with the deflection
or act against the radial velocity of the motion, dr/dt (i.e., radial damping) or
at 90 deg to the radial displacement, as defined by Coriolis forces and available
damping of the whirling motion. Only the generation of a force in phase with the
whirling motion and proportional to the deflection of the motion Fθ , as shown
in the figure, can lead to instability. Note that this is a graphical interpretation of
the force-phasing matrix idea discussed in Chapter 9. A complete discussion of
all of the types of instability for rotary shafts is beyond the intended scope of the
present text. The interested reader is referred to the work of Ehrich, from which
the present material is excerpted. The intent herein is only to identify the various
destabilizing effects and the physics involved in generating forces that are in phase
with the direction of whirl.
This solution, however, would limit the use of the lighter-weight designs possible
with supercritical operation. Another approach is to minimize the number of shaft
elements connected by couplings that provide damping in their interfaces.
krθ
w = (10.89)
cr
where the cross-coupling stiffness and radial damping are dependent on the eccen-
tricity ratio ε0 , which measures where in the journal housing the journal is
operating. A zero value denotes a centered position, whereas a unit value denotes
contact of the journal with the wall. Data on the variations of stiffness and damping
characteristics with eccentricity ratio are usually provided by the bearing manu-
facturers. For typical plain journal bearings the whirl ratio typically comes out to
approximately one-half. The whirl ratio is then used to approximate the onset of
instability (OSI):
ωn
OSI = (10.90)
w
where ωn is the rotor shaft critical speed. Thus, instability would be expected
to occur at approximately twice the lowest shaft critical speed. As discussed in
detail by Ehrich, a number of devices have been developed to alleviate this type
of instability.
a contact force opposite to the shaft’s rotation, and hence, in the direction of
the whirling motion. Furthermore, the coulomb friction force can be expected to
increase with increased deflection. Thus, both conditions are met for establishing
the existence of an instability. The obvious remedy for this type of instability is to
provide lubrication. If that is not feasible, abradability of either of the two elements
should be provided. Also, whirl of the shaft with its housing can be controlled by
maintaining the natural frequencies of the two elements to as dissimilar as possible.
References
Section
10.1 Brosens, P. J., and Crandall, S. H., “Whirling of Unsymmetri-
cal Rotors,” Journal of Applied Mechanics, Vol. 83, Sept. 1961,
pp. 355–362.
Crandall, S. H., and Brosens, P. J., “On the Stability of a Rotor with
Rotationally Unsymmetric Inertia and Stiffness Properties,” Journal
of Applied Mechanics, Vol. 83, Dec. 1961, pp. 567–570.
Gladwell, G. M., and Stammers, C. W., “On the Stability of an
Unsymmetrical Rigid Rotor Supported in Unsymmetrical Bearings,”
Journal of Sound and Vibration, Vol. 3, No. 3, 1966, pp. 221–232.
Wilkerson, J. B., “An Analytical Study of the Mechanical Stability
of Two-Bladed Rigid Rotor Systems,” Naval Ship Research and
Development Center Report 3351, Jan. 1970.
10.2 Dowell, E. H., Curtiss, H. C., Scanlan, R. H., and Sisto, F., A Mod-
ern Course in Aeroelasticity, Sijthoff & Noordhoff, Rockville, MD,
1978.
Peters, D. A., “Flap-Lag Stability of Helicopter Rotor Blades in For-
ward Flight,” Journal of the American Helicopter Society, Vol. 20,
No. 4, 1975, pp. 2–13.
10.4 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, 1972, pp. 2–14.
Peters, D. A., “Flap-Lag Stability of Helicopter Rotor Blades in For-
ward Flight,” Journal of the American Helicopter Society, Vol. 20,
No. 4, 1975, pp. 2–13.
10.5 Childs, D. W., “Identification and Avoidance of Instabilities in High-
Performance Turbomachinery,” Von Karman Institute for Fluid
Dynamics, lecture, Rhode-Saint-Genese, Belgium, 1987.
Ehrich, F. F., Handbook of Rotordynamics, McGraw–Hill, New York,
1992.
Ehrich, F. F., “Self-Excited Vibration,” Shock and Vibration Handbook,
Third Edition, edited by C. M. Harris, McGraw–Hill, New York,
1988, Chap. 5.
Problems
10.1 For the unsymmetrical rotor instability phenomenon consider the com-
pletely isotropic case ( i = s = 0), but with both internal and external
viscous damping (both isotropic) ci and ce , respectively. Denote nondi-
mensional forms of these damping sources as
the axial inertia term A and the nondimensional damping values σi and σe .
(c) Relate the result of part b to Bishop’s findings for stable operation
of supercritical shafting.
and, second, that a substantial drag rise is incurred. Using the ratio of the
drag-divergence Mach number to the advancing blade-tip Mach number
(= Mdd /MR ), formulate a Fourier-series representation for the value of
the lift-curve slope for radii that experience Mach numbers greater than or
equal to the drag-divergence value.
10.3 Derive an expression for the (periodic) flap damping of an articulated rotor
blade (i.e., δUp = −xRδ β̇) taking into account reversed-flow effects. [Note:
Group the linear aerodynamic descriptor constants together into a single
aerodynamic effectivity constant Aa (= ρ/2 acR3 R).]
10.4 Derive the inertial loading terms in the equations of motion for rotor
weaving using a Newtonian approach and the theorem of Coriolis.
Derive the additional terms in the equations of motion for rotor weaving to
account for pitch-flap coupling.
Hint: The basic equation set can be represented in abbreviated matrix
form as
β Lβ
[L( p̃)] = = {0} (10.95)
θ Lθ
10.6 Show that for equal values of the structural coupling factors Rβ and Rζ ,
they are also equal to the structural coupling factor given by Peters:
1/KβB − 1/KζB
R≡ (10.99)
[1/KβB + 1/KβH ] − [1/KζB + 1/KζH ]
387
flapping and flatwise bending the collective rotor modes represent a measurement
of the out-of-plane displacement of the rotor center of gravity from the undeflected
rotor plane. Likewise the in-plane lead-lag and edgewise bending cyclic rotor mode
descriptions provide a corresponding measurement of the in-plane position of the
rotor center of gravity position. Similarly, the cyclic flatwise bending rotor modes
define the pitching and rolling rotational motion of the rotor, and the collective
edgewise bending rotor modes define the yawing rotational motion of the rotor
from which the effective torsional inertia effects can be defined.
The importance of the collective and cyclic components of the rotor mode
description arises from the fact that it is only these modes that elastomechanically
couple with the pylon. A complete rotor mode description of the rotor could include
a Fourier-series expansion that would contain second (and higher) harmonic com-
ponents. However, these portions of the expansion involve neither translations
nor rotations of the rotor center of gravity and, hence, do not produce hub loads
(either forces or moments); these rotor modes are essentially “reactionless.” For
this reason further consideration of them will be omitted herein.
that the propeller-nacelle whirl flutter phenomenon was the most probable cause
of the loss of the aircraft. Retrofit stiffenings were made to the remaining aircraft
of this type, based on knowledge of the stabilizing trends for this phenomenon,
and thereafter no more aircraft were lost.
The study of this phenomenon forms a convenient transition from the material
of the preceding chapter to that of the present one. The basic characteristics of
the more generally defined rotor-nacelle whirl flutter are easily defined using a
relatively simple dynamic representation; at the heart of the formulation is the
dynamics of a relatively simple gyroscopic element. As such, the theory builds not
only on the gyroscopic theory tools covered in an earlier chapter but also on the
material relating to the unsymmetrical rotor of the preceding chapter.
In θ̈ + Cθ θ̇ + Kθ θ − Ix ψ̇ = Mθ (11.3)
Ix θ̇ + In ψ̈ + Cψ ψ̇ − Kψ ψ = Mψ (11.4)
where Mθ and Mψ are, respectively, the applied moments in pitch and yaw, taken
about the pivot point, arising from the aerodynamic loads.
UT = r + ṡ + V sin α1 (11.5)
UP = ẇ + V cos α1 (11.6)
where
Fig. 11.3 Front view geometry of deflected hub and radially located typical section.
Fig. 11.4 Air velocity components and airload components at the typical blade
section.
The square of the effective resulting velocity (that producing dynamic pressure) is
then approximated by
Ue2 = (2 r 2 + V 2 ) + 2rṡ + 2V ẇ + 2rV α1 + (higher-order terms) (11.10)
U2
The lift at the airfoil section can then be expressed in terms of the effective
dynamic pressure and the angle of attack of the section:
dL(r, t) 1
= = ρUe2 · [clα α] ·
c
dr 2
cl chord
dynamic
pressure
1
= ρclα c (U 2 + 2V ẇ + 2rṡ + 2rV α1 ) · (β − γ ) (11.11)
2
where
β − γ = β − tan−1 (UP /UT )
r V V2
= β − tan−1 (V / r) − 2 ẇ + 2 ṡ + 2 α1 (11.12)
U U U
≡ α0
After the various formulations are combined, the resulting expression can be
expanded wherein only the linear terms are retained. The following expression
for the lifting load distribution is then obtained:
dL(r, t) 1 r 2V
= ρclα cU α0 − 2 1 −
2
α0 ẇ
dr 2 U r
V 2r V2 2r
+ 2 1+ α0 ṡ + 2 1 + α0 α1 (11.13)
U V U V
b ψ̇ θ̇
Ly = Ka A1 ψ − aA1 + A2 (11.15)
2
b θ̇ ψ̇
Lz = Ka A1 θ − aA1 − A2 (11.16)
2
b ψ̇ θ̇
My = Ka R A2 ψ − aA2 + A3 (11.17)
2
b θ̇ ψ̇
Mz = Ka R A2 θ − aA2 − A3 (11.18)
2
where b is the number of blades, and
Ka = 21 ρclα R4 2 (11.19)
The various integrals given in the preceding expressions are defined as follows:
1 c (J/π )2
A1 = dη; A1 = (J/π )A1 (11.20)
0 R (J/π )2 + η2
1 c (J/π )η2
A2 = dη; A2 = (J/π )A2 (11.21)
0 R (J/π )2 + η2
1 c η4
A3 = dη; A3 = (J/π )A3 (11.22)
0 R (J/π )2 + η2
where J = πV / R ≡ the propeller advance ratio.
The moments about the pivot are then given by
Mθ = −My + aRLz (11.23)
Mψ = Mz + aRLy (11.24)
b θ̇
Mθ = Ka R −(A3 + a A1 ) − A2 ψ + aA1 θ
2
(11.25)
2
b ψ̇
Mψ = Ka R −(A3 + a2 A1 ) + A2 θ + aA1 ψ (11.26)
2
Fig. 11.5 Summary of key results of rotor whirl trend studies of propellers with rigid
blades (after Reed).
For conditions of backward whirl, θ and ψ̇ are in phase with each other, and
ψ and θ̇ are out of phase. For these conditions the expressions just given for
aerodynamic moment show that the second terms (−)A2 ψ and (+)A2 θ in each of
these expressions are out of phase with their respective rate-dependent terms and
therefore act as negative dampers. These terms are then the source of the flutter
instability.
The following detailed interpretations can be drawn from the trends depicted
in Fig. 11.5:
1) The extent of the susceptibility of the rotor to the instability is directly related
to the isotropy of the pylon structure. The more nearly isotropic the structure,
the more likely the characteristic response of the hub will be a circular whirl.
Increasing amounts of anisotropy serve to distort this response into an elliptic form
with commensurately less ability of the motion to generate the already identified
destabilizing driver forces.
2) For small amounts of damping, the effect of incremental damping is sig-
nificantly greater than that for moderate amounts. Thus, it is quite important to
provide the pylon structure with at least a modicum of damping. However, beyond
a certain point, incremental damping does not offer a viable means for stabilizing
the rotor.
3) The stabilizing effect of the pivot to hub distance aR is attributed to the
additional aerodynamic damping afforded by the motion of the rotor in its plane
(Ly and Lz caused, respectively, by ẏ and ż motion).
soft flapwise stiffness rotor case were successfully predicted in theory by Young
and Lytwyn. Their formulation uses a straightforward Lagrangian procedure and
includes a fairly standard application of (incompressible) quasi-steady lift and drag
airloads. The reader is referred to the literature for a detailed description of the
equations of motion used to model the instability. Alternatively, the generalized
equations of motion for ground and air resonance given in Appendix D can be
used as a basis for defining an approximate formulation to the problem. The key
findings of the work of Young and Lytwyn are summarized in Figs. 11.7 and 11.8,
which were extracted from their work.
Figure 11.7 shows that for blade flapping frequency ratios above 1.1 the effect of
flapping flexibility is to increase the amount of nacelle stiffness required to stabilize
the usual (backward-) whirl mode of instability. Figure 11.8 shows the presence
of the forward-whirl instability for a range of flapping frequency ratios less than
or equal to approximately 1.07. The figure also indicates that for frequency ratios
in this range the prop rotor can experience instability in both the forward- and
backward-whirl modes.
The analysis of actual real-world prop rotors of the tilt-rotor variety must nec-
essarily require a more sophisticated analysis than that provided by either the work
of Young and Lytwyn or a modification of Appendix D. Not only are the (omitted)
effects of unsteady aerodynamics significant to the stability boundaries, but the
details of specific couplings (peculiar to any given prop-rotor configuration) can
greatly influence or even stabilize an existing instability (see Wernicke and Gaffey).
Fig. 11.7 Influence of blade flapping frequency ratios greater than 1.1 on the
(backward-whirl) stability of propeller rotors (after Young and Lytwyn).
Fig. 11.8 Influence of blade frequency ratios from 0.90 to 1.12 on the (forward-whirl)
stability of propeller rotors (after Young and Lytwyn).
combination of two factors to the incompressible lift-curve slope clα (F1 and F2 ,
(=c · F · F ), where
respectively) to define a “corrected” lift-curve slope clα lα 1 2
(AR)
F1 = 1/ 1 − M 2 (r); F2 = (11.27a)
(AR) + 2
and where
M(r) ≡ Mach number at station r
(AR) ≡ Mach number corrected aspect ratio of the blade section,
(AR)
F1 · F2 = (11.28)
2 + (AR) 1 − M∞
2 [1 + (π 2 /J 2 )η2 ]
where M∞ is the forward-flight Mach number and η = r/R. This combined factor
is then inserted under the integral in the expressions given earlier for the coefficients
A1 , . . . , A3 , so that it forms part of each of the integrands.
When available, wind-tunnel measured propeller derivatives should be used to
check the preceding results. Furthermore, in addition to the approximations just
given to the compressibility and finite-span effects, additional modifications can
be added to the analysis to approximate unsteady aerodynamic effects. Methods
for the inclusion of these effects are covered in a subsequent chapter and therefore
will not be treated in any further detail at this point.
Fig. 11.9 Schematic presentation of the modeling detail used in the DYN4 analysis
(Popelka et al.).
Fig. 11.10 Schematic of the basic dynamic subsystems defining the rotor-nacelle
aeroelastic problem.
−−
where [G2 ] = [H13 ][H33 ]−1 [H31 ] and is composed of an elastomechanical part
and a part caused by the unsteady airloads:
H11 | H12
[Ĥ(s)][Ĥ(s)] = − − − − − = [Ĥ(s)]elasto−mechan. + [Ĥ(s)]aero. (11.35)
H21 | H22
where
[G2 (ω)] = [H13 (ω)][H33 (ω)]−1 [H31 (ω)] (11.39)
It should be stressed that each of the three component matrices in Eq. (11.38)
is obtainable from separate sources. Also, the matrices comprising Eq. (11.39)
(which are obtained from the selected rotor aeroelastic analysis) can be alterna-
tively obtained by using a time-history solution of the computer code to establish
responses that have converged to a purely sinusoidal state. This is especially useful
for treating nonlinearities in the rotor equations. Finally, Eq. (11.38) is now in the
proper form for application of the multiple-degree-of-freedom frequency-domain
methodology described in Chapter 9.
Fig. 11.12 Simplified dynamic description of fuselage for ground resonance analysis.
The dynamic description of the flexible rotor blades is achieved using cyclic rotor
modes as described in an earlier section (see Fig. 11.1).
without derivation. Note that the first two equations define the force equilibrium of
the nonrotating hub, and the second two define the force equilibrium of the cyclic
in-plane rotor modes.
Hub longitudinal force Fx :
where the effective masses in the x and y directions, Mx and My , are respectively
given by
Stability solution. The solution for the stability descriptors is standard and
makes use of the solution of either an expanded characteristic polynomial (in this
case an eighth-degree polynomial) or a matrix eigenvalue solution. The roots λj
will have the usual form
λj = σj + iωj (11.46)
λj = ± σj ± iωj (11.47)
For this latter case the eigensolution clearly indicates an instability. These results
are conveniently shown on a Coleman diagram shown in Fig. 11.13; this diagram
shows the variation with rotor rotational speed of the resulting coupled frequencies
(both blade and fuselage modes) as seen in the nonrotating coordinate system.
Note that the regions of instability always occur at the intersections of the degen-
erate “regressive” lead-lag mode with one or more of the fuselage modes. For these
conditions the regressive lead-lag mode is denoted as “degenerate” because the
rotor has become supercritical; the blade frequency has become less than the rotor
rotational frequency. At these conditions the normally regressive mode has changed
direction and become a slower moving progressive mode. More specifically, the
following characteristics can be noted from this figure:
1) The basic structure of this diagram is similar to what the various modes would
look like with zero coupling (i.e., if the ratios of S48 to each of the hub masses,
Mx and My → 0). For zero coupling the intersections between the blade modes
and the pylon modes would show no deviations. However, with finite coupling the
various intersections become coalescences. For coalescences that occur above the
1/rev (lP) line (ω = ), the coupling is smooth, and only pure frequencies result.
Only for coalescences that occur below the 1/rev line can the coupling result in
instabilities.
2) The degree of inertia coupling as defined in terms of the earlier identified
terms is a major measure of the susceptibility of the system to the instability.
3) The shaft critical speeds are indicated as the points where the 1/rev line
intersects the pylon frequencies.
4) At the point where the coupled blade mode crosses the horizontal axis, that
frequency has a value of zero and represents a condition wherein a steady force
can excite the system. Note that this condition is also the point where the rotating
blade mode frequency is exactly equal to the rotor rotational frequency.
5) The figure implies one effective means of avoiding the instability: Do
not operate the rotor near the instability regions that are closely approxi-
mated by conditions wherein [the blade lead-lag (rotating) frequency] + [one of
the pylon mode frequencies] = [the rotor rotational frequency].
Isotropic case. Here we set the hub and rotor center of gravity displacements
to complex forms:
For the product of damping to function properly, the blade and fuselage damping
must be expressed in a similar fashion:
Next, let us assume a neutral stability point right on the stability boundary:
λ = iω, where ω̂ is the frequency on the Coleman plot taken at the center of
instability point, = ω̂ = ωx = ωy = − ω . (Note that here ω is an alternate
expression of the edgewise natural frequency ωv .) With these simplifying assump-
tions the characteristic equation is expressible as a real part and an imaginary part,
both equal to zero. Expressing the damping in the pylon to be some isotropic value
c̃ then permits the real part of the characteristic equation (after some algebra) to
be written in the following compact form:
˜ 3 ω̂3
c̃c = (11.53)
ω
where
My = (mFy + mR ) (11.56a)
Also,
Using similar algebraic manipulation, one can then show that for the infinitely
anisotropic case,
3
3v ωy
c̃y c ≥ (11.57)
2ω
which is the product of damping criterion for the infinitely anisotropic case, as
written here for roll flexibility. The result is the same for pitch flexibility only
infinitely anisotropic case, except that ( )y → ( )x . Note that this result is identical
to the isotropic case except for the factor of two in the denominator.
= /ωy (11.59a)
λζ = cζ /Iζ ωζ ; cζ = blade lag damper rate (11.59b)
λy = cy /My ωy ; cy = equivalent pylon damper rate (11.59c)
in the y direction
1 = ωζ2 / 2 ; ωζ = blade lead-lag frequency (11.59d)
(=ω = ωv )
Note that in Fig. 11.15 the values of hub (pylon) damping required for stability, as
determined by the product of damping criterion, are included for comparison. As
developed therein for rigid lead-lagging blades and an infinitely anisotropic pylon,
the product of damping criterion is expressed as
1 1
λy λζ ≥ 3 √ −1 (11.60)
2 1
As shown in Fig. 11.15, the inadequacy of the criterion is evident at both the
high and low values of blade-blade damping wherein the hub and blade damping
values are most dissimilar. The agreement is also quite apparent, however, when
the damping values are approximately equal.
The usefulness of the Gabel and Capurso results (as well as other such graphic
results) lies in their ability to provide quick estimates of the ground-resonance
characteristics for each pylon mode separately, at least for cases wherein the modes
are reasonably well separated in frequency. By combining the results obtained
separately for each pylon mode, the combined Coleman plot (such as that depicted
in Fig. 11.13) can be approximated.
M = Fdamper l cos γ
Fdamper = clinear dżdamper = clinear l cos γ dθ̇ (11.62)
M = clinear l2 cos2 γ dθ̇ = (cos2 γ )Ef cref dθ̇
Thus, the skewing degradation factor κN is equal to cos2 γ , where γ is equal to half
the angle between the successive blades 2π/N. The results of the two effectiveness
factors are shown in Table 11.1.
The results shown in Table 11.1 would indicate that there is no payoff for
using a blade-to-blade damper configuration. Yet, as shown in Fig. 11.17, such a
configuration has been incorporated in the hub design of a (four-bladed) production
helicopter.
Although there is no payoff in terms of increased effectiveness, this configura-
tion is practical for two reasons: first, a longer moment arm, l, can be used while
still keeping the hub relatively compact, and second, the design results in a simpler
and, therefore, lighter hub. Furthermore, because the dampers are attached further
outboard they will be easier to inspect and service.
N 3 4 5 6
Ef 3 2 1.38 1
κN 0.25 0.5 0.66 0.75
κN ·Ef 0.75 1 0.90 0.75
Fuselage:
Mx = 131.479 lb-s2 /ft, My = 28.687 lb-s2 /ft,
ζx = 0.05, ζy = 0.05,
ωx = 35.0 rad/s, ωy = 82.1 rad/s
Rotor:
R = 28 ft, ref = 221.68 rpm,
b = 4, ζv = 0.005,
Mb = 3.8801 lb-s2 /ft, (ωv )ref = 26.979 rad/s,
S48 = 60.214 lb-s2 , Kv 0.256,
S49 = 1144.0 lb-s2 2-ft
Operational condition:
= 68.068 rad/s (R = 650 ft/s)
Ground-resonance eigensolution results:
λGR = 1.31428 + i31.8026
eigenvector:
φGR = 1 + i0, 0.1712 − i0.1061, 0.0900 + i0.2289, 0.2275 − i0.0954
is driving the fuselage mass by virtue of the centrifugal force arising from the
whirl motion.
Figure 11.19 presents the results of the force-phasing matrix calculation made
using the eigenvector information along with the mass, damping, and stiffness
matrices for the system. This figure shows that the only energy flow path and critical
drivers are those that result from the cyclic in-plane rotor modes. The fuselage
degrees of freedom do not directly take part in the energy flow path, but are instead
driven by the rotor modes through the mass matrix coupling terms m13 and m24 .
The terms defining the energy flow path are blade inertia and damping related that
arise from the rotor mode transformation defined by Eq. (11.2). Also, although not
shown, the diagonal stiffness terms for the cyclic in-plane rotor modes are negative,
commensurate with the rotor being supercritical. The role that fuselage dynamics
play in the physics is thus seen to be that of an “enabler” in that the fuselage provides
a low impedance support for the rotor center-of-gravity’s motion at the coincident
frequency of − ωe (=ωFx ). These results are not particularly significant as far as
defining any new methods for stabilization, rather they underscore well-established
knowledge gleaned pragmatically from looking at trends of stability eigenvalues.
The results also underscore the basic well-established method for stabilization:
avoid any frequency coalescence between the rotor and fuselage and increase the
impedance of the fuselage (by adding damping).
H11 (s) | H12 (s)
XF
−−−−−−− = {0} (11.63)
XR
H21 (s) | H22 (s)
where X F represents the state variables of the fuselage plus the rotor as a point
mass and X R represents the variables defining the blade lead-lag rotor modes. The
partitions of Eq. (11.63) can be rewritten as two matrix equations, which then
(ω) {XF } = [H11 (ω)]−1 [H12 (ω)] [H22 (ω)]−1 [H21 (ω)] {XF } (11.66)
where
Note that the methodology works equally well if the lead-lag vector X R is used
in Eq. (11.66) and when
systems the choice of amplitude for the sinusoidal test forces would be of some
concern. It would be prudent to keep the amplitude small enough to maintain
linearity and yet large enough to keep the results out of the “round-off noise.”
The issue of the presence of nonlinearities in the system is implicitly addressed
in this methodology, in that the fuselage mobilities thus calculated are linear-
effective mobilities. (Note that these would consequently vary with response
amplitude for nonlinear systems). Although this procedure forms a “neat” way
of addressing the nonlinearities, it needs to be stated that this form of linearization
is no less valid than (and is actually equivalent to) the use of describing func-
tions, which are routinely used as a standard basis for linearization, as described
by Savant. Furthermore, as a result the new methodology easily enables the pre-
diction of limit-cycle conditions resulting from nonlinearities. The analysis of
limit-cycle conditions requires that suitable variation of the sinusoidal test loads
be made. Lastly, although not a nonlinearity per se, the condition of a failed or par-
tially failed lead-lag damper (the doubly anisotropic problem) is also addressable
by this methodology without recourse to Floquet theory (Bielawa).
ωM = − ωe (11.70)
where ωe is the blade lead-lag natural frequency (at rotor speed). It is then a
straightforward programming exercise to evaluate the phases (or alternatively, the
imaginary parts) of the characteristic multipliers to identify the real-axis crossing
value, as is described in Chapter 9.
the hub separately in both the fore and aft and lateral directions, and, for each
direction of excitation, responses should be measured in both the fore and aft and
lateral directions. The rotor impedance matrix [G2 (ω)] on the other hand can be
calculated with considerable accuracy [using well-established ground resonance
equations and Eq. (11.68b)]. With the two matrices available, the characteristic
multiplier calculation procedure can then be made to indicate the presence of
instability, as well as a measure of the stability level using the techniques given in
Chapter 9.
yR -direction force:
−2S48 ẋR + S48 ( ÿR − 2 yR ) + S49 (¨ + 2ζv ωv0 ˙ + ωv2 ) = 0 (11.73)
Fig. 11.24 Typical Coleman diagram for ground resonance of two-bladed anisotropic
rotor and isotropic fuselage.
shaft critical speeds are the actual intersection points with the horizontal axis
as shown. However, the rotor speed range between these two shaft critical speeds
defines (in the rotating coordinate system) an additional aperiodic instability range
characteristic of two-bladed rotors, not present with the isotropic rotors. These
two shaft critical speeds are seen in the rotating coordinate system as a divergence.
Therefore, the values of shaft critical speed are then obtained by omitting the time
derivative and damping terms in the characteristic equation and solving for the two
rotor speeds C1 and C2 :
C1 = ω (11.75)
This critical frequency is seen to exclude any of the blade bending descriptors.
It is, in fact, the shaft critical speed that results when the shaft is deflecting in a
direction parallel to the blade axis (with zero blade bending). The second shaft
critical speed can be obtained in a compact form by making use of the simplified
relationship for the blade rotating natural frequency in terms of the nonrotating
frequency ωvNR and the rise factor Kv :
Thus,
1/2
Kv ω − ωvNR + (Kv ω2 − ωv2NR )2 + 4(2 + Kv )ω2 ωv2NR
2 2
3
c2 =
2(2 3 + Kv )
(11.77)
This shaft critical speed does involve blade bending descriptors and corresponds
to the case wherein the shaft is deflecting normal to the blade axis and thereby
involves blade bending as well.
3) For the conditions wherein λ = i, a steady force in the fixed coordinate
system can excite a 1/rev (1P) response in the rotating coordinate system. The ram-
ification of this 1P response condition is that a 2P response will also be manifested
in the fixed coordinate system.
4) The implied methods for avoiding ground response with two-bladed rotors
are the same as those for the three-or-more-bladed case: Keep the blade natural
frequency subcritical (at a value above 1P), avoid values of rotor speed that result
in frequency coalescences with the pylon (fuselage) natural frequencies, reduce
the mass ratio, and provide damping in both the blade and pylon systems.
expression:
∗∗ γ ∗
β + β +ω2β β = 0 (11.79)
8
The (nondimensional) characteristic roots for this simple second-order system then
become
γ ! γ "2
λ=− ± i ω2β − (11.80)
16 16
where, using the rise factor representation for the rotating natural frequency.
1
ω2β = (ω2 + Kβ 2 ) (11.81)
2 βNR
A typical value for the rise factor Kβ of the first (flatwise) bending mode is 1.05.
These relationships can be combined to give the following expression for the
dimensional damped characteristic root:
γ
λ = λ = − ± i ωβ2NR + [Kβ − (γ /16)2 ]2 (11.82)
16
The behavior of the damped frequency ωβD [= Im(λ)] is shown in Fig. 11.26.
As this figure indicates, there is a certain value of Lock number for which the
flatwise frequency remains invariant. For Lock numbers greater than this value,
the damped frequency will eventually go to zero (at which point the characteristic
root pair becomes two aperiodic roots) for a sufficiently high value of rotor speed.
Fig. 11.26 Behavior of uncoupled flatwise bending frequency with rotor speed and
Lock number.
3) The coalescence of the edgewise modes with the pylon modes as a prime fea-
ture of the instability is not a distinguishing feature of air resonance. An instability
can be established without a clear-cut coalescence.
4) The air resonance phenomenon involves substantial coupling of the modes,
several of which are of relatively low frequency. Indeed, the modeling is capable
of simulating some of the low-frequency dynamics appropriate to evaluations of
vehicle dynamic stability and handling qualities. Although these roots are unstable,
they are not generally considered a structural dynamics problem because of the
pilot’s and/or autopilot’s ability to interact and stabilize them.
number ensures that the rotor has the correct aerodynamic damping and aerody-
namic coupling characteristics, and the advance ratio ensures that the scaling of
forward flight speed is correct in relationship to rotor rotational speed. The Froude
number ensures that the gravity effects, in terms of gravity springs and the rotor
thrust, are properly scaled in relation to the other three basic forces. The Froude
number is typically in the order of 500–700 and becomes increasingly important
with rotor size. Because the Froude number is relatively large compared with
the other nondimensional parameters, strict scaling of the gravitational terms can
sometimes be relaxed if their effects (as they relate to the phenomenon at hand)
can be approximated.
Scale factors of the pylon. For the rotor pylon to be properly scaled relative
to the rotor itself, it must present to the rotor a properly scaled impedance. This
can be achieved by matching 1) the mass ratios between the rotor mass and that
of the pylon, 2) any couplings existing between the hub degrees of freedom, and
3) the pylon natural frequencies (as nondimensionalized by the rotor frequency).
Mass ratios. Inspection of the air resonance equations shows that two mass
ratios of importance to the air-resonance instability phenomenon are 1) the already
discussed mass coupling parameter (involving rotor in-plane blade mode general-
ized mass and in-plane pylon effective mass) 3 and 2) a new parameter 4 . This
new parameter involves the rotor flapping blade mode generalized mass and pylon
rotational inertia, as taken about an effective focal point heff (see Fig. 11.12):
bS122
4 = (11.87)
2Ieff S10
where
and where the integrals S10 and S12 are respectively given by
R R
S10 = R2 m γw2 dr; S12 = R m rγw dr (11.89)
0 0
Pylon natural frequencies. The requirements for scaling the pylon impedance
also require that the natural frequencies (with respect to rotor rotational frequency)
of the pylon (with the constraint of it being focused at a point heff below the hub)
must be maintained. In the present context this scaling principle becomes important
when we might wish to alter the properties of an unscaled rotor also having an
unscaled focal point. Thus, for two configurations, ( )1 and ( )2 , that have the same
effective mass Meff , but are operating at different rotor speeds, 1 and 2 , the
frequency criterion then becomes that of maintaining the same effective rotational
stiffness.
The pylon stiffnesses in pitch and roll can each be approximated as a sum of
an explicit spring rate (around the focal point) Kp and an implicit one caused by
rotor flapping flexibility Kr . Furthermore, this rotor flapping spring rate Kr can
be conveniently expressed as a factor K̄r multiplying the rotor speed squared.
This factor is frequency dependent and proportional to the number of blades
and the already defined blade integration constant S12 . The K̄r factor depends
principally on Lock number γ , the nondimensional frequency of vibration ω,
and the blade flapping natural frequency ωw , again nondimensionalized by rotor
speed :
b
Kr = 2 K r S12 , γ , ω, ωw (11.90)
2
Then, the invariancy of pylon frequency criterion can be written as
Kp + K r 2
= invariant (11.91)
Ieff 2
2 , and M
But, Ieff equals Meff heff eff itself must also be invariant. Therefore, for
two different configurations that must both present the same impedance to the
rotor,
Kp / 2 + K r Kp / 2 + K r
2
= 2
(11.92)
heff heff
1 2
Noting that K̄r must be the same for both configurations, we can then
rewrite the expression to separate out the explicit stiffness rate for the second
configuration:
2 2
h eff K p heff
Kp2 = 22 2 1
+ K̄r 2
−1 (11.93)
heff1 21 heff1
It is readily apparent that, for 1 = 2 and heff1 = heff2 , the two explicit stiffnesses
are equal.
Fig. 11.28 Analytical predictions of air resonance stability variation for the BO-105
model with feathering axis precone (after Burkam and Miao).
Fig. 11.29 Experimental stability boundaries for the BO-105 model, showing
variations in collective angle and precone angle (after Burkam and Miao).
5) Active pitch control: The already present active stability augmentation sys-
tems in rotor control systems form the basic technology upon which an active air
resonance stabilization system can be configured. The frequency response required
of such a controller is well within the state of the art, and suitable feedback signals
can be obtained by accelerometer technology instrumentation. However, attention
must be paid to ensuring that such a stabilization scheme does not degrade existing
control system characteristics.
rates of the hub. The preceding expression defines an obliquely truncated cylin-
drical function over the rotor disk; the spanwise variable x defines the “tilt” of the
truncation plane.
The κν term represents an empirically formed function to account for the fore
and aft tilt of the truncation plane in forward flight caused by rotor lift (even in the
absence of rotor hub moments). Stated in physical terms, rotor lift in forward flight
produces an “upwash” in the forward part of the rotor disk and a “down-wash” in
the aft part of the disk. This fore and aft inflow distribution would be typical of any
lifting surface in forward flight. The empirical representation for κν , as formulated
by Payne, is as follows:
4µ
κν = (11.96)
3.6|λ0 | + 3µ
p
= 2 Vx2 + νz2 νi + he ν̇i (11.97)
ρ
where he is the effective height of a column of air that is being directly accelerated
(∂/∂t part of D/Dt), in contrast to the air that is being accelerated by convection
(U∂/∂x part of D/Dt). The actual value of he to be used depends on the type
of motion, but this issue will be deferred for the present. Invocation of the sixth
assumption given earlier allows the magnitude of the particle transport velocity
through the disk element to be expressed in the following manner:
λ0 ν̃i
Vx2 + νz2 ∼
= R µ2 + λ20 + (11.98)
µ2 + λ20
where
Vz + ν0
λ0 = (11.99)
R
Fig. 11.30 Idealized stream tube air flow through the rotor disk.
Here ν̃i , represents the part of the inflow velocity that can be assumed to be “small”:
Now let us expand the momentum equation, retaining linear terms where feasible
and nondimensionalizing by the rotor tip speed R. Also, from this point on the
nondimensionalization symbology is discarded:
p ∗
= 2 µ2 + λ20 ν0 + he0 ν 0
ρ(R)2
µ2 + λ20 + λ0 ν0 ∗
+ x 2 νc + he1 νc cos ξ
µ2 + λ20
µ2 + λ20 + λ0 ν0 ∗
+ x 2 νs + he1 νs sin ξ (11.101)
µ2 + λ20
Integrated hub loads. The integrated hub loads (rotor thrust, pitching, and
rolling moments) are simply formed from the incremental pressures. A simple
(very approximate) way of dealing with three-dimensional effects at the blade tips
is to delimit the extent of the integration at the blade tip means of the tip loss factor
B (typical value = 0.97). Also, at the same time the hub loads can be put into the
usual appropriate nondimensional rotor load coefficient form:
Rotor thrust:
1 2π B
CT = px dx dξ (11.102)
πρ(R)2 0 0
−1 2π B
CM = px 2 cos ξ dx dξ (11.103)
πρ(R)2 0 0
1 2π B
CL = px 2 sin ξ dx dξ (11.104)
πρ(R)2 0 0
CL µ2 + λ20
+ λ0 ν0 16 ∗
= νs + νs (11.107)
B3 2 µ2 + λ20 45π
CT
= | | ν0 + A (11.108)
2B2
where A is the nonsingular time-derivative part, and where
| |= µ2 + λ20 (11.109)
A
1 = sgn(CT )| ˆ | ν̂0 + (11.110)
CT /2B2
where
(· · · )
(· ˆ· ·) = (11.111)
|CT |/2B2
The basic thrust equation can then be rewritten to isolate the offending singularity as
1 A
ν̂0 = sgn(CT ) 1− (11.112)
|ˆ| CT /2B2
A
ν̂0 = sgn(CT )F ( ˆ ) 1 − (11.113)
CT /2B2
where the F ( ˆ ) is denoted the rotor-induced velocity function and for steady con-
ditions is the value of the equivalent induced velocity. In practice, the momentum
equation is reasonably valid in two operating conditions wherein there is substan-
tial generally uniform flow through the rotor disk. One is the thrusting propeller
state, wherein the flow is in a direction opposite to the thrust. The other valid case
is the windmill brake state, wherein the thrust is in the same direction as the flow.
These two cases define the two valid branches of the momentum equation that are
separated by the singular zero through-flow case and that must be joined by some
suitable empirical relationship.
Thus, the construction strategy is to join these two branches of the momen-
tum equation by some smooth nonsingular functionality. One way of achieving
the required approximation to this function is to use steady induced velocity test
data for rotors operating at or near the vortex ring state. Two sets of data for
rotors in vertical flight conditions are available for this purpose. A set of data col-
lected and presented in the work of Gessow and Myers and another set by Castles
and Gray (available in report form, see Bibliography for both works) provide the
basis for devising empirical fits to achieve a smooth transition between the two
branches of the momentum equation. The resulting complete functionality is given
in Fig. 11.31.
Observations that can be made from Fig. 11.31 are as follows:
1) Of the two experimental data sets available, the data set of Castles and Gray
shows the most marked deviation from the momentum equation. This curve is
probably more reliable in that the details of the test results are more accessible
and show features that are more consistent with actual identifiable flow details. All
subsequent observations are made relative to these data.
2) By normalizing the total through flow by the square root of the rotor thrust
coefficient (to give ˆ ), a measure of the relative magnitude of the thrust of the
rotor can be made. It is in the region of low absolute value of ˆ that the thrust can
be deemed to be “high,” in this case high relative to the valid momentum analysis
values.
3) The curve is strictly valid only in the vertical flight regime. However, the
empirical part of the curve deviates from the momentum equation only when the
nondimensional inflow is small, which occurs only in vertical flight. Indeed, for
both forward translational flight (helicopter operation) or forward axial flight (tilt-
rotor operation), the inflow parameter will assume high (negative) values, which
is seen to reduce the curve to the momentum analysis values.
4) The experimental bridge between the two momentum equation branches
at inflow parameter values near +1 is seen not only to have an infinite slope
but to be actually multivalued. This type of functionality, as described in a later
chapter, typically gives rise to hysteretic behavior in the dynamics. Indeed, such
behavior could well provide a mathematical vehicle for describing the unstable
flow conditions in the vortex-ring state.
5) The composite function assumes the role of a universal function servicing a
variety of rotor-related machines, both flight and nonflight (wind turbine) related.
6) The F = ˆ line corresponds to the hovering flight condition, and it is seen
that the fit of the data produces modification to the valid momentum value even at
hover.
7) In rotorcraft it is difficult to achieve “high” thrust conditions in the wind-
mill brake state (small positive values of ˆ ) because the actual inflow is typically
substantial, and the thrust is limited to the weight of the vehicle. However, in the
case of a stationary wind turbine neither of these limitations is ensured; indeed, the
thrust of a wind turbine is limited only by its stall conditions. Thus it is quite possi-
ble for a wind turbine to generate high thrust conditions with attendant deviations
from the momentum validity values of induced velocity.
8) For high positive values of the inflow parameter (representative of wind
turbine operation), the deviation of the experimental values from the momen-
tum validity values is consistent with the turbulent windmill state. In this
state the momentum equation requires a smooth deceleration and significant
streamtube expansion in the flow that is not attainable as a result of real flow
effects.
CT |CT | v0
A= − (11.114)
2B2 2B2 F ( ˆ )
Then the final (still nonlinear) working form of the thrust equation is
4 ∗ |CT | v0 CT
v0 + − 2 =0 (11.115)
3π 2B F ( ˆ ) 2B
2
This equation is now well behaved in that the F ( ˆ ) function never equals zero.
Also, examination of the moment equations shows that, for the critical vertical
flight condition where λ0 goes to zero (and concurrently, µ = 0), the similar terms
for vc and vs are seen to be proportional to the parameter V (a combination of v0
and λ0 , defined in the following section as the inflow mass flow parameter), which
is at least mathematically well behaved. However, more needs to be learned about
the behavior of the actual rotor moment characteristics in this critical condition.
Form 2:
∗
v 0
v0 CT
∗
[Ma ] v c + [L]−1 vc = −CM (11.117)
∗ vs CL
vs
where, based on the preceding development, the various matrices are given by
8
0
3π
16
[Ma ] = (11.118)
45π
16
0
45π
µ2 + λ20 + λ0 v0
V= (11.120)
µ2 + λ20
Pitt model. One air mass dynamics model that is often used because of its
high degree of mathematical consistency is that devised by Pitt and Peters. In this
model the [L] matrix is written in a form that smoothly transitions from hovering
flight to forward flight:
$
1 15π 1 − sin χ
− 0
2 64 1 + sin χ
$
1 15π 1 − sin χ
[L] =
4 sin χ
0
(11.121)
V 64 1 + sin χ 1 + sin χ
4
0 0
1 + sin χ
where
(δvj = [δv0 , δvc , δvs ] and 2) a perturbational form of the air mass equations with
coupling terms relating the perturbations of the rotor hub loads accruing from per-
turbations of the vector of the selected rotor aeroelastic variables (δqR ). In matrix
form this implementation would have the following form:
..
A 11 (λ) . A 12 (λ) δqR
.
· · · · · · · · · · · · .. · · · · · · · · · · · · . . . . . . = 0 (11.123)
.. δvj
A (λ)
21 .
22 A (λ)
where
A11 (λ) ≡ basic aeroelastic equations for perturbational blade (modal) responses
= [M]λ2 + [C]λ + [K] (11.124)
A12 (λ) ≡ additional excitations to modal responses from perturbational air mass
dynamics velocities (arising from perturbational quasisteady blade air-
load distributions)
A21 (λ) ≡ perturbations in the integrated hub loads caused by perturbations in the
blade modal responses
∂(CT , −CM , CL )
=− (11.125)
∂(qR )
∂(CT , −CM , CL )
A22 (λ) = [Ma ]λ + [L]−1 − (11.126)
∂(vj )
References
Section
11.1 Hohenemser, K. H., and Yin, S.-K., “Some Applications of the
Method of Multiblade Coordinates,” Journal of the American
Helicopter Society, Vol. 17, No. 3, 1972, pp. 3–12.
11.2 Houbolt, J. C., and Reed, W. H., “Propeller-Nacelle Whirl Flut-
ter,” Journal of the Aeronautical Sciences, Vol. 29, No. 3,
1962, pp. 333–346.
Johnson, W., “A Comprehensive Analytical Model of Rotorcraft
Aerodynamics and Dynamics, Part I: Analysis Development,”
NASA TM-81182, 1980.
Kvaternik, R.G., “Studies in Tilt Rotor VTOL Aircraft Aeroe-
lasticity,” Ph.d. Dissertation, Dept. of Solid Mechanics,
Structures and Mechanical Design, Case Western Reserve
Univ., Cleveland, OH, 1973.
References (continued)
Popelka, D. A., Sheffler, M. W., and Bilger, J., “Correlation of
Test and Analysis for the 1/5-Scale V-22 Aeroelastic Model,”
Journal of the American Helicopter Society, Vol. 32, No. 2,
1987, pp. 21–33.
Reed, W. H., “Propeller-Rotor Whirl Flutter: A State-of-the-Art
Review,” Journal of Sound and Vibration, Vol. 4, No. 3, 1966,
pp. 526–544.
Wernike, K. G., and Gaffey, T. M., “Review and Discussion of
‘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, 1967,
pp. 55–57.
Young, M. I., and Lytwyn, R. T., “The Influence of Blade Flap-
ping Restraint on the Dynamic Stability of Low Disk Load-
ing Propeller-Rotors,” Journal of the American Helicopter
Society, Vol. 12, No. 4, 1967, pp. 38–54.
References (continued)
Burkam, J. E., and Miao, W., “Exploration of Aeroelastic Stabil-
ity Boundaries with a Soft-in-Plane Hingeless Rotor Model,”
Journal of the American Helicopter Society, Vol. 17, No. 4,
1972, pp. 27–35.
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, 1969, pp. 33–41.
Hunt, G. K., “Similarity Requirements forAeroelastic Models of
Helicopter Rotors,” Aeronautical Research Council, London,
CP-1245, 1972.
Lytwyn, R. T., Miao, W.-L., and Woitsch, W., “Airborne and
Ground Resonance of Helicopter Rotors,” Journal of the
American Helicopter Society, Vol. 16, No. 2, 1971, pp. 2–9.
Lytwyn, R. T., “Aeroelastic Stability Analysis of Hingeless
Rotor Helicopters in Forward Flight Using Blade and Air-
frame Normal Modes,” Proceedings of the 36th Annual
National Forum of the American Helicopter Society, Paper
79-23, 1979.
Miao, W.-L., and Huber, H. B., “Rotor Aeroelasticity Coupled
with Helicopter Body Motion,” Proceedings of Specialists’
Meeting on Rotorcraft Dynamics, NASA Ames Research
Center, 1974.
11.5 Castles, W. Jr., and Gray, R. B., “Empirical Relation Between
Induced Velocity, Thrust and Rate of Descent of a Helicopter
Rotor as Determined by Wind-Tunnel Tests on Four Model
Rotors,” NACA TN 2474, 1951.
Gaonkar, G. H., Sastry, V. V. S. S., Reddy, T. S. R., Nagab-
hushanam, J., and Peters, D. A., “The Use of Actuator-Disc
Dynamic Inflow for Helicopter Flap-Lag Stability,” Journal
of the American Helicopter Society, Vol. 28, No. 3, 1983,
pp. 79–88.
Gessow, A., and Myers, G. C., Aerodynamics of the Helicopter,
Ungar, New York, 1967.
Lamb, H., Hydrodynamics, Dover, New York, 1932.
Pitt, D. M., and Peters, D. A., “Theoretical Prediction of
Dynamic-Inflow Derivatives,” Vertica, Vol. 5, No. 1, 1981,
pp. 21–34.
Rao, B. M., and Jones, W. P., “Application to Rotary Wings of a
Simplified Aerodynamic Lifting Surface Theory for Unsteady
Compressible Flow,” Proceedings of Specialists’ Meeting on
Rotorcraft Dynamics, NASA Ames Research Center, 1974.
Problems
11.1 A simple rotor-nacelle propulsion system has a four-bladed rotor with
constant chord blades and the following specifications:
11.2 Using the basic (rigid-blade) rotor-nacelle whirl flutter equations of motion
(as simplified to the isotropic case), show that the whirl instability must be
in the retrograde direction. (Hint: Combine the two degrees of freedom into
a single complex variable, Z = ψ + iθ , and assume sinusoidal motion.)
11.4 Verify that the product of damping stability criterion given in Sec. 11.3.2
reduces to the criterion used by Gabel and Capurso given in Sec. 11.3.3.
11.5 Derive the product of damping stability criterion given for the two-bladed
rotor case.
11.6 Using the momentum equation for hub (pitching) moment in hover, derive an
expression for the lift deficiency function [≡(CM )unsteady /(CM )QS ]. Note:
use the following assumptions and considerations:
1) A separate expression for pitching moment as a result of pitch rate
(using strip theory) is required:
4) ∗
vc = 0 (11.127d)
11.7 Consider the tail rotor of a helicopter in yawing (hovering) flight at stan-
dard sea-level conditions. The (two-bladed) tail rotor specifications are as
follows:
Find the percent loss in first harmonic aerodynamic effectivity (lift deficiency
function) for respective helicopter yaw rates of ±100 deg/s. Note: rotor
inflow for the rotor ascending (i.e., moving in the same direction as the
thrust) at an ascent speed Vz is given by
CT
λ= (11.128)
2B2 (λ + Vz / R)
Nondimensional Reduced
Phenomenon Rotor type frequency ω/ frequency k
relate to the variability of flow in the two directions have been assimilated from the
fixed-wing case. However, the effects of large response amplitudes of the airfoil
section in the rotary-wing case have not as yet been fully addressed, and the fixed-
wing solutions, which have almost universally assumed infinitesimal motion, are
suspect and of limited usefulness.
4) Near- and far-field effects: Because of the rotor-blade’s ability to be influ-
enced by past disturbances from both itself and other blades, the rotor-blade wake
is composed of a near-field wake that consists of the trailing and shed vortex system
extending a few chord lengths behind the trailing edge of the blade and a far-field
wake that is incorporated with those from the other blades into a total rotor wake.
The modeling of this rotor wake has been approached from two directions: 1) the
concept of air mass dynamics, as treated in an earlier section, and 2) a detailed
flow calculation built up using vortex elements.
portion of this chapter; later portions of the chapter deal with two aeroelastic
instabilities, each of which require an appropriate form of unsteady aerodynamics
for an accurate analysis: bending-torsion flutter and stall flutter.
Fig. 12.2 Real and imaginary parts of the Theodorsen function F(k) and G(k),
respectively.
second part contains circulatory terms and looks like a static airfoil coefficient
except with a multiplication by the Theodorsen function C(k). This term in these
expressions has the form of an attenuation factor on what would otherwise be
static (circulation-dependent) airfoil coefficients. Hence, the Theodorsen function
is generalized to represent a lift-deficiency function, with the understanding that
alternate lift-deficiency functions from other sources might be used in the same
manner as that stated in these equations.
Virtual mass terms. The inertia-like parts of the expressions given earlier
for lift and pitching moment are usually referred to as the noncirculatory or virtual
mass terms and represent the reactive forces exerted by the “equivalent” mass
and inertia of the air mass surrounding the airfoil. For rotary-wing applications
the doubly differentiated acceleration-like terms (¨) are typically small relative to
the blade’s mass properties and are thereby neglected. The single-differentiated
velocity-like terms (˙) are relatively small in the expression for lift, but are large in
the expression for the pitching moment. In fact, the principal source of damping
in pitch comes from this virtual mass term:
1
Mθ = · · · − Ub 2 − a θ̇ − · · · (12.7)
be motionless in both pitch and plunge. The unsteadiness of the problem is instead
afforded by the sinusoidal vertical gust pattern imbedded in the air mass into which
the airfoil is moving, as shown in Fig. 12.3.
Thus, if the coordinate axes are fixed to the airfoil, the vertical gust can be
represented by a (vertical) velocity distribution flight U of the following form:
The waviness of the sinusoidal gust can alternatively be expressed using the wave-
length of the imbedded gust pattern or the frequency ω, with which the wave
passes any point of the airfoil:
ω = 2π U/ (12.14)
The results of Sears’ formulation give expressions again for unsteady lift and
pitching moment about the midchord. Sears’ formulation is unlike the Theodorsen
formulation, however, in that it is devoid of virtual mass terms (which is reasonable
because the airfoil is not moving) and similar to it in the appearance of a lift-
deficiency functionality to describe the unsteady features of the problem:
Figure 12.4 shows the variation of the Sears function in the complex plane together
with the Theodorsen function for comparison. Note that the asymptotic values for
the Sears and Theodorsen functions are, respectively, (0 + 0i) and (0.5 + 0i) as
k approaches ∞. As can be seen by this comparison, for frequencies typical of
helicopter aeroelastic problems there is little difference between these two lift-
deficiency functions. However, the figure does serve to highlight the difference
between the effects of unsteady angles of attack caused by airfoil motion (θ̇ ,
θ, and ḣ/U, i.e., the Theodorsen problem), compared with those of an angle
of attack caused by air mass motion (inflow angle φ, i.e., the Sears problem),
Fig. 12.4 Vector diagram showing the real and imaginary parts of Sears’ φ(k) and
Theodorsen’s C(k) functions as functions of the reduced frequency k.
where the total angle of attack α is generally a combination of both basic types of
effects.
Example 12.1
Consider a simple two-dimensional airfoil taken as a spring mass damper con-
strained to pivot about a (variable) point located at a distance xa behind the
aerodynamic center (quarter-chord). Thus, the airfoil is constrained to have zero
plunge (ḧ = ḣ = 0), as shown in the following sketch:
When the doubly time differentiated virtual mass terms are omitted, the equation
of motion for this system can be stated very simply as
Iθ θ̈ + cθ θ̇ + kθ θ = Mθ (12.20)
where the (sinusoidal) aerodynamic moment Mθ resulting from the θ motion (of
amplitude θ̄ ) is given by
2x
a 2xa ik 1 − 2xa
Mθ = A 1 − ik 1 − C̄(k)eiη(k) − θ̄ (12.21)
c c 2 c
(= F+iG)
or
ik
Mθ = A XR − S θ̄ (12.23)
2
where
R = F − Gk(1 − X) (12.24a)
S = (1 − X)(1 − 2XF) + X(−2G/k) (12.24b)
The AXR quantity is then the effective aerodynamic spring, and the S factor is a
measure of the pitch-damping effectivity.
Because the possible range of X is from −0.5 (leading-edge pivot) to +1.5
(trailing-edge pivot), the R factor is always positive. Therefore, as expected, the
aerodynamic spring stiffness is proportional to the negative of X (i.e., a pivot point
ahead of the quarter-chord yields a stable “restoring” spring). Let us then examine
the behavior of the S factor for four principal pivot axis locations:
X = − 21 (corresponds to pivoting about the leading edge): For this value of X,
the S factor is given by
Fung presents a handy tabulation of the Theodorsen function as well as the function
(−2G/k); from this tabulation it can be shown that for reduced frequencies less
than approximately 0.03, the quantity (−2G/k) is greater than 3(1 + F) (i.e., ≈6).
Therefore, for configurations with negligible mechanical damping, the airfoil has
negative pitch damping and is therefore unstable for these low values of reduced
frequency.
X = 0 (pivot location at the quarter-chord): For this value of X, the aerodynamic
spring rate goes to zero, and the pitch-damping effectivity S is equal to 1. Thus,
the system is unconditionally stable.
X = + 21 (pivoting about the midchord): For this case the aerodynamic spring
is unstable, but the pitch-damping effectivity is given by
which is positive for all values of k. Thus, for an airfoil that has sufficient mecha-
nical stiffness, the airfoil would be stable.
X = +1 (pivoting about the three-quarter-chord point): This case corresponds
to the case wherein the principal part of the pitch damping (accruing from the
virtual mass terms only) would be zero. In this case, however, the pitch-damping
effectivity is equal to (−2G/k), which is always positive, and the airfoil is still
stable. The variations already given serve to demonstrate how the phase lag of
the unsteady aerodynamics can destabilize an otherwise stable configuration and
stabilize an otherwise neutrally stable one.
axial translational speed Vz . If this inflow velocity is assumed constant over the
rotor disk, standard momentum considerations give the following expression for
the inflow velocity u of the rotor:
Vz Vz2 T
u= + + (12.27)
2 4 2Aρ
T
CT = (12.29)
π R2 ρ(R)2
Review of the Loewy theory. The returning wake concepts developed ear-
lier are incorporated in an approximate two-dimensional manner in the theory of
Loewy. Historically, this analysis was the first successful attempt to incorporate
the returning wake into the unsteady aerodynamics problem. Fortunately, because
only the hovering/vertical flight condition was considered, a closed-form solution
was possible. In addition to the assumptions inherent in the Theodorsen problem, it
is assumed here that the returning wake can be represented by a countably infinite
number of infinite (in length) sheets of vorticity parallel to the rotor plane as shown
in Fig. 12.7.
""
Q−1 ∞ ! ∞
γnq (ξ , t)(x − ξ ) dξ
+ 2
−∞ (x − ξ ) + (nQ + q) h
2 2
q=0 n=0
∞ !
#
" ∞ γn0 (ξ , t)(x − ξ ) dξ
+ (12.30)
(x − ξ )2 + n2 Q2 h2
n=1 b
Neglecting the last two terms, which are a result of the returning wakes, reduces the
formulation to that of Sec. 12.2.1 (the Theodorsen problem). The usual unsteady
aerodynamics assumption of sinusoidal motion is made at this point:
The strengths of the trailing and returning wakes are then related to the strength
of the bound vorticity by means of 1) the phasing of the motion between the rotor
blades ψq ; 2) the spacing of the returning wakes h; 3) the number of rotor blades Q;
and 4) the reduced frequency k. After a nondimensionalization of the lengths by
the semichord b, the preceding integral equation then reduces to
! +b ! ∞ −ikx #
1 γ̄a (ξ ) dξ e dξ
v̄a (x) = − C − ik ¯ ¯
+ π k e −ikx
W (12.32)
2π −b x − ξ 1 x−ξ
where the amplitude to the total vorticity over the blade section ¯ is given by
! +1
¯ = eik γ̄a (x) dx (12.33)
−1
and W , which accounts for the total effect of the returning wake, is given by
$ %
Q−1 &(Q−q)/Q
1+ ekhQ ei2πm eiψq
q=1
W= (12.34)
ekhQ ei2πm − 1
In a like manner, as was used for the fixed-wing case (Theodorsen problem),
the preceding integral equation for ¯ is solved using Söhngen’s inversion formula.
The total circulation over the airfoil then becomes
! +1
2 (1 + ξ /1 − ξ )v̄a (ξ ) dξ
¯ −1
= ' ( (12.35)
(2) (2)
iπ 1/2 H1 (k) + iH0 (k) + [J1 (k) + iJ0 (k)] W
Bernoulli’s equation, which relates the solution variable (vorticity) to the desired
working quantity (pressure), for sinusoidal motion is given by
! x
p̄(x)
= −γ̄ (x) − ik γ̄a (ξ ) dξ (12.36)
ρV −1
Use of this equation then yields the following expression, which represents the
desired solution as a function of the known (or prescribed) velocity at the airfoil
section:
(2)
−p(x) 2i/π [H0 (k) + 2J0 (k)W ]
= (2)
ρr (2)
H1 (k) + iH0 (k) + 2[J1 (k) + iJ0 (k)]W
! +1 )
1−x 1+ξ
× v̄a (ξ ) dξ
−1 1+x 1−ξ
! ) #
2 +1 1−x 1+ξ 1
+ C − ik1 (x, ξ ) v̄a (ξ ) dξ (12.37)
π −1 1+x 1−ξ x−ξ
√ #
1 1 − xξ + 1 − ξ 2 1 − x 2
1 (x, ξ ) = n √ (12.38)
2 1 − xξ − 1 − ξ 2 1 − x 2
The preceding equation for p̄ can be shown to be of the same form as a cor-
responding equation for the two-dimensional fixed-wing problem discussed in a
preceding section. Therefore, a modified lift-deficiency function exists and can be
expressed in the following general form:
Fig. 12.8 Real and imaginary parts of the Loewy function vs k as functions of inflow
parameter (m = 0) (after Loewy).
A significant and useful result of this formulation is that the modified (Loewy) lift-
deficiency function C can be directly inserted into the simple expressions for rotor-
blade lift and moment distributions given in Section 12.2.1 to give a more accurate
flutter/aeromechanical stability calculation for hovering or vertically ascending
rotors. Furthermore, the Loewy analysis represents a base upon which even more
accurate formulations can be formulated.
Fig. 12.9 Real and imaginary parts of the Loewy function vs k as functions of inflow
parameter (m = 1/4) (after Loewy).
J1 (k)
C = (12.42)
lim J1 (k) + iJ0 (k)
m→integer
h→0
h 1
C = = (12.43)
lim h+π 1 + (π σ/4|λ|)
m→integer
k→0
Fig. 12.10 Real and imaginary parts of the Loewy function vs k as functions of inflow
parameter (m = 1/2) (after Loewy).
It can be verified that the minimum (inverted) peaks shown in Fig. 12.12 for flap
damping, when computed for nonzero values of k, are closely approximated by
this formula.
3) At nonintegral harmonics of the oscillatory frequency, as the spacing h
approaches zero, and as the reduced frequency k approaches zero (by virtue of
r → ∞), the lift-deficiency function approaches unity:
C =1 (12.44)
lim
m→integer
h→0
k→0
As a practical aeroelastic tool for rotors, the Loewy theory has four inherent
limitations.
First, it does not account for oscillations in the streamwise air velocity and is
therefore not valid for forward-flight conditions. Second, being a two-dimensional
the theory is again based on the Theodorsen formulation approach, it must treat
aeroelastic problems only in the frequency domain and cannot easily treat prob-
lems of arbitrary motion in the time domain. The theory nevertheless represents
a substantial achievement in unsteady aerodynamics and is a pioneering effort in
the development of a truly rotorcraft-oriented theory. Moreover, this theory has
become a touchstone for newly emerging theories that continue to relax the basic
assumptions.
The crucial assumption of sinusoidal motion of the aerodynamic environment
is a basic impediment to an expanded, more generally applicable unsteady aerody-
namic formulation. In the following material concepts are presented that provide
basic arbitrary motion aerodynamic formulations. It is hoped that these concepts
will eventually lead to such a general unsteady aerodynamic formulation.
3) The lift-deficiency function C(k) is defined only for simple harmonic motion,
that is, in the frequency domain.
We require an unsteady aerodynamic formulation that is described entirely for
arbitrary airfoil motion. To this end, four possible approaches present themselves
and are respectively described in the subsequent material: 1) use of the Theodorsen
function together with a rigorous application of Fourier transform (integral) tech-
niques; 2) use of the Wagner function (derivable from the Theodorsen function)
together with the Duhamel integral; 3) direct use of the generalized Theodorsen
function C( p), with a (complex) Laplace transform argument; and 4) use of an
approximate Wagner function, again using Laplace transform techniques.
The principal disadvantage with this method is that the various integrals, although
completely valid for any arbitrary motion, are quite cumbersome to evaluate for any
general arbitrary vertical velocity w3c/4 (t). The method described in the following
section circumvents this disadvantage by performing the required Fourier integral
transform once for a basic type of transient motion, the unit step function, and then
by building up arbitrary motion using superposition.
Wagner function and Duhamel integral. Wagner used the technique of the
previous Fourier integral method for a step change in angle of attack:
0, t<0
w3c/4 (t) = (12.49)
Uα0 , t ≥ 0
The Wagner function and the exponential approximation attributed to Jones are
depicted in Fig. 12.13. Use of the Duhamel or superposition integral enables the air-
loads resulting from any arbitrary angle-of-attack time history to be mathematically
constructed from the Wagner function:
where
! s dw3c/4 (σ )
Wφ = w3c/4 (0)φ(s) + φ(s − σ ) dσ (12.53)
0 dσ
Although these results represent a considerable improvement over the pure Fourier
integral approach, they are as yet sufficiently complex to preclude their use in a
flutter (instability) analysis. This type of formulation typically finds application
mainly in the calculation of transient loads wherein the airfoil motion is known a
priori.
Besides giving a fairly accurate fit to the “exact” Wagner function, this approxi-
mation has two desirable characteristics. First, it has the character of a decaying
functionality, which is entirely consistent with the physics of the unsteady flow
phenomenon; disturbances at the airfoil section are convected away from the sec-
tion by the airstream so that their influences are monotonically lessened. Second,
from a mathematical point of view the approximation has a straightforward easily
calculable Laplace transform. The Laplace transform is defined as
! ∞
L[ f (t)] = F(p) ≡ f (t)e−pt dt (12.59)
0
Therefore,
! ∞
b
L[φ(Ut/b)] = [1 − 0.165e−0.0455s − 0.355e−0.3s ]e−psb/U ds
U 0
0.5p2 + 0.281(U/b)p + 0.0136(U/b)2
= (p) = (12.60)
p[ p + 0.3(U/b)][ p + 0.0455(U/b)]
where p behaves like d( )/dt.
Then, the Laplace transform for lift and moment can be written as
(12.63)
The preceding expressions for the loads (lift and pitching moment) are stated
in the form of transfer functions, wherein the loads are expressed as Laplace
transform operations of the responses. This has the advantage that the (input)
responses do not need to be explicitly stated, thus making the expressions quite
useful for eigenvalue analyses.
where, after performing an integration by parts on Eq. (12.64), the effective angle
of attack αE can be put into the following form:
! s
1
αE = αQS + αQS (σ )φ (s − σ ) dσ
2 0
! s
1
= αQS − αQS − αQS (σ )φ (s − σ ) dσ
2 0
= αQS − αW (12.66)
With the inclusion of the unsteady decay parameter, the effective angle of attack
αE can then be used in a quasi-steady manner. It can be easily verified that the
Wagner problem is duplicated using the unsteady decay parameter when the quasi-
steady angle of attack is a step function. Moreover, the role of this parameter is
where β = (1 − M 2 ).
As stated earlier, and as the following development will make more apparent,
we wish to work with the actual lift-curve slope clα instead of the theoretical
incompressible value (≈2π). Hence, the division by β in the preceding expression
will be omitted because it will be inherently included with the actual (Mach-
number-dependent) lift-curve slope. The compressibility corrected unsteady decay
parameter then becomes
! s
αW ≡ 1/2 αQS − αQS (σ )βφc (s − σ ) dσ (12.69)
0
quasi-steady assumption wherein the inflow angle is calculated using the air
velocity components at the three-quarter-chord point. These two variables are then
used (with appropriate interpolation schemes) to look up the aerodynamic section
coefficients using available static airfoil tables. The wind-tunnel testing of airfoils
over ranges of α and M appropriate to the operational envelope of the rotorblade
sections has become a routine part of new helicopter development. Consequently,
a large body of airfoil data, as functions of α and M, currently exists.
Rotor aeroelastic analyses incorporating such table look-up schemes generally
have the advantage of providing a more accurate modeling of the low- (reduced-)
frequency airloads needed for trim calculations and for low-frequency aero-
mechanical instability analysis. Furthermore, these time-history analyses provide
a straightforward, practical vehicle for solving rotor aeroelastic problems, both
vibration and stability related, which are governed by strong nonlinearities. The
analyses of such nonlinear phenomena, using time-history solution techniques, are
treated in subsequent sections. Note that, as developed in the preceding section,
eigenvalue analyses that use a linear form of airloads (lift-curve slope times angle
of attack) together with a Laplace-transformed Wagner function can be formulated.
For such analyses the effective angle-of-attack approach described herein offers
no advantage since the two methods are actually identical.
The basis for the use of the effective angle-of-attack method is the facility in
solving for the time history of this variable quantity using an appropriate differ-
ential equation. To this end, the convolution integral representation given earlier
is an ideal definition because it allows a transfer function to be formed and can
also be calculated using a recursive numerical time-marching scheme. When we
take the Laplace transform of the effective angle of attack (relative to aerodynamic
time s), the following relationships are formed:
1 0.007508 0.1005
ÂE = ÂQS + β 2 + ÂQS (12.70)
2 p̂ + 0.0455β 2 p̂ + 0.3β 2
where
ÂE = L(αE ); L( ) ≡ Laplace transform with respect to s
ÂQS = L(αQS )
p̂ ≡ (aerodynamic time) Laplace transform variable
[equivalent to d( )/ds]
To complete the formulation, the aerodynamic time s must be related to real
time t. For fixed-wing problems wherein the streamwise air velocity is a constant,
the aerodynamic time is merely proportional to real time by the factor (U/b). For
rotary-wing problems, however, the streamwise velocity is generally variable, and
the relationship between the two timescales must be taken differentially.
ds = [2U(t)/c] dt (12.71)
Then the real-time differential operator p[=d( )/dt] can be formed by means of
the following substitution:
d( ) d( ) dt
p̂ = = · = [b/U(t)] p (12.72)
ds dt ds
1 0.007508 0.1005
AE = AQS + β 2 (U/b) + AQS
2 p + 0.0455β 2 (U/b) p + 0.3β 2 (U/b)
(12.73)
Numerical solution for the effective angle of attack in the time domain.
The calculation of the effective angle of attack in time-history analysis typically
requires some form of (time-marching) numerical integration algorithm for the
unsteady decay parameter. The appropriate time-marching variable for our pur-
poses is the aerodynamic time variable s. Thus, the kth time-marching integration
interval sk must be related to real time t as follows:
2U(t)
(s)k = t (12.74)
c k
Taking this (nondimensional) time interval sufficiently small to ensure both numer-
ical stability and accuracy, and the freestream velocity U(ψ) constant over the
time interval, leads to a recurrence relationship for calculating the unsteady decay
parameter. The numerical recursive representation of the unsteady decay parameter
begins by breaking the parameter into a sum of two parts:
(αW )k = Xk + Yk (12.75)
where the formulations for Xk and Yk begin with the following Duhamel integral
expressions:
! s
X(s) = A1 αQS + a1 β 2 αQS e−a1 (s−σ ) dσ (12.76a)
0
! s
Y (s) = A2 αQS + a2 β 2 αQS e−a2 (s−σ ) dσ (12.76b)
0
where the constants A1 and A2 are respectively given by 0.165 and 0.335 and a1
and a2 are respectively given as 0.0445 and 0.3. Following development which
parallels that given in Leishman, and averaging the quasi-steady angle of attack
over each integration step, leads to the following expressions for marching the
solution from the (k − 1)th step to the kth:
include any plunge motion. But within the limitations of these assumptions, it
is a rigorous theory and forms a credible basis for evaluating all of the various
theoretical methods. This theory is sufficiently involved to preclude its detailed
description in this text.
Van der Wall and Leishman performed a comprehensive comparative study of
all of the various methods in the literature for addressing this problem in 1994 and
introduced a new indicial response methodology based on the use of Duhamel’s
integral [Eq. (12.64)]. As regards the evaluations of the preexisting theories, van
der Wall and Leishman concluded that within the range of reduced frequencies
typically encountered by helicopter rotor blades the various theories examined
gave equivalent results. Figures 12.16a–12.16c compare the results obtained with
the newer indicial response methodology with those obtained with the Isaacs theory
for constant, cosine, and sine variations in pitch angle. The results clearly show
the indicial response method to be quite accurate compared with the “touchstone”
Isaacs theory for a) constant pitch angle, b) sine variation, and c) cosine variation.
Both the apparent mass terms and the circulatory terms must be generalized.
Thus, Eqs. (12.51) and (12.52) can be rewritten as
d(Uθ)
L = πρb2 ḧ + − baθ̈ + 2πρbWφ (12.80)
dt
d(Uθ)
Mθ = πρb2 baḧ − b(1/2 − a) − b2 (1/8 + a2 )θ̈ (12.81)
dt
+ 2πρUb(1/2 + a)Wφ
where Wφ is the Duhamel integral, as given by Eq. (12.53), but the derivative
of the normal velocity at the three-quarter point [Eq. (12.45)] with respect to the
aerodynamic time variable σ is now given by a more general form:
dw3c/4 (σ ) dḣ d [U(σ )α] 1 − 2a dα̇(σ )
= + +b (12.82)
dσ dσ dσ 2 dσ
As presented by Leishman, the approximate Wagner function, defined by Eq.
(12.58), and Duhamel’s integral are used to calculate the time variation of the
equivalent angle of attack αe , that is,
! s
dα
αe (s) = α(s0 )φ(s) + (σ )φ(s − σ ) dσ (12.83)
0 dσ
Fig. 12.16 Comparison of results from Isaacs theory with results using the indicial
response method, varying amplitudes of freestream oscillation, and form of pitch angle
oscillation (after van der Wall and Leishman).
where αe can then be used in a quasi-steady sense. Leishman’s method for obtaining
a solution assumes first that the α(s0 ) term can be neglected because it represents
short-lived startup transient conditions. (Although somewhat different in detail,
this formulation leads to an expression for effective angle of attack that is com-
pletely compatible with that defined using the unsteady decay parameter.) The
solution is taken to be the sum of three terms, where here α is taken to be the
quasi-steady value:
where
! s dα
X(s) = A1 (σ )e−b1 (s−σ ) dσ (12.85a)
s0 ds
! s
dα
Y (s) = A2 (σ )e−b2 (s−σ ) dσ (12.85b)
s0 ds
where s is given by
!
2 t+t Uk + Uk−1
s = U(t) dt = t (12.87)
c t c
Fig. 12.17 Results for a constant angle of attack in an oscillating flow, compari-
son between the indicial response method and using an Euler CFD code (after Jose,
Leishman, and Baeder).
$
N+1
aj (iω) j
j=1
C(ω) = (12.88)
$
N
bj (iω) j
j=1
where N is the order of the approximant. The degree of the denominator is lower
than that of the numerator by 1 because of the known asymptotic behavior at large
frequencies. Note that the coefficient C(ω) is generally complex.
As extracted from the work of Bielawa et al., Fig. 12.18 presents an example of
the lift coefficient variations with reduced frequency for a two-dimensional airfoil
at a high subsonic Mach number. These data were obtained first from a theoretical
prediction, using the work of Jordan, and then from a suitably calculated Padé
approximant (approximating the Jordan results).
Basic methodology. The reason that the Padé approximant is so useful is that
the Padé form, when approximating the aerodynamic coefficient’s variation with
respect to iω, can actually provide a reasonably accurate analytical continuation
of the coefficient off the imaginary axis into the positive and negative halves of
the complex plane. Thus, after the Padé form has been successfully found, subject
to the preceding itemized constraints, then iω can be replaced by λ(= σ + iω),
or, alternatively, by the Laplace transform variable p. Most unsteady aerodynamic
data are presented in the form of variations with respect to reduced frequency k,
which then become analytically continued with respect to the Laplace transform
variable based on aerodynamic time p̂. The steps in utilizing the approximant are
basically as follows:
1) Obtain the variation of the (complex-valued) aerodynamic coefficient(s) with
respect to frequency ω (or k). This variation represents the coefficient taken as
A2 p A3 p
Cx = A1 + + + ··· QS
p + p̃1 p + p̃2
B3 p B4 p
+ B1 + B2 p + + + ··· QS (12.89)
p + p3 p + p4
where for this most general case QS and QS are, respectively, the Laplace
transforms of the quasi-steady inflow and geometric pitch angles comprising
the quasi-steady angle of attack and Cx refers to lift or moment coefficient, as
appropriate.
As with the formulations for the unsteady decay parameter, the Padé formula-
tions can be used either in Laplace variable eigenvalue solutions or in time-history
solutions. The approximants also provide a convenient interpolation formula
for obtaining values of the complex coefficients at frequencies other than those
available from the analytical or experimental source.
Fig. 12.19 Typical blade section showing parameters defining flutter and divergence
characteristics.
degrees of freedom require that chordwise offsets of the aerodynamic and mass
centers relative to the elastic (shear) center yAC and yCG , respectively, be taken into
consideration. These features are depicted in Fig. 12.19, which shows a typical
blade section.
Pitching motion:
where the perturbational lift L and (noncirculatory) moment Mθnc are respectively
given by
* #
c2 c
L= πρ −z̈ + U θ̇ + − yAC θ̈
4 -- - 4- - - - - - - - - -
c c (
+ clα ρU C(k) −ż + Uθ + − yAC θ̇ (12.92)
4 2
c2 c c c 3c
Mθnc = πρ z̈ − U θ̇ − − yAC θ̈ (12.93)
4 4- - 2 4 8
---------------
The underlined terms are neglected because they are small compared with the
already present inertia terms. Furthermore, the circulatory term in the expression
for lift [∼C(k)] can be treated as a frequency-domain lift-deficiency function (for
frequency-domain solution techniques), or as a Laplace transformed function using
the approximate Wagner function (for general matrix eigenvalue solutions). For
example, the preceding flutter equations can be expressed in the following form
using the Laplace transform variable approach:
m myCG c2 0 −1 kz 0
p2 + πρU p+
myCG (Iθ + myCG
2 ) 4 0 c/2 0 kθ
c 0.5p2 + 0.281(U/b)p + 0.01365(U/b)2
+ clα ρU
2 [ p + 0.3(U/b)][ p + 0.0455(U/b)]
1 −(c/2 − yAC ) 0 −1 z̄
× p+ =0 (12.94)
yAC −yAC (c/2 − yAC ) 0 −yAC θ̄
This equation can then be expanded into a matrix eigenvalue problem of canonical
form.
two complex conjugate pairs will generally result. Figure 12.20 shows the typical
behavior of these two root pairs with increases in the nondimensional parameter
2U/cωθ , where
Presented without derivation are the following two general equations for the
flatwise bending and elastic torsion modal responses. Note that the equation
derivation assumes that the effects of control pitch angle θ0 and steady coning
β0 are negligible. The equations initially show only the elastomechanical terms in
detail.
Flatwise bending:
Elastic torsion:
1) This term represents the flatwise stiffening caused by centrifugal force and
is the source of the already considered Southwell coefficient or rise factor.
2) This term represents a moment caused by the centrifugal force operating in
the presence of a moment arm provided by the torsional deflection. A loading is
produced by the spanwise derivative of this moment:
3) This term represents a torsional moment about the (inclined) elastic axis,
where the inclination is caused by flatwise bending. The moment is thereby a
result of a component of the radially directed centrifugal force in the direction
normal to the elastic axis taken together with the c.g. offset.
dL(r) c2 ρ
= πρ r γθk (r)q̇θk + clα cr C̄(k) − γwm (r)q̇wm
dr 4 2
c
+ r γθk (r)qθk + − yAC γθk (r)q̇θk (12.99)
2
Noncirculatory pitching-moment distribution:
dMθ (r)nc c3
= −πρ r γθk (r)q̇θk (12.100)
dr 8
S31ij q̈wi + A4ji q̇wi + S38ij qwi + S78j (q̈θj + ωθ2j qθj ) + A5jm q̇θm − A6jm qθm = 0
(12.102)
where
! R
S10i = m γw2i (r) dr (12.103a)
0
! R
S31ij = m yCG γwi (r)γθj (r) dr (12.103b)
0
! R
S38ij = m ryCG γw i (r)γθj (r) dr (12.103c)
0
! R
S78j = m (kθ2 + yCG
2
)γθ2j (r) dr (12.103d)
0
where ωwi and ωθj are, respectively, the natural frequencies in the ith flatwise
bending and jth torsion modes (at rotor speed). The aerodynamic related integrals
are defined as follows:
! BR
ρ
A1im = clα C̄(k) crγwi (r)γwm (r) dr (12.104a)
2 0
! BR 2 c
ρ c
A2ij = π + clα cC̄(k) − yAC γwi (r)γθj (r) dr (12.104b)
2 0 2 2
! BR
ρ
A3ij = clα C̄(k) cr 2 γwi (r)γθj (r) dr (12.104c)
2 0
! BR
ρ
A4ij = clα C̄(k) cyAC rγwi (r)γθj (r) dr (12.104d)
2 0
Effective c.g. offset. For the purpose of examining the results, it is convenient
to define an effective c.g. offset (relative to the ith flatwise mode and the jth torsion
mode):
S31ij
( yCG )eff = (12.105)
S10i S78j
a0 λ4 + a1 λ3 + a2 λ2 + a3 λ + a4 = 0 (12.107)
and both real and imaginary parts of the characteristic equation must equal zero
(as is discussed in an earlier section). The two parts of the characteristic equation,
R (iω) and I (iω), each define two real-valued equations in the unknown variable
Z(=ω2 ). If some single parameter (such as the torsion mode frequency ωθ ) is
simultaneously varied in both of these parts of the characteristic equation and
the two equations are solved for Z, then neutral stability points (i.e., the stability
boundary values of ω2 ) will be the intersection points of the two curves.
where the [ZB ] matrix would contain all of the flutter equation terms excluding the
unsteady airloads {FA }. The unsteady airloads description is typically proportional
to some definition of dynamic pressure and a function of (reduced) frequency, and
in the case of the Loewy function to harmonic number m. The [QA ] matrix can
be made as rigorous as desired. In the case of rotor blades, although there will
generally be a spanwise variation of freestream velocity, thereby resulting in a
commensurate spanwise variation in reduced frequency, this need not pose any
inherent limitation to a rigorous evaluation of the matrix. With the variation in
spanwise freestream velocity, the dynamic pressure descriptor in Eq. (12.109)
would be based on the tip speed R.
Following the development of Chapter 9, Eq. (12.109) can be written as
Then, upon relating {X1 } to {X2 }, using the techniques described in Chapter 9 for
this method, the final useful form is obtained:
- .
(ω) {X} = [ZB (ω)]−1 21 ρ(R)2 [QA (k(r, ω), m, . . .)] {X} (12.112)
The analysis for flutter then becomes the determination of the locus of characteristic
multipliers as the frequency ω is varied. If any of the loci cross the real axis at
a value greater than unity, the system is unstable.
with divergence and flutter, with the two major dynamic parameters: (nondimen-
sional) blade torsion mode frequency ωθ / and c.g. offset yCG . Note that for a
given range of aft (–) c.g. offsets, two values of torsion frequency typically exist
for which the blade is neutrally stable in flutter.
2) Use of the simplified lift-deficiency function C(k) = 1.0 will generally pro-
duce conservative results. However, this generality should prove to be unreliable
when the flutter frequency is integral harmonic.
3) Based on the use of the force-phasing matrix technique, it can be shown that
for both the flutter and divergence instabilities similar drivers of the motion are at
work. The torsion degree of freedom is driven by inertia moments associated with
blade vertical acceleration ∼z̈ (flutter) or by centrifugal force and bending ∼2 z
(divergence), each taken together with the c.g. offset. The flatwise bending degree
of freedom is driven in each case by the aerodynamic lift generated by the torsion
deflection θ .
the development of methods for “prescribing” the distortion from a priori cal-
culations is currently an active area of methodology development. Despite the
differences in assumptions made in each of the various three-dimensional theories
regarding these four basic issues and in other procedural assumptions, underlying
similarities and unifying principles exist. These unifying principles are issues that
are addressed herein. Each of the various unsteady rotor aeroelastic theories devel-
oped assuming three dimensionality requires a basic understanding of one or more
of the following considerations.
2) Scalar formulation: Any one component of the velocity field can be expres-
sed in terms of the vortical strength of an elemental length of line vortex normal
to the component of velocity:
r sin κ
dq = ds = ds (12.117)
4π R3 4π R2
Fig. 12.24 Schematic of vortex structure for a lifting airfoil section in an unsteady
environment.
Trailing-wake vorticity:
∂b
[γtw ]T.E. = − (12.118)
∂r
Shed-wake vorticity:
1 ∂b
[γsw ]T.E. = − (12.119)
U ∂t
Blade-section circulation:
! T.E.
b = γb (r, ξ , t) dξ (12.120)
L.E.
Thus, the acceleration and velocity potential both satisfy the same differential
equation. The acceleration doublet is the appropriate elemental solution because
this solution element is directly proportional to an incremental pressure or lift. The
mathematical expression for such a doublet solution to the Laplace equation is as
follows:
p ∂ 1 p 1
ψ= = ∇ ·n (12.123)
4πρ ∂n R 4πρ R
where, as shown in the accompanying sketch, the n vector is normal to the surface
with the incremental pressure p:
= I(ψ) (12.124)
Pressure distribution:
Then
!!
v̄a (r, x) = p(η, ξ ) K(η, ξ , r, x, ω) dη dξ (12.127)
blade kernel function
surface
velocities flowing through the rotor disk. Consider the rotor advance and inflow
ratios:
V
µ= cos αR (12.128)
R
where αR is the rotor angle of attack. Rotor inflow λ is then obtainable from simple
momentum considerations:
CT
λ = −µ tan αR + (12.129)
2B2 λ2 + µ 2
The skew angle of the wake χ is then given by
∞
"
= [nc cos nψ + ns sin nψ] (12.131)
n=1
2) The vortex system consists of the bound vortices on the (infinite in number)
blades, the trailing tip vortices, and the spanwise constant shed vortices. This
vortex system forms a disk and an attached semi-infinite cylinder “filled” with
shed vorticity (see Fig. 12.28).
The pertinent details of the derivation are based on applications of the Biot and
Savart equation and the geometry shown in Fig. 12.29.
The development of the theory follows in the classical approach outlined ear-
lier wherein the air velocities induced at the blade sections (the known boundary
conditions) are formulated as functions of the (unknown) strengths of the assumed
system of vorticity.
Vertical air velocity induced by the trailing vortices is
t [ξ − r cos(ψ − φ)] ds
dvt = (12.132)
4π [ξ − 2rξ cos (ψ − φ) + r 2 + h2 ]3/2
2
where the differential arc length ds and vertical spacing of the trailing vorticity h
are respectively given by
2π v0
ds = ξ dφ; h= (12.133)
QR
Vertical air velocity induced by the shed vortices is
s r sin(ψ − φ) ds
dvs = − (12.134)
4π [ξ − 2rξ cos(ψ − φ) + r 2 + h2 ]3/2
2
or, finally,
∞
1 QR " n
vt = [η nc cos nψ + ηn ns sin nψ] (12.139)
4π R 2πv0
n=1
Thus, the effect of the trailing vorticity on the induced velocity is seen to depend on
the span and becomes more concentrated at the blade tips with increasing harmonic
content.
Integrated effect of the shed vorticity. The effects of the shed vorticity must
be taken not only vertically down the wake and azimuthally, but radially as well:
! 1 ! 2π ! ∞
1 d d [η sin(ψ − φ)] dz dφ dη
vs =
4π R 0 0 0 dφ dz [η2 − 2ηη cos(ψ − φ) + η2 + z2 ]3/2
! 1 ! 2π
1 QR d η sin(ψ − φ) dφ dη
= (12.140)
4π R 2πv0 0 0 dφ [η − 2ηη cos(ψ − φ) + η2 ]
2
In this case the integration must be broken up into two parts to avoid the singularity
when η = η :
∞
1 QR "
vs = [nc cos nψ + ns sin nψ]
4π R 2π v0
n=1
! η (n−1) ! 1 #
η ηn
× n dη + n (n+1) dη (12.141)
0 η η η
= (2−ηn )
Therefore,
∞
1 QR "
vs = (2 − ηn )[nc cos nψ + ns sin nψ] (12.142)
4π R 2πv0
n=1
where
vn
Lnw = 21 ρU 2 × 2b × 2π × (12.145)
U
The results of the preceding derivation, together with the simple expression for the
total nth harmonic of lift Ln (=ρUn ), then yield, after considerable cancellation
of terms,
σπ
Lnw = Ln (12.146)
4|λ|
where σ is the rotor solidity (=2bQ/π R2 ) and λ is the (uniform) inflow ratio. Then
the nth harmonic of (total) lift can be written as
σπ
Ln = LnQS − Ln (12.147)
4|λ|
or, in the final most useful form,
Ln 1
= (12.148)
LnQS 1 + (σ π /4|λ|)
This result is thus seen to have the form of a lift-deficiency function, applicable
to harmonic airloads in hover. Note that this result duplicates a result obtained by
Loewy using an entirely different approach (two-dimensional vs three-dimensional
for the Miller theory). From a practical standpoint this result is highly useful for
determining the loss of cyclic (1P) damping in a rotor in a hover or vertical flight
regime.
Drag coefficient. Within the linear angle of attack, as described earlier, the
drag coefficient cd typically assumes a more or less constant value cd0 up until αs
is encountered. For angles of attack beyond αs , the drag increases markedly with
angle of attack up to a value of approximately 90 deg, the point at which the airfoil
is acting essentially as a plate flat to the wind. For further increases in angle of
attack, the drag coefficient then decreases (see Fig. 12.31).
Lift coefficient. The most striking characteristic of airfoil lift in the presence
of a positive angle-of-attack rate α̇ is that the airfoil stalls both at an angle of attack
and at a corresponding lift coefficient that are respectively higher than those for
the static case (see Fig. 12.33).
loading needed to align the lifting load normal to the freestream direction. When
an airfoil is pitching upward into the high-angle-of-attack stall region, this high
curvature flow region on the leading edge begins to separate and in doing so forms
a vortex near the leading edge. As the angular rate goes to zero and then to negative
values, the vortex detaches from the leading edge and is convected downstream
but near the upper surface of the airfoil. These events are graphically shown in
Fig. 12.36. The suction formerly associated with the (unstalled) leading edge now
becomes initially a mechanism for a concentrated source of increased lift that tra-
verses along the upper surface of the airfoil. The magnitude of the increase depends
on a variety of factors such as the strength of the vortex and its distance from the
upper surface. The streamwise movement of the vortex depends on the airfoil
shape and pitch rate. The relative distance between the vortex and the airfoil varies
according to the kinematics of the airfoil.
As the vortex leaves the leading edge and traverses the chordwise length of the
airfoil, the overlift mechanism caused by the vortex suction changes its moment
arm relative to the mean aerodynamic center point xAC so that the overlift vortex
suction produces an increasingly more negative (nose-down) pitching moment, that
is, dynamic moment stall. The moment coefficient during dynamic moment stall
is significantly greater in magnitude than that normally obtainable during static-
stall conditions. As the vortex leaves the trailing edge, its effect on the airfoil is
subsequently diminished so that the lift abruptly drops off, and dynamic lift stall
occurs. Coincidentally, the moment coefficient typically increases to values that
are more typical of a static-stall condition, and the airfoil then remains stalled until
the angle of attack drops sufficiently so that reattachment of the flow can occur.
The dynamic-stall phenomenon is strongly governed by the rate of change of the
angle of attack, and the higher this rate, the higher the strength of the vortex, and
the higher the degree of overlift and overmoment.
Fig. 12.36 Flow visualization of a NACA 0012 airfoil undergoing dynamic stall:
Re = 3 × 10−1 ; α0 = 10 deg, α = 10 deg, and k = 0.049 (courtesy of U.S. Army
Aeroflightdynamics Laboratory).
Typically, it is found that K(k, α0 , α), the work caused by moment hysteresis,
is positive. This means that the airfoil is capable of extracting energy (or work)
from the airstream. The extent to which the airfoil can experience positive work
(negative damping) is indicated by the reduction of the quantity Q(k, α0 , α) to
zero and then to negative values. This characteristic is shown qualitatively by
Fig. 12.37.
The actual range of negative damping thus depends principally on reduced
frequency and mean incidence angle. Carta and Ham present, for a NACA 0012
airfoil, the tabulation of an equivalently defined function. Figure 12.38, which
shows a two-dimensional damping surface plot of this function, is reproduced
from their findings.
An alternate and perhaps more useful method for describing the airfoil pitch
damping is to formulate an equivalent damping coefficient. Let the two-
dimensional aerodynamic pitch-damping coefficient be denoted as Cθ , where
1 1 kπ
Mθ = − ρU 2 c2 × Cθ × θ̇ = ρU 2 c2 × ϒ × × θ̇ (12.152)
2 2 2
and where, from potential flow theory, Cθ → kπ/2, as α0 → 0. As gleaned
from the results of Carta and Ham, the pitch-damping effectivity factor ϒ can
be plotted vs k, for varying values of the mean angle of attack, as shown in
Fig. 12.39.
Fig. 12.39 Equivalent damping characteristics for a NACA 0012 airfoil: M = 0.2,
α = 6 deg, and αstall = 13.5 deg (after Carta and Ham).
the assumption of simple sinusoidal motion is suspect. For analyses of the rotor
in the time domain, with the attendant nonlinearities, an analytical representation
of the dynamic-stall process is necessary. To this end, such a modeling should
adhere to the following guidelines:
1) The formulation should be expressible in the form of differential equations
that are solvable in the time domain.
2) The formulation should account for the various aspects of the physics as
outlined earlier. In particular, the modeling of the physics should account for the
delay in the stall with overlift characteristics, moment stall preceding lift stall, and
suitable reattachment to potential flow conditions.
3) Being a “dynamic” phenomenon, the formulation should contain function-
ality with respect to angle of attack α, angle of attack rate α̇, and a time-variable
parameter that gives a decay-like behavior with time. The aforementioned unsteady
decay parameter αW (see Section 12.3.2) satisfies this requirement and can be used
for this purpose.
4) The methodology should be “tuned” to fit existing measured (real-world)
dynamic-stalled airfoil data. This suggests a semi-analytical approach wherein
suitable physical phenomena are mathematically modeled and then calibrated
using experimental data. Such an approach is not uncommon in aerodynamic
theory; the use of a measured static lift-curve slope clα , in place of the theoretical
value, 2π , is just such a case in point. The use of a lift-curve slope implies a linear
functionality, and the use of a measured clα instead of 2π is just such a calibration
of a functionality deduced from both theory and experimentation.
5) The use of an assumed (nonlinear) functionality to model generically the
physics together with measured unsteady stall airfoil data implies a curve-fitting
procedure. Use of some form of nonlinear least-squares fitting procedure can be
used.
6) Although dynamic stall does exhibit unique characteristics over and apart
from those of static stall, it still exhibits many of the qualitative characteristics of
static stall. The static-stall functionality in a sense provides a nonlinear “signature”
for the airfoil, which is already available and which could be used to help define
the dynamic-stall functionality. To this end, the desired dynamic-stall functionality
could incorporate a distortion of the static characteristics.
7) In addition to dynamic-stall functionality aimed at distorting the static-stall
characteristics, other supplementary functionality is required to model the vortex
shedding phenomena as discussed earlier.
where αss , is the static-stall angle of attack. These parameters are then used to
define supplementary time-domain functionality of the unsteady coefficients:
Lift:
clU (t) = clS (α − α1 − α2 ) + a0l α1 + cl1 + cl2 (12.154)
Pitching moment:
Drag:
smt ; and 3) angle of attack for flow reattachment αRE , where the flow is taken to
be reattached when α̇ is negative and α is less than αRE .
The specific dynamic-stall modeling parameters in Eqs. (12.154–12.156) are
all taken to be rational functions of the three basic parameters [i.e., (α/αss ),
A ≡ bα̇/U, and αW ) using a variety of unknown constants that are evaluated (cali-
brated) using the experimental data. The selection of the rational functions is based
on engineering judgement and providing sufficient algebraic terms to enable the
mathematics to find the more important ones. For example, the following linear
and nonlinear terms have been selected with constants that must be evaluated:
α1 = (P1 A + P2 αW + P3 )αss (12.157)
2
α α
cl1 = Q1 A + Q2 αW + Q3 + Q4 (12.158)
αss αss
The constants P1 , P2 , . . . , Q4 are all determined using nonlinear least-squares
curve-fitting techniques with the experimental data. Examples of the ability of this
approach to fit the experimental hysteretic loops with a high degree of accuracy
are shown in Fig. 12.40, as extracted from the work of Bielawa and Gangwani
et al. Because the synthesized method is formulated in the time domain, it can
easily address nonsinusoidal angles of attack. Figure 12.41 shows the results of
applying the methodology to a ramp function in angle of attack. Although exam-
ples of applying the method to nonsinusoidal motion are sparse in the literature,
the results of Fig. 12.41 show the potential of the method to give accurate simula-
tions for conditions that were not actually used to quantify the parameters in the
method. Gangwani presents results of applying this method to four airfoils often
used in rotary wing applications. Although variation was found in the various
empirical parameters for the airfoils synthesized, the parameter orders of magni-
tude and signs were found to be consistent. Leishman presents a comprehensive
survey of the various methods that have been devised for modeling the dynamic
stall phenomenon.
Within any given Mach-number Reynolds-number test condition, the accuracy
of any one of the loops curve-fitted by this process is comparable to the accuracy of
any other loop because all of the loops are curve fitted at the same time. Generally,
the various constants falling out as a result of the nonlinear curve-fitting calculation
are therefore tabulated functions of Mach-number. This method has the advantage
that it is relatively easy to implement in a time-history form of solution for a
nonlinear aeroelastic formulation. The number of Mach-number variable constants
is relatively small compared with the high degree of accuracy obtainable for a
fairly complex aerodynamic phenomenon. The principal disadvantage is that, for
any new (dynamically) untested airfoil, the unsteady coefficients needed for the
analytical functionality are not available. The use of unsteady parameters from a
similar airfoil taken together with the static coefficient data could be used as an
approximation, but one of uncertain accuracy.
Fig. 12.41 Correlation of synthesized unsteady aerodynamic data with test data for
a ramp function in angle of attack (after Gangwani).
load phenomenon that afflicts both helicopters and propellers is that of stall flutter,
which is the major result of the unsteady stall phenomenon described earlier. The
phenomenon typically manifests itself in these two elements of rotorcraft by the
substantial resulting increases in blade torsional and pitch link loads. The important
characteristics of stall flutter are as follows:
1) The phenomenon essentially involves a single-degree-of-freedom oscillation
in one of the blade torsion/feathering modes.
2) In propellers the phenomenon principally occurs at the start of the air-
craft takeoff roll at the time when the propeller blades are at maximum angle of
attack because of the lack of inflow angle, which normally results from propeller
translation through the air.
3) In rotorcraft the phenomenon principally occurs in main rotor blades only
for periodically recurring periods of time when the blade is in the retreating blade
(third and fourth) azimuthal quadrants. The requirement of the rotor to maintain
roll trim in high-speed flight necessitates that the blades operate with high angles
of attack in the retreating blade quadrants (see Fig. 12.42).
4) Because the conditions for the phenomenon are typically of short duration,
the oscillations are normally divergent only for short periods of time, and, hence,
the oscillations are not usually destructive in a catastrophic sense. The oscillation
generally assumes the character of a complex multiharmonic limit-cycle oscillation
as depicted in Fig. 12.43.
5) For rotorcraft the occurrence of stall flutter is strongly dependent on the
advance ratio µ and the blade loading, as measured by CT /σ (see Fig. 12.44).
From the results of numerous experimental studies, the source of the self-excited
oscillations has been identified to be the substantial loss of pitching-moment
damping because of the hysteretic nature of dynamic stall. An understanding of why
a stalled rotor blade should lose pitch damping is given in the preceding section.
As can be appreciated by the mathematical strategems required to synthesize the
Fig. 12.42 Typical angle-of-attack distribution over the roter disk for a helicopter in
high-speed flight—140 kn with nonuniform inflow; µ = 0.33.
Fig. 12.43 Typical time history of a blade pitch link showing the occurrence of stall
flutter in a helicopter.
Fig. 12.44 Stall flutter limited rotor lift capability in forward flight based on full-scale
wind-tunnel and flight tests.
where
! 1
S78j = R γθ2j (x)m (ky210 + kz210 ) dx (12.160)
0
Although expressible in this relatively compact equation, the stability solution for
stall flutter is generally complicated by the complexity of the function F (θ , θ̇ , . . .).
Because solution of this equation represents an exercise in nonlinear analysis,
further consideration of stall flutter is deferred until the next chapter.
References
Section
12.2 Bisplinghoff, R. L., Ashley, H., and Halfman, R. L., Aeroelasticity,
Addison–Wesley, Cambridge, MA, 1955.
Fung, Y. C., An Introduction to the Theory of Aeroelasticity, Wiley,
New York, 1955.
References (continued)
Kemp, N. H., “On the Lift and Circulation of Airfoils in Some
Unsteady-Flow Problems,” Journal of the Aeronautical Sciences,
Vol. 19, No. 10, 1952, pp. 713, 714.
Loewy, R. G., “A Two-Dimensional Approximation to the Unsteady
Aerodynamics of Rotary Wings,” Journal of the Aeronautical
Sciences, Vol. 24, No. 2, 1957, pp. 81–92.
Luke, Y. L., and Dengler, M. A., “Tables of the Theodorsen Circula-
tion Function for Generalized Motion,” Journal of the Aeronautical
Sciences, Vol. 18, No. 7, 1951, pp. 478–483.
Sears, W. R., “Some Aspects of Non-Stationary Airfoil Theory and Its
Practical Application,” Journal of the Aeronautical Sciences, Vol. 8,
No. 3, 1941, pp. 104–108.
12.3 Beddoes, T. S., “A Synthesis of Unsteady Aerodynamic
Effects Including Stall Hysteresis,” Vertica, Vol. 1, 1976,
pp. 113–123.
Bielawa, R. L., Johnson, S. A., Chi, R. M., and Gangwani, S. T.,
“Aeroelastic Analysis for Propellers,” NASA CR-3729, 1983.
Bisplinghoff, R. L., Ashley, H., and Halfman, R. L., Aeroelasticity,
Addison–Wesley, Cambridge, MA, 1955.
Edwards, J. V., Ashley, H., and Breakwell, J. B. “Unsteady Aero-
dynamic Modeling for Arbitrary Motions,” Proceedings of the AIAA
Dynamic Specialist Conference, Paper 77-451, 1977.
Greenberg, J. M., “Airfoil in Sinusoidal Motion in a Pulsating Stream,”
NACA TN-1326, 1947.
Isaacs, R., “Airfoil Theory for Flows of Variable Velocity,” Journal of
the Aeronautical Sciences, Vol. 12, No. 1, 1945, pp. 113–117.
Isaacs, R., “Airfoil Theory for Rotary Wing Aircraft,” Journal of the
Aeronautical Sciences, Vol. 13, No. 4, 1946, pp. 218–220.
Jones, R., Flannelly, W. G., Nagy, E. J., and Fabunmi, J. A., “Exper-
imental Verification of Force Determination and Ground Flying on
a Full-Scale Helicopter,” U. S. Army Applied Technology Lab.,
USAAVRADCOM-TR-81-D-11, Fort Eustis, VA, May 1981.
Jones, R. T., “Operational Treatment of the Non-Uniform Lift Theory
in Airplane Dynamics,” NACA Technical Note 667, 1938.
Jordan, P. F., “Aerodynamic Flutter Coefficients for Subsonic, Sonic
and Supersonic Flow (Linear Two-Dimensional Theory),” Royal
Aircraft Establishment Reports and Memoranda No. 2932, London,
April 1953.
Jose, A. I., Leishman, J. G., and Baeder, J. D.,“On Calculating
Unsteady Airloads with Time-Varying Free-Stream Mach Num-
bers,” Proceedings of the 61st Annual National Forum of the
American Helicopter Society, 2005.
References (continued)
Leishman, J. G., Principles of Helicopter Aerodynamics, rev. ed.,
Cambridge Univ. Press, New York, 2005.
McGuire, D. P., “The Application of Elastomeric Lead-Lag Dampers
to Helicopter Rotors,” Lord Library No. LL2133, 1976.
McGuire, E. P., “Fluidlastic Dampers and Isolators for Vibration
Control in Helicopters,” Lord Library No. LL-6502, 2001.
van der Wall, B., and Leishman, J. G., ”The Influence of Variable Flow
Velocity on Unsteady Airfoil Behavior,” Journal of the American
Helicopter Society, Vol. 39, No. 4, 1994, pp. 288–297.
12.4 Ham, H. D., “Helicopter Blade Flutter,” AGARD Rept. 607 (revi-
sion of Pt. III, Chap. 10 of the AGARD Manual of Aeroelasticity),
1973.
Miller, R. H., and Ellis, C. W., “Blade Vibration and Flutter,” Jour-
nal of the American Helicopter Society, Vol. 1, No. 3, 1956,
pp. 19–38.
Rodden, W. P., and Johnson, E. H., “MSC/NASTRAN—Aeroelastic
Analysis,” Users Guide, Ver. 68, MacNeal-Schwendler Corp.,
Los Angeles, CA, 1994.
12.6 Beddoes, T. S., “Onset of Leading Edge Separation Effects Under
Dynamic Conditions and Low Mach Number,” Proceedings of the
34th Annual National Forum of the American Helicopter Society,
May 1978.
Carlson, R. G., Blackwell, R. H., Commerford, G. L., and Mirick, P. H.,
“Dynamic Stall Modeling and Correlation with Experimental Data
on Airfoils and Rotors,” Proceedings of Specialists’ Meeting on
Rotorcraft Dynamics, NASA Ames Research Center, 1974.
Carta, F. O., and Ham, N. D., “An Analysis of the Stall Flutter Insta-
bility of Helicopter Rotor Blades,” Proceedings of the 23rd Annual
National Forum of the American Helicopter Society, May 1967.
Dowell, E. H., Clark, R. C., Cox, D., Curtiss, H. C. Jr., Scanlan, R. H.,
Edwards, J. W., Hall, K. C., Peters, D. A., Scanlan, R., Simiu, E.,
Sisto, F., and Strganac, T. W., A Modern Course in Aeroelasticity,
Fourth Edition, Kluwer Academic Publishers, Norwell, MA, 2004.
Gangwani, S. T., “Prediction of Dynamic Stall and Unsteady Air-
loads for Rotor Blades,” Journal of the American Helicopter Society,
Vol. 27, No. 4, 1982, pp. 57–64.
Gangwani, S. T., “Synthesized Airfoil Data Method for Predicition of
Dynamic Stall and Unsteady Airloads,” NASA CR-3672, 1983.
Halfman, R. L., Johnson, H. C., and Haley, S. M., “Evaluation of High
Angle-of-Attack Aerodynamic—Derivative Data and Stall-Flutter
Prediction Techniques,” NACA TN-2533, 1951.
Problems
12.1 The pitching moment that a rotor can generate is directly limited by atten-
uations caused by unsteady aerodynamic considerations. Expressions for
the rotor pitching moment, including unsteady aerodynamic effects, can
be obtained using both the Loewy/Miller lift-deficiency function (for har-
monic airloads) and the air mass dynamics theory in the preceding chapter
(Section 11.5).
(a) Derive expressions for the ratio of rotor pitch damping (with unsteady
aerodynamic effects) to that using only the quasi-steady assumption
using i) the Loewy/Miller lift-deficiency function and ii) air mass
dynamics, respectively.
(b) Explain any differences obtained using each of the two unsteady
aerodynamic formulations considered in part a.
12.2 Consider the tail rotor of a helicopter in yawing (hovering) flight at sea level
(see Problem 4.1). In this maneuver the (two-bladed) tail rotor develops
once per rev (1P) flapping responses as a result of the gyroscopic excita-
tions. A properly designed tail rotor must account for these responses with
regard to determining an appropriate setting of the flapping limit stops. The
1P excitations created by the gyroscopics are equilibrated by the aerody-
namic loads, which are necessarily unsteady. The rotor and flight condition
specifications are as follows:
radius R = 3.25 ft; thrust T = 275 lb; tip speed R = 712 ft/s
chord c = 0.65 ft; tip loss B = 0.95; ρ = 0.002378 lb-s2 /ft4
tail rotor position aft of aircraft c.g. TR = 20.5 ft
λ + Vz
λ = CT / 2B2 (12.161)
R
principally from the airloads on the advancing and retreating sides, where
at the 0.75R spanwise location the tangential velocity UT is respectively
proportional to R(0.75 ± µ).]
12.4 The design of a high-speed (articulated) helicopter rotor blade with a radius
of 20 ft is to incorporate a swept tip of 35 deg over the outer 10% of the
radius. It is to have an aspect ratio (R/c) of 12 and use 11% thick airfoil
sections. The blade is to have a uniform spanwise mass distribution, be
chordwise balanced on the line of section quarter-chords (including the aft-
swept quarter-chord in the tip region of the blade), and weigh 100 lb. A flutter
analysis of the blade shows that, because of the swept tip, the blade will have
bending-torsion flutter susceptibility involving the first bending mode with
the first torsion mode. However, the torsion mode does not couple unstably
with the second bending mode. Devise a design modification incorporating
an incremental mass distribution m and appropriate chordwise location
y10CR , which will decouple the torsion mode from the first flatwise mode,
but leave the favorable (stable) coupling with the second flatwise mode
intact. [Note that the devised incremental mass distribution must be physi-
cally realistic and compatible with the established blade geometry (i.e., the
additional mass must “fit” within the selected airfoil section).] The first and
second bending and torsion mode shapes are, respectively, given by
13.1 Introduction
The aeroelasticity of rotating wings is an especially abundant source of non-
linear dynamics problems. In this branch of aeroelasticity, significant nonlinearities
can be found in each of the three basic types of forces traditionally defining the
subject matter (i.e., inertia, elastic, and aerodynamic). Nonlinearities arise in the
inertia loadings of rotor blades because of their freedom to flex with signifi-
cant amplitudes in a rotating coordinate frame in at least four modes of elastic
deformation: transverse (beam) bending in two directions, axial torsion, and axial
extension. The internal elastic forces in rotor blades are also rich in nonlinearities.
The relatively high aspect ratios of rotor blades taken together with the inherent
twist (both built-in and elastic) and the aforementioned flexing amplitudes define
a variety of elastic nonlinearities.
However, perhaps of all of the nonlinearities present in rotor aeroelastic pheno-
mena, those arising from aerodynamic sources are the least tractable. Principal
sources of aerodynamic nonlinearities are the dynamic stall effects treated in
the preceding chapter and those of transonic flow occurring on the advancing
blades in high-forward-flight conditions. Indeed, even the unsteady aerodynamic
phenomena not associated with dynamic stall and advancing blade compressibility
effects are subject to nonlinearities because blade motion in the air mass is not
infinitesimal.
Beyond these basic sources of nonlinearities, other sources exist as well.
An important one is that afforded by the (articulated) blade lead-lag dampers.
Although these dampers do provide a measure of “linear” damping, they are
still essentially hydraulic devices and are characterized by velocity squared and
saturation-type nonlinearities. The control systems of rotors are additional sources
of nonlinearities because of the kinematics of the swash plate and pitch control
push rods, as well as inherent nonlinearities (both intentional and necessary) in
the power boost and stability augmentation systems.
The study of nonlinear dynamics, generally, and nonlinear vibrations, specif-
ically, has evolved into an extensive field of technology. Much is now known
of nonlinear dynamics, and a variety of analysis techniques have been formu-
lated. Nonlinear dynamics, generally exhibit characteristics that are quite unique
to the point that they can be considered to be “hallmarks,” that is, they have
become virtually synonymous with nonlinear phenomena. The nonlinear dynamic
characteristics of rotors are quite consistent with these more general hallmark
529
where the [C] and [K] matrices and the {F} vector are periodic with period 2π :
The vector {FNL (ψ, X, Ẋ)} represents the system nonlinearities that might or might
not be expressed in simple analytic form.
Because the nonlinearities existing in rotors are not all explicit or analytically
defined, a variety of methods must be used for their analysis. Thus, in the fol-
lowing material the distinction between explicit and implicit nonlinearities will
be stressed. Three major methods for analyzing the nonlinear dynamics typical
of rotary-wing systems are presented in this chapter: 1) simple linearizations,
2) direct numerical solutions, and 3) quasi-linearization. Furthermore, in support
of the direct numerical solution methodology, a section is devoted to the numerical
extraction of stability descriptors from the resulting response time histories.
where the motion X(t) is taken to be simple harmonic with a frequency of ω and
an amplitude of X̄. The required expression for the equivalent damping can then
be easily written as
WNL
CE = (13.4)
X̄ 2 ωπ
The simplified analysis of stall flutter offered in the preceding chapter constitutes
a solution of this type.
In contrast to the equivalent damping representation given earlier, the equivalent
stiffness approximation to the nonlinearity cannot be predicated on the basis of
equivalence of energy dissipation. Thus, nonlinearities of this type should not have
dependencies on velocity. The techniques used to obtain an equivalent stiffness
approximation can best be formulated by moving on to the next topic.
which we denote the describing function. This quantity, which can be real or
complex, is generally evaluated by expanding the output force caused by a simple
harmonic input motion in the form of a Fourier series based on the period of the
assumed frequency of motion ω:
∞
fNL (X, Ẋ) = Re An e inωt
(13.6)
n=1
for
X = X0 + X̄eiωt (13.7)
where
2π/ω
1
An = fNL (X0 + X̄ eiωt , iωX̄ eiωt )einωt dt (13.8)
π 0
GN = A1 /X̄ (13.9)
The preceding formulations then allow one to extract from the analytic (or even
tabulated) representation of the nonlinearity an amplitude-dependent value for
an equivalent stiffness. For an equivalent stiffness KE the appropriate source of
stiffening would come from the real part of the describing function:
where
It can be verified that the imaginary part of the describing function is equivalent
to the calculation of energy dissipated per cycle; hence, this component of the
describing function also yields the value of the equivalent damping:
The great advantage of the describing function approach is that describing func-
tions can be found for nonlinearities that cannot be stated analytically (as is the
case with various nonlinear aerodynamic phenomena such as dynamic stall). The
disadvantage of the technique is that it is only valid in the frequency domain and
it discounts the effects of the higher harmonics on the fundamental.
not allow for extensive variations in oscillatory amplitude (only one amplitude,
6 deg, was used), these results are prima facie independent of amplitude. Thus, the
results inherently assume that the effective damping moment varies linearly with
pitch amplitude, as does the linear (virtual mass) damping moment. The validity
of this approximation is moot, but, in the absence of more substantial unsteady
airfoil data, it does not appear unreasonable, especially because helicopter stall
flutter limit-cycle instabilities have been observed to be of this order of magnitude.
The variation of the measured zero-angle-of-attack damping values with
reduced frequency ( for this airfoil and for a pitch amplitude of 6 deg), as normal-
ized by the theoretical Theodorsen value of π k/2, is also presented in Fig. 12.35.
This variation is presented as a reference and is indicated by the dashed line in
the figure. Note that the measured variation is significantly less than the theo-
retical values, but improves with reduced frequency. Because the test article that
generated these data was an airfoil section having an aspect ratio of only 1.35, it
is reasonable to expect that inherent three-dimensional end effects could reduce
the two-dimensional pitch damping that might otherwise be available. Also, the
recovery trend of the experimental two-dimensional damping values to the theoret-
ical ones with reduced frequency is consistent with findings of various researchers
to the effect that increasing unsteadiness (k → ∞) tends to reduce the relative
importance of spanwise cross talk and thereby makes the airfoil sections behave
more two-dimensionally.
Because the full Fourier decomposition of the unsteady moment coefficient is
not available, the static-moment coefficient characteristics (Fig. 13.1) must be used.
Figure 13.2 is constructed by using the preceding expression for KE , but as applied
to Fig. 13.1. This task represents an approximation of unknown accuracy, as taking
the imaginary component of the first harmonic of the Fourier decomposition would
truly represent that component of the unsteady moment in phase with displacement.
In the absence of such data and the fact that the equivalent stiffness calculation is
independent of hysteresis, the use of the static characteristics for the calculation
of the equivalent stiffness offers a reasonable approximation.
Fig. 13.1 Static pitching coefficient characteristics for a NACA 0012 airfoil: M = 0.2.
Fig. 13.2 Equivalent stiffness characteristics for a NACA 0012 airfoil: M = 0.2, and
αstall = 13.5 deg.
The determination of the stall flutter characteristics using Figs. 13.2 and 12.39,
then, is as follows:
1) Determine the linear elastomechanical dynamic characteristics in terms of
generalized inertia Iθ and blade torsion natural frequency ωθ .
2) Determine the linear pitch damping (as given by the virtual mass effects) for
a selected rotor azimuth angle ψ̂ (typically taken to be 3π/2):
1
πρR4 c 3 2
Cθ = R γθ (x)[x + µ sin ψ̂] dx (13.13)
8 µ R
4) For each mean angle of attack within the selected range, iterate on amplitude
of motion ᾱ and reduced frequency k to determine a combination that will give
a value of equivalent damping (= ϒ · Cθ , where ϒ is obtained from Fig. 12.35),
which is either zero or negative for each of the selected mean angles of attack
α0 . Specifically, for each iteration value of amplitude ᾱ calculate the effective
frequency of motion:
ωθ2 + Kθ (α0 , ᾱ)aero
νθ = (13.16)
Iθ
where
1 c 2
ρR3
(Kθ )aero = (R)2 Kcm (α0 , ᾱ) γθ (x) (x + µ sin ψ̂) dx (13.17)
2 µ R
and where Kcm (α0 , ᾱ) is obtained from Fig. 13.2. The effective reduced frequency
keff is then given by the following expression:
c
keff = νθ /[R(x̃ + µ sin ψ̂)] (13.18)
R
which, together with each of the selected values of α0 , define the values of the
parameter ϒ, as per Fig. 12.35. Stall flutter instability is indicated by a zero or
negative value for ϒ.
Note that the limits of integration for the various integrals given earlier are
from µ to 1. Because the stall flutter instability is confined to the reversed-flow
quadrants of the rotor disk, these limits define the blade spanwise extent that is still
operating in forward-flow conditions. The development already given constitutes,
at best, only a rough cut at a rotor blade stall flutter calculation. For a more rigorous
calculation of stall flutter, as well as providing solutions to a variety of complicated
nonlinear instability phenomena, recourse must be made to numerical transient
time-history solutions of the full nonlinear equations of motion. Techniques for
the numerical integration of these equations, which have been found to be well
suited to rotary-wing aeroelastic applications, are presented in the next section.
rapid stability boundary estimations. In a sense this scenario defines two phases:
physical modeling and mathematical modeling. Both of these phases are important
for a thorough understanding of the phenomenon at hand. The thrust of the present
section is in describing some of the basic techniques found useful in obtaining
transient solutions of the complete nonlinear equations of motion, given earlier as
Most practical solution methods currently operational use some form of numer-
ical integration in time (or alternatively, nondimensional time, ψ = t) with a
uniform step size dt (or dψ). In the following material time will be assumed to be
nondimensionalized as per this definition.
For methods of the second type, the equations of motion are used as part of the
actual numerical integration scheme, and the resulting matrix equation set is highly
method dependent. An example of this equation transformation is presented in the
following material.
These general formulas give rise to a “family” of methods all identifiable as Adams
methods:
a) First-order backward-differencing:
ẋk+1 = ẋk + ψ 23 ẍk − 21 ẍk−1 (13.22a)
xk+1 = xk + 1
2 ψ(ẋk+1 + ẋk ) (13.22b)
b) Second-order backward-differencing:
ẋk+1 = ẋk + ψ 12 23
ẍk − 43 ẍk−1 + 5
12 ẍk−2 (13.23a)
xk+1 = xk + ψ 12 5
ẋk+1 + 23 ẋk − 1
12 ẋk−1 (13.23b)
c) Third-order backward-differencing:
ẋk+1 = ẋk + ψ 2455
ẍk − 24
59
ẍk−1 + 37
24 ẍk−2 − 38 ẍk−3 (13.24a)
xk+1 = xk + ψ 24 9
ẋk+1 + 24
19
ẋk − 5
24 ẋk−1 + 1
24 ẋk−2 (13.24b)
The Adams method is a technique of the first basic type wherein the highest-
order differentiated terms are calculated from the equations of motion, and
lower-order terms are subsequently integrated using one of the integration schemes
already given. Note that the use of increasing orders of backward differencing does
not necessarily ensure more accuracy or stability. For many applications first-order
differencing is preferable by virtue of its relative simplicity.
A second example is the numerical integration scheme used in the Newmark
method. This scheme can be seen to be based on the assumption of linear
acceleration over each time interval:
where α and δ are parameters that can be determined in order to guarantee integra-
tion accuracy and stability. One set of values producing an unconditionally stable
integration, which has found wide application, is δ = 21 and α = 41 . The Newmark
method is a technique of the second basic type, wherein the equations of motion
are used in the integration algorithm itself, and the accelerations must be calculated
using differencing techniques. Typical usage of the Newmark method requires the
use of appropriate differencing equations to form the {˙}k+1 and {¨}k+1 quantities.
A third example is the harmonic acceleration method, wherein the accelera-
tion is assumed to be simple harmonic over each time interval, with a specified
characteristic frequency ω̄. This frequency can be thought of as an “integration”
frequency. For modal formulations the integration frequencies will take on dif-
ferent values for each (modal) degree of freedom and are typically taken to be
the natural frequencies of each of the respective modes. The resulting integration
algorithms are as follows:
cos ω̄ ψ − cos 2ω̄ ψ 1 − cos ω̄ ψ
ẋk+1 = ẋk + ẍk − ẍk−1 (13.26a)
ω̄ sin ω̄ ψ ω̄ sin ω̄ ψ
1 − cos ω̄ ψ
xk+1 = xk + (ẋk+1 + ẋk ) (13.26b)
ω̄ sin ω̄ ψ
This scheme can be seen to be a modified form of the first of the optional forms
of the Adams method (involving only the first backward difference) given earlier.
For the harmonic acceleration method the acceleration is assumed to follow a sine
wave, whereas for the first backward-difference Adams method the acceleration
inherently follows a parabola. Indeed, for ω̄ ψ → 0 the algorithms are equivalent.
The harmonic acceleration method has the ability to duplicate exactly the free
responses of the degrees of freedom oscillating at their respective intrinsic natural
frequencies. Thus, although calculated responses at other frequencies would be
less accurate, such responses would inherently be attenuated by the second-order
response characteristics of the modal degrees of freedom.
The fourth example is a variant of one of the versions of the Runge–Kutta
method. The version described herein is based on a fourth-order Runge–Kutta
scheme with Gill coefficients and relates to a first-order form of the differential
equations of motion. This method was described in Chapter 9 (Section 9.4.5), and
the basic equations are duplicated here for completeness:
which results in a first-order form of the differential equations. The actual solution
algorithm involves the use of the usual integration interval h = ψ, as well as
one-half of this interval h/2:
h 1
{Yk+1 } = {Yk } + {k1 } + 2 1 − √ {k2 }
6 2
1
+ 2 1 + √ {k3 } + {k4 } (13.28)
2
where the vectors {k1 }, {k2 }, {k3 }, and {k4 } are given by
∗
{k1 } = {Yk } = { f (ψk ,Y k )} (13.29a)
h h
{k2 } = f ψk + , Y k + k1 (13.29b)
2 2
h h 1 √
{k3 } = f ψk + , Y k + √ 1 − √ k1 − (1 − 2)k2 (13.29c)
2 2 2
h √
{k4 } = f ψk + h, Y k + √ −k2 + (1 + 2)k3 (13.29d)
2
One distinction of Runge–Kutta methods is that they are self-starting, that is,
solutions can be advanced in time ({ }k → { }k+1 ) without the need for knowledge
of the response in the past { }k−m .
3) Uncoupling of the responses: The basic (matrix) differential equation of
motion set typically contains a mass matrix [M] and a stiffness matrix [K], both
of which are generally well-populated and often unsymmetrical. In all of the
methods discussed earlier, at some point in the solution flow, comprising any
single time-integration step, a decoupling must be performed. The methods that
work explicitly with accelerations, such as the Adams, harmonic acceleration, and
Runge–Kutta methods, must first have the accelerations decoupled so that the
“present-time” accelerations {Ẍk } can be made available for use in the numerical-
integration schemes. Thus, simultaneous equation solutions must be performed
at each time step to uncouple the accelerations; this is indicated by the follow-
ing expression (note that, although a premultiplication by [M]−1 is indicated, the
proper operation should nonetheless be a simultaneous equation solution):
where
h
[D] = [M] + [C0 ] + ah2 [K0 ] (13.32)
2
[B] = 2[M] − (1 − 2α)h2 [K0 ] (13.33)
h
[F] = [M] − [C0 ] + ah2 [K0 ] (13.34)
2
Because the Newmark method is defined for constant linear dynamic matrices, the
M, C 0 , and K 0 matrices in the preceding expressions must be the components of
the respective total system matrices that are constant. Consequently, the excitation
vector { fk } must be defined as follows:
{ fk } = − [C(ψk )] − [C0 ] {Ẍk } − [K(ψk )] − [K0 ] {Ẍk }
+ {F(ψk )} + {FNL (ψk , X k , Ẋ k )} (13.35)
where C 0 and K 0 represent the constant components of the total damping and
stiffness matrices, respectively. From this development two difficulties arise when
the algorithm is applied to the general nonlinear problem. First, the constant parts
of the damping and stiffness matrices must be formed and extracted from the total
matrices (at each discrete time point). Second, the solution for {Xk+1 } requires
knowledge of { fk+1 }, which is not generally available within the time-step solution
flow. One way to solve this problem is to extrapolate from { fk } to { fk+1 } using
backward-differencing techniques.
4) Initial conditions and startups: The operations already described relate to
repetitive calculations made within the solution flow defined by an arbitrary kth
point in time. Inherent in these operations is the assumption that somehow the
solution has been going on for a sufficient period of time so that all required past
quantities { }k−m are available. However, for start-up calculations of the solution
such quantities would not generally be available. Therefore, the fourth basic opera-
tion needing exposition is the definition of adequate start-up techniques. It should
be stressed that the need for defining start-up techniques is limited to only the
schemes that require quantities in the past { }k−m . Thus, because the Runge–Kutta
method does not require such quantities, start-up calculations are not a consider-
ation for this method. However, the other methods clearly require some sort of
startup procedure.
Three methods can be defined for starting the solution. First, an integration
method that does not require previous time quantities, such as the Runge–Kutta
or some lower-order method, can be used to establish sufficient points in time as
might be required by the nominal technique. An example of such a lower-order
method is Euler’s method, which essentially uses a zeroth-order differencing on
the integration of the acceleration:
ẋ1 = ẋ0 + ψ ẍ0 ; x1 = x0 + 1
2 ψ(ẋ1 + ẋ0 ) (13.36)
The second method is to use appropriate backward-differencing techniques to
approximate the missing required previous time quantities:
ẋ−1 = ẋ0 − ψ ẍ0 ; x−1 = x0 − 1
2 ψ(ẋ−1 + ẋ0 ) (13.37)
A third method is to rewrite the algorithm in such a way as to utilize the ini-
tial conditions. This method is especially useful for application to the Newmark
method. Because the difference equation contains displacement terms for three
consecutive time intervals and none for velocity, it is possible to formulate an
alternate algorithm that utilizes the velocity information. For this purpose the
following formulas are offered:
where
[P] = [M] + (h/2)[C] − 21 − α h2 [K]
− 41 − α h3 [C][M]−1 [K] (13.39)
[Q] = h [M] − 41 − α h2 [C][M]−1 [C] (13.40)
[R] = h2 21 − α h2 [I] + 41 − α h[C][M]−1 (13.41)
Accuracy vs cost vs stability. In a sense the three attributes are not indepen-
dent of each other in that all of the attributes are directly related to the integration
step size h. Generally, the solution can be made as accurate as desired and/or the
algorithm might (or might not) be made stable (if unstable) by the selection of a
suitably small step size. Conversely, the cost of obtaining a numerical solution is
inversely related to the step size; smaller step sizes require more points in time
for the calculations and commensurately more computer CPU time (=$). Another
factor impacting on computer costs is the degree of complexity of the algorithm.
The need to make many time-history solution calculations would make simplicity
an obviously desirable characteristic.
Thus, the proper selection of an integration method (and, correspondingly, step
size) would entail making some sort of a trade-off study. The fact that a multiplicity
of methods is presently in use would indicate that several factors are at work in
making this trade-off study: those relating to technical issues (form of equations
used, type of excitation, etc.), as well as those of a more subjective nature (e.g.,
familiarity with any given method and desirability for compatibility with other
existing methods).
α̇b α̈b2
A≡ ; B≡ (13.42)
U U2
The second method considered (for which results are shown in the figure) is based
on the so-called time-delay method for modeling the unsteady effects of dynamic
stall. This method is based on the basic hypothesis that a maximum quasi-steady
angle of attack exists at which the pressure distribution and the boundary layer are
in equilibrium. During further increases 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. This
method is characterized by an heuristically derived nonanalytical computational
scheme that can only be modeled using a nonlinear time-history solution approach.
The stall flutter signature shows that over one rotor revolution the response
grows exponentially, but is then quenched in the first quadrant to grow once again
in the retreating blade quadrants, etc. This response pattern is highly periodic and,
as such, would have to be strictly classified as a harmonic response. However,
because locally the response is in an energy absorption mode of operation, it
should be considered to be an instability. Although Fig. 13.3 shows the calculated
responses using two different analytical time-domain simulations of the unsteady
stalled airfoil data, other more rigorous methods of simulating these data have been
developed and have been applied to both helicopter and propeller stall flutter. The
calculation for the response could use any relatively straightforward numerical-
integration scheme for the elastomechanics. One challenge in predicting stall flutter
is in calculating the highly azimuthally variable and inflow-dependent impressed
angle of attack. A significant portion of this angle of attack is as a result of the
inflow angle, which in turn, is a function of the variable inflow environment of the
rotor. This impressed angle of attack would be expected to function as a highly
sensitive triggering mechanism for the instability.
Fig. 13.3 Correlation of CH-53A helicopter blade stresses and push rod loads using
alternate nonlinear calculation schemes for dynamic stall (after Carlson et al.).
by the five open or closed square symbols on the figure. These five cases were
selected with blade pitch angle-rotor speed combinations that would appear to
result in both stable and unstable calculated responses. Two extreme cases were
selected to overlap partially (either in blade pitch angle or rotor speed) with one case
that was deemed to be a representative condition strongly associated with signifi-
cantly higher pitch angles. These two extreme cases were deemed to be probably
stable conditions. Additional cases were introduced to enhance the definition of
the apparent stall flutter boundary. The calculated results for the five conditions
are shown either with an open or closed symbol denoting stability or instability,
respectively. Where appropriate, the calculated 21 PTP torsion limit-cycle torsion
stresses are shown parenthetically.
These results, while showing the practical usefulness of such an approach to
analyzing nonlinear phenomena, can identify special considerations that must be
made in order to obtain reasonable solutions. In the course of performing the
calculations, it was noted that the degree of stall flutter response obtainable was
a strong function of the value selected for structural damping. With a sufficiently
high value of damping (0.02), the blade was unconditionally stable, and the stall
flutter condition could not be induced. The final value of damping used (0.008) was
selected on the basis that it produced stable limit-cycle oscillations that neither grew
nor diminished once the instability “locked in.” The more detailed a description of
the phenomenon that is used, the more detailed all of the aeroelastic descriptions
need to be. In the figure a pitch-angle disparity of approximately 6 deg is noted
in the apparent stall flutter boundary pitch angle between the experimental and
analytical results. Such a disparity can come from several sources and represents a
general lack of refinement in predicting either the stall characteristics of the airfoil
or the detailed inflow description and the resulting mean angles of attack of the
various blade sections.
Fig. 13.5 Pictorial description of “split” tip-path plane oscillation (after Paul).
the instability and not the unstable slope of the pitching-moment characteristic.
Second, blade mode coincidence with even or odd multiples of 1/2P was not
a necessary condition for instability. Blade mode detuning for odd multiples
of 1/2P would in some instances raise the threshold of the blade oscillation to
a higher Mach number.
The equation
X̄ f
[J] − R (13.46)
ω fI
are then used to iterate from initial guesses of {X̄} and ω, (i.e., {X̄}0 and ω0 , respec-
tively) to the correct values of {X̄} and ω. The resulting solution for these variables
generally defines a limit-cycle condition that might or might not be stable. This
is illustrated in the following example, wherein for a certain type of nonlinearity
the locus of limit-cycle amplitudes has been plotted vs a system parameter . An
instability is known to exist for values of this parameter between the values of
1 and 2 for a zero value of the amplitude {X̄}. For the parameters selected in
this example, it is found that in the region near 2 the solution for {X̄} is triple
valued. Here the range of limit-cycle oscillations increases beyond 2 to the point
A beyond which the system abruptly stabilizes to point B. For subsequent reduc-
tions in the parameter , to values below 2 , the system again becomes unstable,
and {X̄} grows until point C is reached, wherein the response abruptly jumps to
the limit-cycle locus at point D. Such behavior is a classic example of a hysteretic
nonlinear response characteristic (see Fig. 13.7).
Fig. 13.7 Locus of limit-cycle amplitudes for a nonlinear system with a hysteretic
response behavior.
knowledge of these frequencies can enable the method to ignore such frequencies
selectively beyond the FFT phase. Third, the method is relatively simple to
implement; FFT and DFA algorithms are readily available.
One basic disadvantage of the method is that it is relatively expensive. Typically,
the method would find use in low-frequency stability investigations, and several
rotor revs would be needed to obtain enough points to enable the FFT and DFA
calculations to be reasonably accurate. A second disadvantage is that it is not a
mathematically rigorous method, and some “art” is sometimes needed to obtain
credible stability predictions. Often the plot of n|An | vs time is not a straight
line, and the calculation is thereby somewhat ambiguous. This deficiency can
sometimes be overcome by “weighting” one end of the trace or another (or even
the middle portion). Third, as with the log-decrement method, the moving-block
method will only identify the lightly damped modes and is blind to the highly
stable modes.
One technique that facilitates obtaining a good smoothing of the An values
is to use a filtering window such as a Hamming window or a Hanning window.
The method has proven capable of identifying aliased frequencies for time histories
that were generated by dynamic simulations with a high Floquet-like behavior; this
fact should be used for filtering out aliased frequencies in the critical frequency
selection part of the implementation.
where
µk = eak (13.56)
From the second form of the preceding approximation, it can be shown that each
of the constants µ1 , . . . , µn is a solution of the following algebraic equation:
Each of the fk ordinates is known (they are the numerical values that are being
curve fitted), and the equation set can be solved for the nα values providing that
N ≥ 2n. If N = 2n, a direct solution arises; if N > 2n, then the solution would be
approximate in a least-squares sense. Up to this point the solution is concerned with
obtaining the n exponential functions, µk (= eakt ). To complete the calculation,
the Cn coefficients must be calculated; this can be accomplished using a stan-
dard least-squares calculation. One idiosyncrasy of the method is that it does not
handle nonviscous damping well (most likely because nonviscous damping does
not produce responses of a simple exponential form). However, one substantial
advantage of this method is that it will detect more highly damped modes than will
the moving-block method.
References
Section
13.1 Urabe, M., “Galerkin’s Procedure for Nonlinear Periodic Systems,”
Archives of Rational Mechanics and Analysis, Vol. 20, 1965,
pp. 120–152.
13.2 Bisplinghoff, R. L., Ashley, H., and Halfman, R. L., Aeroelasticity,
Addison–Wesley, Cambridge, MA, 1955.
Bathe, K.-J., Finite Element Procedures in Engineering Analysis,
Prentice–Hall, Englewood Cliffs, NJ, 1982.
Ralston, A., and Wilf, H. S. (eds.), Mathematical Methods for Digital
Computers, Wiley, New York, 1960.
Hildebrand, F. B., Introduction to Numerical Analysis, McGraw–Hill,
New York, 1956.
13.3 Paul, W. F., “A Self-Excited Rotor Blade Oscillation at High Sub-
sonic Mach Numbers,” Journal of the American Helicopter Society,
Vol. 14, No. 1, 1969, pp. 38–48.
13.4 Hildebrand, F. B., Methods of Applied Mathematics, Prentice–Hall,
New York, 1952.
13.5 Hammond, C. E., and Doggett, R., “Demonstration of Subcriti-
cal Damping by Moving-Block/Randomdec Applications,” NASA
SP-415, 1976, pp. 59–76.
Hildebrand, F. B., Introduction to Numerical Analysis, McGraw–Hill,
New York, 1956.
Ibrahim, S. R., and Mikulcik, E. C., “A Method for the Direct Identifi-
cation of Vibration Parameters from the Free Response,” Shock and
Vibration Bulletin, No. 47, Pt. 4, 1977, pp. 183–198.
Problems
13.1 Although the action of a viscous damper can be simply approximated as a
resisting force proportional to velocity (by a constant of proportionality A),
a more realistic modeling of this action should include “saturation” effects at
high velocities. Consider a rotor lead-lag damper whose (nonlinear) damper
force is modeled by
A
FNL = − tan−1 (k )
˙ (13.59)
k
where the saturation value is given by
Aπ
FNLsat = − (13.60)
2k
13.2 An often specified operational “flight condition” for helicopters is the abil-
ity to land and take on passengers while perched on a sloping terrain with
the rotor rotating (at design rotor speed, though not actually lifting). In such
a condition the rotor’s tip-path plane must be maintained horizontally. Con-
sequently, in this condition the rotor blades are flapping with amplitude β̄
(= slope angle of the terrain) relative to the shaft. Because there are no Cori-
olis accelerations in the tip-path plane, the nominal lead-lag motion in this
plane is zero. However, relative to the shaft axis (where the lead-lag dampers
are located and must function), there is nominal blade lead-lag motion
(Coriolis accelerations) as a result of the flapping. This nominal lead-lag
motion can be thought of as a Hook’s joint effect, which manifests itself as
a periodic variation in the rotor speed of any one of the rotor blades. The
definition of the effective (linear) blade lead-lag damper characteristics is
required for the ground-resonance calculations for this operating condition.
The effective variability of the rotor speed, as seen by the damper, and the
specific assumed lead-lag damper saturation characteristics defined earlier
in Problem 13.1 are to be included in this calculation.
(a) Show that the lead-lag motion as seen by the shaft axis lead-lag dampers
is given by
[1 − (1 − cos β) sin2 ψ]
˙4 = (˙ + ) − (13.61)
[1 − (1 − cos β) cos2 ψ]
where
β = β̄ cos ψ (13.62)
and ˙1 and ˙4 are the lead-lag motions measured in the tip-path plane
and shaft axis plane, respectively. [Hint: Consider the blade motion
with three consecutive coordinate transformations going from the 1
(rotating tip-path plane), 2 (nonrotating tip-path plane), 3 (nonrotating
shaft axis plane), and 4 (rotating shaft axis plane) coordinate systems.]
(b) Derive an expression (valid to fourth order in β̄) for the effective (rotor
period, time-averaged) linear lead-lag damping coefficient (small
values of ¯1 ) subject to the inclusions of the two aforementioned effects.
13.3 As developed in detail in the work of Bathe, the “stability” of any numeri-
cal integration algorithm can be determined from the eigenvalues of the
integration approximation operator matrix  defined for that algorithm.
With this matrix any vector of a suitable combination of dynamic quantities
(displacements, velocities, accelerations, etc.) at one time step {X}t can be
defined in terms of the same vector at the next time step {X}t+δt . For an
unforced system this matrix is defined by the following relationship:
13.4 A conventional rigid rotor blade (articulated only in flapping about a flapping
hinge) is operating at high thrust in hover. Its lifting airloads are assumed
to be quasi-steady and governed by a two-dimensional lift coefficient
approximated by
14.1 Introduction
Testing for rotorcraft aeroelastic and aeromechanical stability continues to be a
necessary tool for the successful development of new designs. Although aeroelastic
analyses have continued to develop with increased breadth and sophistication, they
are still not accurate enough to be used routinely as design tools for the analyses of
man-rated aircraft. Not only must these analyses be validated experimentally, but
instability issues inherent in new rotorcraft concepts must be identified as a guide
to new or continued aeroelastic methodology development. Furthermore, although
testing of a prototype at full scale for aeroelastic stability is still performed, it is
with the intention of providing proof-of-design validations rather than exploring
new stability issues or sizing the design parameters to preclude the occurrence of
any possible instability. Such testing at full scale is typically very cost prohibitive
and is best done only where absolutely necessary. Thus, multiple reasons exist for
testing rotorcraft for aeroelastic instability at model scale.
For such testing to have relevance to full-scale behavior, it must be performed
with the correct attention to the scaling of the appropriate parameters, both with
the rotor and the airframe. Such scaling is generally an exercise in knowing which
issues can be compromised and which cannot, for it is impossible to scale every-
thing at scale factors of anything less than unity. Construction of aeroelastic models
requires attention to issues of practical construction technique. Often a properly
scaled model can end up being literally impossible to fabricate for certain important
ranges of parameters. Although certain test practices continue to be standard, the
state of the art in test procedures is constantly growing, and new resources must be
integrated into the test engineer’s tool kit. As with any other type of testing, there
are certain test preparations, test procedures, and data-reduction techniques appro-
priate to rotorcraft aeroelastic stability testing that must be understood. All of these
issues are covered in this chapter, along with material on aeroelastic considerations
to be given to nominally nonaeroelastic testing.
and gravity forces (Hunt). Assuming complete geometric modeling, these inter-
actions can be stated mathematically in terms of the following nondimensional
parameters, which should be maintained invariant:
Frequency scaling:
E E
λf = = (14.1a)
m2 ρ(R)2
Lock number:
ρacR4
γ = (14.1b)
Ib
Advance ratio:
V
µ= (14.1c)
R
Froude number:
2 R
F= (14.1d)
g
Mach number:
R
M= (14.1e)
a∞
The frequency scaling parameter ensures that the blade has the correct natural
frequencies in bending in relationship to rotor frequency. The Lock number ensures
that the rotor has the correct aerodynamic damping and aerodynamic coupling char-
acteristics, and the advance ratio ensures that the scaling of forward-flight speed
is correct in relationship to rotor rotational speed. The Froude number ensures that
the gravity effects, in terms of gravity springs and the rotor thrust, are properly
scaled in relation to the other three basic forces, and the Mach number ensures
that the correct compressibility effects are encountered on the advancing blade
portions of the rotor disk. The Froude number is typically in the order of 500–700
and becomes increasingly important with rotor size. Because the Froude number is
relatively large compared with the other nondimensional parameters, strict scaling
of the gravitational terms can sometimes be relaxed if their effects (as they relate
to the phenomenon at hand) can be approximated.
2
bS48 (generalized ground-resonance
3 = ; coupling parameter) (14.2)
2Meff S49
2
bS12
4 = (14.3)
2Ieff S10
Noting that K̄r must be the same for both configurations, we can then rewrite the
expression to separate out the explicit stiffness rate for the second configuration:
2 2
heff Kp1 h eff
Kp2 = 22 2
+ K̄r 2
−1 (14.7)
heff 1 21 heff 1
Dimensionless parameters:
3/2
Reynolds number Re λV λ λp /λµ λV λ λ λ
√
Mach number M λV /λa λV λ 1
Froude number F λ2V /λg λ λ2V /λ 1 1/λ
Velocities √
Linear V λV λV √λ 1
Angular λV /λ (= λ ) λV /λ 1/ λ 1/λ
Structural characteristics
Density σ λσ 1 1 1
Elastic modulus E λσ λ2V λ2V λ 1
Forces
Aerodynamic A λp λ2V λ2
Elastic B λE λ2 λ2V λ2 3 λ2
λ
Inertial (radial & oscillatory) C λσ λ2 λ4
Gravitational W λσ λ3 λ3 λ3
Model responses
Displ. amplitude a λ /λ λ λ λ λ
TESTING FOR ROTOR AEROELASTIC STABILITY
3) The need to approximate the gravity springs in pitch and roll, together
with the concurrent need to approximate free-flight trim conditions, places con-
straints on both the elastic restraints and preloads about the gimbal. For some
combinations of required spring rates and preloads, special gadgetry can be either
impractical or too expensive.
An additional constraint on the design of the gimbal arrangement is the need for
providing some form of “snubbing” of the model pitch and roll amplitudes so that
unstable motion, when it occurs, is delimited and thereby kept to noncatastrophic
values.
But according to Table 14.1, the ratio of rotor lift coefficients must always be unity
for either type of scaling:
CLMS
=1 (14.11)
CLFS
This scaling of total rotor lift and rotor lift coefficients becomes difficult because
of the nonlinear relationships defining the inflow and consequently the local inflow
angle distributions. In this regard three points should be noted:
1) The inflow ratio λ typically varies as the rotor thrust coefficient and would
thus be invariant with Froude vs non-Froude (e.g., Mach) scaling.
2) By virtue of the aforementioned difference in scaling of the rotor lift for the
two types of scaling, for the same geometry (including blade pitch angles) the two
types of scaling will produce different inflow velocities because of differing values
of lift.
3) Only for Froude scaling does weight scale in the same ratio as does aero-
dynamic lift.
Hence, in general the unit scaling of (required) lift coefficients for non-Froude-
scaled rotors would be difficult to achieve for the same geometric pitch angles. This
feature would be especially important for model tests wherein both rotor/fuselage
interactions and blade pitch angle similarity are important.
An additional aspect of aerodynamic performance is the matching of (scaled)
hub moments to achieve trimmed flight. In the usual implementation of model
fuselage pitch and roll articulation, gimbals with rotational springs about the gim-
bal axes are used (see Fig. 14.1). These springs generally have fixed scaled rates
according to the requirements for invariance of the pitch and roll frequencies
relative to the rotor rotational frequency and usually have low spring rates. The
maintenance of scaled hub moments in forward-flight (wind-tunnel) conditions
requires that some trim moments must be equilibrated by moments about the gim-
bal axes. The dual constraints of low spring rates and transmissibility of variable
trim moments pose special problems in the design of the gimbal axes.
14.3.4 Damping
The sources of damping occurring in the full-scale helicopter in free flight are
either structural [complex modulus of elasticity E = (1 + ig)Enom ], aerodynamic,
and/or frictional. Table 14.1 shows that for either type of scaling the aerodynamic,
elastic, and inertial forces all scale by the same scale factors. Consequently, sources
of damping that are either of the (classical) structural and/or aerodynamic types
should, to first-order effects, scale properly. However, the damping caused by fric-
tional sources tends to scale inversely with the scale factor. This can be appreciated
by noting how this form of damping would scale based on simple ideas using the
results of Table 14.1. Assuming that in a rotational joint (or bearing) the shear
stress caused by friction is constant with scale, then the nondimensional damping
is governed by the following relationships:
−1
[(M/θ̇ )/Iω]MS −2 λ , Froude scaling
= λV (14.12)
[(M/θ̇ )/Iω]FS 1, Mach scaling
Thus, the use of bearings and/or constructions having frictional joints with Froude-
scaled models should be avoided or minimized as much as possible for an accurate
scaling of internal damping.
in objective and/or scope can exist between various rotor aeroelastic tests, the
required instrumentation generally consists of elements in each of the following
types.
Electrical and hydraulic shakers. Shakers are most often used for vibration
related testing of structures. However, they have a valid application to aeroelas-
tic stability testing by providing the means for measuring component impedance
from which total system stability can be calculated (using Nyquist techniques).
Furthermore, for some response applications shakers can be used as sources of
standard input, either sinusoidal or pulse. Such shakers consist of two basic parts,
the base and a plunger, with a sinusoidal linear relative motion of two parts of the
shaker whose amplitude and frequency are electronically (servo) controlled.
For tests wherein actual stable or unstable time-history traces are sought, a standard
excitation is truly required in order to produce measurable responses all with the
same initial energy level. This uniformity of initial conditions thereby ensures that
the use of the stability evaluator already discussed will produce consistent results.
For cases of this type of (transient) testing, one appropriate form of excitation
would be an impulse, suitably calibrated to some standard magnitude. Actually,
any form of standard excitation could be used, even sinusoidal wherein after steady
sinusoidal motion is established over some fixed period of time, the sinusoidal
excitation is then sharply terminated.
Another form of stability testing is to test each of the various subsystems sep-
arately in the frequency domain and then to use the multiple-degree-of-freedom
Nyquist criterion (see Section 9.3.3). In this form of testing, frequency-response
functions (either in the form of mobilities or impedances) are obtained. Thus,
with tests in the frequency domain the excitation must generally be some form
of sinusoidal force and/or moment. However, in such cases although the pri-
mary responses will be sinusoidal and thus not generally unstable with regard to
the phenomenon being tested, other instability issues can arise. In this case an
appropriate form of monitoring of the responses would be to perform online fast-
Fourier-transform (FFT) calculations to determine if there are harmonic responses
at frequencies other than the fundamental one being used for the basic har-
monic excitation. Such monitoring is well within the state of the art of current
digital-signaling-processing (DSP) techniques and can be accomplished using
high-speed (digital) implementations of FFT algorithms on microprocessors whose
architectures are optimized for DSP operations.
(motion) of the model using some form of active power; and 3) a source of intel-
ligence to activate the caging motion when appropriate. These three activities can
now be implemented with state-of-the-art electronics and real-time microproces-
sors. However, a detailed description of generic methods for such implementations
is highly situation dependent and beyond the scope of this text. Generally, as shown
in Fig. 14.3, these three activities can be implemented using any of the various
motion sensors identified earlier along with some kind of powered mechanical
Fig. 14.4 Folded Abbé inverting prism (after Stetson and Elkins).
special indexing on the two shafts with computer control to maintain the 2 to 1
rotation relationship. The coaxial alignment had to be maintained free of significant
vibration, which necessitated the use of an air bearing support. Figure 14.5 presents
a pictorial of the arrangement selected for imbedding the prism within the hollow
shaft motor. Apart from the special mechanical design considerations, a significant
effort was expended to “orchestrate” the components electronically. Figure 14.6
presents a simplified schematic view of the interconnective block diagram for the
system.
Although the system was originally used to investigate the aeroelastic char-
acteristics of turbojet engine bladed-disk compressor stages, the system has
Fig. 14.5 Installation of the prism within the hollow-shafted motor (after Stetson and
Elkins).
Fig. 14.6 Schematic representation of the system block diagram (after Stetson and
Elkins).
Fig. 14.7 Derotated image test: object speed = 5000 rpm (after Stetson and Elkins).
Fig. 14.8 Inclusion of pylon damping in a nonaeroelastic test facility to preclude the
occurrence of ground resonance.
would be attached through its load cell to a structural ground point and would
form the basis of a configuration that might experience classical ground resonance
instability even though the full-scale rotorcraft would be instability free. In such a
case sufficient care should be put into the design of the support structure (including
the stiffness characteristics of the load cell) to ensure a stable configuration. As
shown in Fig. 14.8, one method is to fabricate the support structure so that an ample
source of pylon damping is available between the load cell and the ground point.
In this manner the stabilizing loads of the damper would not introduce errors in
the measurement of rotor hub loads (except for the effects of hub motion itself on
the loads).
An underlying issue in all of these considerations is that some form of aeroelastic
analysis should precede the fabrication of the test articles and that, on the basis
of this analysis, appropriate features must be designed into these test articles to
ensure stability. Inasmuch as the stability must be guaranteed, the inclusion of the
stabilizing features should be conservative.
References
Section
14.1, 14.2 Bielawa, R. L., “Validation of a Method for Air Resonance Testing
of Helicopters at Model Scale Using Active Control of Pylon
Dynamic Characteristics,” Journal of the American Helicopter
Society, Vol. 34, No. 2, 1989, pp. 33–42.
Hunt, G. K., “Similarity Requirements for Aeroelastic Models of
Helicopter Rotors,” Aeronautical Research Council, London,
CP-1245, 1972.
Stewart, H. L., Pneumatics and Hydraulics, Audel, (Div. of Bobbs-
Merrill), Indianapolis, IN, 1966.
14.4 Stetson, K. A., and Elkins, J. N., “Optical System for Dynamic
Analysis of Rotating Structures,” Air Force Aero Propulsion
Laboratory Technical Report, AFAPL-TR-51, 1977.
Problems
14.1 A given full-scale rotor blade has the following spanwise (constant) section
property distributions:
Mass distribution:
Also,
14.2 Consider the full-scale rotor defined in Problem 14.1, wherein a scaled
model aeroelastic stability test is to be performed. This test is to determine
the effects of advance ratio µ as they impact on flap-lag stability, espe-
cially with regard to the periodicity of the aerodynamic coefficients (Floquet
theory effects). A test range of advance ratios of 0.4 to 0.6 is required. The
design and construction of the model rotor is to be, as in Problem 14.1,
scaled to a 9-ft diameter. However, the tests are to be performed in a wind
tunnel limited to a useful speed range of 150–240 ft/s.
Determine appropriately scaled bending stiffness (both flatwise and
edgewise) and mass distributions needed to accomplish these tests.
14.4 As discussed in Section 14.2.2, the proper scaling of the natural frequen-
cies of the pylon structure of a model requires knowledge of the implicit
rotational spring caused by the rotor as quantified by K̄r . Using material
from Appendix D, derive an expression for this spring rate for the hovering
flight condition.
14.5 Figure 14.4 defines a typical installation of a multiaxis load cell for mea-
suring rotor loads. Generally such load cells are strain-gauge instrumented
and therefore require a finite degree of compliance in order to achieve
accurately measurable strains. Successful designs of these load cells main-
tain linearity despite exhibiting considerable coupling between the degrees
of freedom. Thus, the hub load vector {F}, [= Fx , Fy , Fz , MxF , MyF T ]
can be related to the incremental hub displacement degree of freedom vec-
tor, {δXh }, [= δ x , δ y , δ z , , δθxF , δθyF T ] by means of a stiffness matrix
defined by the load cell compliance:
15.1 Introduction
The use of elastomeric materials in the control of vibration and in rotor-
stabilization devices is well established in current rotorcraft design practice. In
this chapter we examine the basic properties of elastomeric materials and identify
the critical issues relating their application to rotorcraft. The key component in
all elastomeric devices is the elastomer itself, some form of a rubbery material
that behaves very much like a structurally damped metal, only with significantly
reduced elastic modulus but with a correspondingly enhanced loss factor. The
material is typically bonded in thin layers alternately with equally thin layers of
metal (or composite) to form a device that can serve both as a limited rotation
bearing and/or as a source of damping. The majority of elastomeric devices used
in rotorcraft applications typically utilize the elastomeric material in the shearing
mode of deformation.
Fig. 15.1 Typical spherical elastomeric bearing for radial retention of the rotor blade
and provision for articulated motion (Reproduced courtesy of the Lord Corporation).
579
flexure in the flapwise direction and a source of lead-lag damping. Thus, this
use of elastomerics allows a considerably more compact and lower maintenance
rotor system. An additional advantage claimed for elastomeric devices is that their
failure modes are “graceful” and visually evident in the form of degradation of
the elastomeric material. As attractive as the use of elastomeric devices is, how-
ever, they do have downsides that have to be addressed, as discussed later in this
chapter.
Fig. 15.5 Hypothetical plots of modulus amplitude and loss-factor data vs reduced
frequency (after Jones).
Fig. 15.7 Bell 412 Damper complex spring rate characteristics for single-frequency
motion (after Felker et al.).
the dual simultaneous motion at this nominal frequency and at another frequency.
The normal operation of rotors will generate once-per-rev in-plane oscillations as
a result of drag airloads on the advancing side in forward flight. Thus, the lead-
lag dampers will of necessity undergo once-per-rev motion at the same time that
stabilization might be needed at the lower lead-lag natural frequency. In the case
of the Bell 412, the once-per-rev frequency is 5.4 Hz. Experimental data were
obtained and published by Felker et al. for variations in the dual-frequency motion.
For this case experimental data were obtained for the damper to be driven by a
displacement composed of two sinusoidal motions:
The resulting data and the analytical predictions presented indeed confirmed the
degradation of the 3.3-Hz damping in the presence of varying degrees of oscil-
latory motion at 5.4 Hz. The analytical approach used by Felker et al. was to
curve fit time-domain expressions for effective spring and damper coefficients
k and c, respectively, using the measured single-frequency data. The resulting
analytical expressions for both k and c were simple polynomials in the instan-
taneous displacement x. The final expressions for damper force then combines
these polynomial expressions together with appropriate signum functions on
x and ẋ.
Numerous researchers have also endeavored to devise an accurate modeling
of the damper characteristics in the time domain (e.g., Gandhi and Chopra,
Brackbill et al.) in order to address this problem. Two difficulties with this
approach, however are, first, the need for extensive detailed testing of prospective
dampers to provide sufficient data to be curve fitted to the nonlinear time-domain
modeling, and, second, the lack of universality in applying the modeling coef-
ficients so obtained to alternate damper configurations. An alternate approach is
presented herein that uses the usual single-frequency data, as is normally obtained
directly from the damper manufacturer. However, this method does not require
intermediary transformations of K and K into the time domain and is essentially
formulated in the frequency domain.
δ1 + δ2 δ2 ω2 − δ2 ω2 δ1 ω1 + δ2 ω2 ω1 − ω1
ω1 = ω1 + = + δ2 (15.5a)
δ1 + δ2 δ1 + δ2 δ1 + δ2 δ1 + δ2
δ1 + δ2 δ1 ω1 − δ1 ω1 δ1 ω1 + δ2 ω2 ω1 − ω2
ω2 = ω2 + = − δ1 (15.5b)
δ1 + δ2 δ1 + δ2 δ1 + δ2 δ1 + δ2
Equations (15.8) and (15.9) thus represent a sinusoidal function that is both ampli-
tude and phase modulated. These equations are then used in a single-frequency
formulation, recognizing that, with the basic assumption just presented, the single-
frequency amplitude, as expressed by Eq. (15.8), is the instantaneous amplitude
and the appropriate frequency is ω̄. Thus, the expression for the instantaneous
damper force is given by
Note that K and K are evaluated, as sinusoidally varying functions of time, using
the single-frequency data presented in Fig. 15.7, along with the instantaneous
amplitude δinst .
Following the procedure formulated by Felker et al., one then retrieves the
resulting values of K and K (appropriate to one of the frequencies) by using a
harmonic analysis of F(t). Thus, taking the stabilization pertinent frequency as the
“1” frequency, the resulting K and K values can be evaluated as follows:
2 1 T
K1 = F(t) cos ω1 t dt (15.12)
δ1 T 0
2 1 T
K1 = F(t) sin ω1 t dt (15.13)
δ1 T 0
Fig. 15.8 (a) Damper stiffness, (b) Damper damping. Comparisons of predicted
and measured Bell 412 damper characteristics at 3.3 Hz for dual-frequency motion
(experimental data from Felker et al.).
Figures 15.8a and 15.8b clearly show the degradations in both stiffness and
damping that motion at 3.3 Hz sustains when increasing amplitudes of 5.4 Hz are
present. The correlation of predicted values with experiment appears to be compa-
rable with what was achieved with the curve-fitted damper and spring coefficient
method of Felker et al.
Aside from the damper attenuation characteristics caused by dual-frequency
motion, the use of elastomeric dampers does introduce an additional stiffness
into the lead-lag dynamics. Even with only single-frequency motion the effective
stiffness and damping values decrease with damper amplitude (see Fig. 15.7).
Accordingly, the lead-lag dynamics must be examined for suitable frequency
placement and damping levels for worst-case scenarios involving both high initial
displacements and the degree of dual-frequency motion.
Fig. 15.9 Comparison of loss factors of typical Fluidlastic® and purely elastomeric
lead-lag dampers as functions of dynamic amplitude (after McGuire).
References
Section
15.3 Jones, R. M., Mechanics of Composite Materials, McGraw–Hill, New
York, 1975.
15.4 Brackbill, C. R., Lesieutre, G. A., Smith, E. C., and Ruhl, L. E.,
“Characterization and Modeling of the Low Strain Amplitude and
Frequency Dependent Behavior of Elastomeric Damper Materials,”
Journal of the American Helicopter Society, Vol. 45, No. 1, 2000,
pp. 34–43.
Felker, F. F., Lau, B. H., McLaughlin, S., and Johnson, W., “Non-
linear Behavior of an Elastomeric Lag Damper Undergoing Dual-
Frequency Motion and its Effect on Rotor Dynamics,” Journal of
the American Helicopter Society, Vol. 32, No. 4, 1987, pp. 45–53.
Gandhi, F., and Chopra, I., “Analysis of Bearingless Main Rotor Aeroe-
lasticity Using an Improved Time Domain Nonlinear Elastomeric
Damper Model,” Journal of the American Helicopter Society,
Vol. 41, No. 3, 1996, pp. 267–277.
McGuire, D. P., “The Application of Elastomeric Lead-Lag Dampers
to Helicopter Rotors,” Lord Library No. LL2133, 1976.
16.1 Introduction
As lifting devices go, rotary wings differ considerably from conventional fixed
wings. Basically, rotary wings are devoid of the various cutouts associated with
fixed wings and are principally designed to take the attendant tension loads and
to have favorable frequency placements. Beyond these issues, the analysis of
rotorcraft stability issues is seen to be highly dependent on the interactions of the
blade kinematics and the resulting coupled blade responses. At the basic level, the
calculation of blade stiffness and mass distributions is mandatory. Also, even for
blades constructed of conventional (isotropic) materials we need to know where
the various elastomechanical points of interest (e.g., mass center, shear center)
are located on the blade sections. Finally, with the introduction of composite
construction, additional couplings accruing from anisotropic properties become
available, which are either detrimental or to be exploited for aeroelastic response
enhancement. This chapter addresses all of these issues with a special emphasis
on the use of composite material construction. In most of what is presented, the
use of conventional (isotropic) materials can be considered as a special case of
composites.
With the current availability of quality finite element tools, the temptation to
restrict one’s efforts solely to their use is quite persuasive. However, blind use of
only one tool is often a prescription for a failed or, on a more positive note, a more
expensive design. The material we will examine serves two functions: first, it will
provide useful first cuts at any new blade design to identify the critical issues that
must be addressed. And second, it will provide the basis for reality checks on the
more comprehensive methods the designer will eventually apply.
Figure 16.1 identifies the basic issues addressed in this chapter. The underlined
quantities are those needed to define the elastic characteristics of the blade. In their
more basic form, these are the stiffness distributions. More general relationships
are possible, however, and will be investigated.
As a rule of thumb, the blade section needs to be mass balanced to some point
near the quarter-chord and therefore has some form of balance weight near the
section leading edge that can contribute to the section’s elastic properties. Lastly, a
common design practice is to use a very thin skin for the aft portion of the section,
which is then strengthened and stabilized with the use of filler material (often
honeycomb). The aft skin is capable of contributing to the bending stiffness, and
the honeycomb filler, while providing negligible bending stiffness, can provide
shear stiffness.
multiple-cell beams of arbitrary geometry and built-up prismatic bars. Figure 16.1
begins the examination by defining the forces (T , Vy , and Vz ) and moments (Mx ,
My , and Mz ), as well as the average section strains (u , γxy
o , γ o ) and curvatures, (φ ,
xz x
v , and w ) that define the section stiffness properties. Initially, we will neglect
warping effects.
With the exclusion of warping effects, Eq. (16.1) represents the most general
expression for the section stiffness properties. In the sections to follow, the var-
ious elements of the [Cij(6) ] matrix will be developed. Note that, as written, this
definition for the beam stiffness is the composite equivalent of an isotropic (metal)
Timoshenko beam. For an isotropic beam, most of the elements in this matrix
would be zero. Such is not generally the case for composite beams, however.
Then, upon setting Vy and Vz to zero the following reduced equation is obtained:
u
T
T
u
φ Mx
Mx −1
φx
x (4 × 4) (4 × 4)
= 11 −→ = 11
(16.3)
−w
M
M
−w
y y
v Mz Mz v
where the resulting (reduced) 4 × 4 stiffness matrix with only the significant
nonzero terms identified is
(4) (4)
C11 C12 0 0
.. (4)
.
(4)
C22
(4)
C23 C24
(4 × 4) −1
11 = Cij (4)
= (16.4)
···
(4)
C33 0
(4)
(symm.) ··· C44
16.3.2 Assumptions
All of the basic features of composites, as presented in Appendix E, can be
brought to bear in formulating a theory for single-cell beams. In particular, the
following assumptions are made:
1) The beam is composed of “thin-shell” elements, where everywhere along the
contour of the section s, the thickness of the beam wall approaches a very small
value compared with the beam’s outer dimensions, thereby defining structural
membranes.
Fig. 16.2 General idealized features of a closed single-cell thin wall beam.
2) The constitutive relations for the structural membrane stress resultants and
strains are given by
Nxx
A11 A12 A16 εxx
Nss = A22 A26 εss (16.5)
Nxs (sym.) A66 γxs
where the Nxx and εxx are the axial stress resultant and strain, respectively, Nss
and εss are the “hoop” stress resultant and strain, respectively, and Nxs and γxs
are the shear flow and shear strain along the contour s. For a laminate of N plies,
the various Aij stiffness coefficients in Eq. (16.5) are calculated by summing the
products of the plane off-axis stress stiffnesses Qij for each of the k plies together
with its thickness tk (as described in Appendix E):
N
(k)
Aij = Qij tk (i, j = 1, 2, 6) (16.6)
k=1
8) The hoop stress resultant is zero, enabling the elimination of the hoop strain:
Nxx K11 K12 εxx
= (16.8)
Nxs K21 K22 γxs
where
K11 y ds = 0 (16.10a)
K11 z ds = 0 (16.10b)
Principal axes. The y and z axes that uncouple the respective bending
deformations are defined as the principal elastic axes and are determined by the
condition
K11 yz ds = 0 (16.11)
follows:
C11 = K11 ds (extensional stiffness) (16.12a)
2Ae
C12 = K12 ds (extension—twist coupling stiffness) (16.12b)
c
C13 = C14 = 0 (choice of axis) (16.12c)
dy
C15 = K12 ds (extension—Y -shear coupling stiffness) (16.12d)
ds
dz
C16 = K12 ds (extension—Z-shear coupling stiffness) (16.12e)
ds
4Ae
C22 = K22 ds (torsional stiffness) (16.12f)
c
2Ae
C23 = K12 z ds (torsion—Y -bending coupling stiffness) (16.12g)
c
2Ae
C24 = − K12 y ds (torsion—Z-bending coupling stiffness) (16.12h)
c
2Ae dy
C25 = K22 ds (torsion—Y -shear coupling stiffness) (16.12i)
c ds
2Ae dz
C26 = K22 ds (torsion—Z-shear coupling stiffness) (16.12j)
c ds
C33 = K11 z2 ds (flatwise bending stiffness) (16.12k)
where
2Ae
= rn ds ds (16.13)
c
Ae is the enclosed area of the single closed-cell beam, and c is the circumference.
(i)
C11 = K11 si (16.14a)
i
(i)
C12 = − K12 ρi si (16.14b)
i
(i) (i)
C13 = K11 zmid si (16.14c)
i
(i) (i)
C14 = − K11 ymid si (16.14d)
i
(i) dy
C15 = − K12 si (16.14e)
ds i
i
(i) dz
C16 = − K12 si (16.14f)
ds i
i
(i)
C22 = K22 ρi2 si (16.14g)
i
(i) (i)
C23 = − K12 ρi zmid si (16.14h)
i
(i) (i)
C24 = K12 ρi ymid si (16.14i)
i
(i) dy
C25 = K22 ρi si (16.14j)
ds i
i
(i) dz
C26 = K22 ρi si (16.14k)
ds i
i
(i) (i)
C33 = K11 [zmid ]2 si (16.14l)
i
(i) (i) (i)
C34 = − K11 zmid ymid si (16.14m)
i
(i) (i) dy
C35 = − K12 zmid si (16.14n)
ds i
i
(i) (i) dz
C36 = − K12 zmid si (16.14o)
ds i
i
(i) (i)
C44 = K11 [ ymid ]2 si (16.14p)
i
Fig. 16.5 Beam deflection calculations as a result of combined lift and blade weight
(after Hodges et al.).
Fig. 16.6 Beam twist calculations as a result of applied torque at the blade tip (after
Hodges et al.).
also included the effects of warping. This subsection presents the additional for-
mulations need for including warpage. This additional formulation requires the
inclusion of an additional term to the expression for the axial deformation:
Fig. 16.7 Beam twist calculations caused by centrifugal force (after Hodges et al.).
where U(x) is the (average) extension of the section, βy and βz are the section
rotations, and ψ is the torsion-related warping function, defined as
s
2Ae
ψ= s− rn ds (16.16)
c 0
and where
dy dz
rn = − z −y (16.17)
ds ds
(Note: It can be readily verified that ψ is identically zero for circular sections.)
The warping function is included in the formulation by expanding the constitu-
tive equations to a 7 × 7 matrix representation, where the generalized strains are
now given by
The quantity Qw is defined as the generalized warping function and for all practical
cases is zero. One exemption would be the imposition of a constraint to keep
a section from warping. This could alternatively be achieved by setting φx to
zero. The additional Cij terms are then given by the following:
C17 = K11 ψ ds (16.20a)
2Ae
C27 = K12 ψ ds (16.20b)
c
C37 = K11 zψ ds (16.20c)
C47 = − K11 yψ ds (16.20d)
Elimination of the warping deformation terms from the reduced constitutive equa-
tions can still be achieved using the techniques defined by Eqs. (16.2) and (16.3),
wherein the matrix partitioning now includes the generalized warping function Qx
along with the horizontal and vertical shears Vy and Vz .
Fig. 16.8 (a) Loads carried by hexagonal honeycomb on the faces normal to the X3
axis, (b) single shell with unequal wall widths (after Gibson and Ashby).
where t and l are defined in Fig. 16.8. Gibson and Ashby present experimental
results showing good correlation with Eq. (16.21).
Fig. 16.9 Torsion model for the beam section depicting the two systems of shear flows
and details for the kth cell.
The starting point for the development is the relationship between the shear
strain in each wall and the shear-stress resultant (i.e., shear flow). For composite
walls with nonzero shear-extension coupling [i.e., K12 = 0, in Eq. (16.8)], the
shear flow can be rewritten by eliminating the axial strain from the constitutive
K122
β= (16.23)
K11 K22
To make the following development more universal, we rewrite Eqs. (16.22) and
(16.24) as
Rotation constraints. Figure 16.9b shows the combined shear flows on each
wall of the kth cell. The two different shear flows combine over each wall to produce
for each cell a rotational twist rate and a vertical shearing strain. As regard the cell
twist rates, they must all be compatible with that of each of the other cells and to
the total section:
The twist rate of each cell is based on the integration of its shear strain γxsk around
its contour sk . Well-established theory gives for the arbitrary kth cell the following
expression for this twist rate:
1
φk = γxsk ds (16.27)
2Ak ck
The collection of twist rates for all of the cells can then be put into a matrix form as
∗
φ1 q1 ∗
q̃1
φ2
q2 ∗
∗
q̃ 2
.
.
..
..
..
.
∗
φ
q
k−1 k−1 q̃ ∗
∗ ∗ ∗ ∗ ∗ ∗ ∗ k−1
φ = φk = S q + S̃ q̃ = S qk + S̃
q̃k∗
φ
∗
qk+1
k+1
∗
q̃
..
..
k+1
.
.
.
.
∗
.
φ
q
n−1
n−1
q̃ ∗
φn qn ∗ n−1
(16.29)
where
1
1
.
φ = .. φT = {U} φT (16.30)
1
1
S ∗ ∗
S12 0 0 ··· 0 ··· 0
11
S ∗ ∗ ∗ ··· 0
21 S22 S23 0 0
.. .. .. .. .. .. .. .. ..
. . . . . . . . .
∗
[S ] = ∗ ∗ ∗
0 0 0 Sk−1,k Sk,k Sk,k+1 0 0
..
.. .. .. .. .. .. .. .. ..
. . . . . . . . . .
0 ··· ··· ···0 ∗
Sn−1,n ∗
Sn,n
(16.31)
where it is understood that the evaluation of H is made using the properties for
the wall indicated by the line integral. The matrix for the secondary (translation
constraint) shear flows is given by
∗ ∗
S̃11 S̃12 0 0 0 ··· 0 0 0
∗ ∗ ∗
S̃21 S̃22 S̃23 0 0 ··· 0 0 0
. ..
. .. .. .. .. .. ..
∗
. . . . . . . .
S̃ = .. ..
∗ ∗ ∗ ∗ ∗ ∗
S̃k,1 S̃k,2 S̃k,3 · · · S̃k−1,k S̃k,k S̃k,k+1 . .
. .. .. .. .. .. .. .. ..
.. . .
. . . . . .
∗
S̃n,1 ∗
S̃n,2 ∗
S̃n,3 ··· ··· ∗
S̃n,n−2 ∗
S̃n,n−1
(16.35)
where
∗ 1 ds
S̃1,1 =− (16.36)
2A1 cLE H
∗ 1 ds ds ds
S̃k,k =− + + (16.37)
2Ak ckU H ck−1,k H ckL H
∗ 1 ds ds
S̃k,i =− + 1≤i ≤k−1 (16.38)
2Ak ckU H ckL H
∗ 1 ds
S̃k,k+1 =− (16.39)
2Ak ck,k+1 H
∗ 1 ds
S̃n,m =− 1≤m ≤n−1 (16.40)
2An cTE H
where k = 2, … , n − 1.
Translation constraints. Figure 16.10 depicts the translations that the cell
area centroids must undergo in order 1) to maintain the centroids on a straight
line that also passes through the shear center and 2) to keep the cell septa walls
contiguous.
From the geometry shown in Fig. 16.10, one can see that the (translational)
shear strain that each cell must undergo can be related to the twist rate and the
Fig. 16.10 Pictorial representation of the translational shearing strains of the cells
needed to maintain section shape and the cells properly contiguous to each other.
The z-direction shear strains are likewise related to the shear flows using the H
factor defined by Eq. (16.25). The following matrix equation defines the required
translational shear relationships:
y1 − ySC 0 0
..
.
y2 − ySC 0
φ = yk − ySC {U} φT
..
.
0 ..
.
0 0 yn − ySC
∗
q1 ∗
q̃1
q2 ∗
∗
q̃
..
2
.
.
.
∗
.
k−1
q
q̃ ∗
∗ ∗ ∗ ∗ ∗ ∗ k−1
= [T ] q + [T̃ ] q̃ = [T ] qk + [T̃ ∗] (16.42)
∗
q̃k∗
q
k=1
∗
q̃
..
k+1
.
.
.
∗
.
n−1
q
∗
∗ q̃n−1
qn
where
T ∗ ∗
T12 0 0 ··· 0 ··· 0
11
T ∗ ∗
T22 ∗
T23 0 0 ··· 0
21
.. .. .. .. .. .. .. .. ..
∗
. . . . . . . . .
T = ∗ ∗ ∗
0 0 0 Tk−1,k Tk,k Tk,k+1 0 0
..
.. .. .. .. .. .. .. .. ..
. . . . . . . . . .
0 ··· ··· ···0 ∗
Tn−1,n ∗
Tn,n
(16.43)
∗ ∗
T̃11 T̃12 0 0 0 ··· 0 0 0
∗ ∗ ∗
T̃21 T̃22 T̃23 0 0 ··· 0 0 0
.
. .. .. .. .. .. .. ..
∗
. . . . . . . .
T̃ = .. ..
∗ ∗ ∗ ∗ ∗ ∗
T̃k,1 T̃k,2 T̃k,3 · · · T̃k−1,k T̃k,k T̃k,k+1 . .
. .. .. .. .. .. .. .. ..
.. .
. . . . . . .
∗
T̃n,1 ∗
T̃n,2 ∗
T̃n,3 ··· ··· ∗
T̃n,n−2 ∗
T̃n,n−1
(16.44)
The various elements can then be brought together to form a matrix eigenvalue-like
statement of the problem:
q∗ q∗
| |
| |
∗ S̃ ∗ | − {U} − −∗ − − − −
S | −0− − | −0 − |− 0− ∗
− − − |
− − − − − −−
| q̃ + y SC | | q̃ =0
− − −
T ∗ | T̃ ∗ | − {yk }
0 | 0 | {U}
− −
−
φT φT
(16.53)
or, more compactly as
∗
q −1
| |
− − − ∗ ∗
S | −S̃ − −| −−−{U}
q̃∗ = −ySC − − − | |
−−
− − −
∗ ∗
T | T̃ | − {yk }
φT
q∗
| |
− − −
−0− − | −0 − |− 0− ∗
+ | | q̃
0 | 0 | {U} − − −
φT
w1
w2
0
.
= −ySC [V ]−1 − − − φT = −ySC {W } φT = −ySC .. φT
{U}
w2n−1
− − −
w2n
(16.54)
The shear center location can then be immediately calculated from the last row of
Eq. (16.54):
1
φT = −ySC w2n φT −→ ySC = − (16.55)
w2n
Once the shear center is found using Eq. (16.55), the solution for the torsional
stiffness proceeds from Eq. (16.54). We rewrite Eq. (16.54) so that the shear flows
Solution for the torsion stiffness. With the solution for the shear flows in
hand, the calculation of the torsion moment resisting the applied moment Mx
can be readily accomplished. The torsion moment per twist rate is obtained by
summing the individual torques within each cell and the torques accruing from
the secondary translational constraint shear flows. Because all of the assumed shear
flows are circular, the resisting moment can then be easily written as
q∗
Mx = 2 A Ã − − − = 2 A Ã Q∗ φT
--
--
(16.57)
q̃∗
where the n elements of the {A} vector contain the individual cell areas:
A1
A2
..
.
{A} = Ak (16.58)
..
.
A
n−1
An
The n − 1 elements of the à vector consist of the areas defined in Fig. 16.9a for
the secondary shear flows:
!n
1 Ai
!n
Ai
2
.
.
.
à = !n (16.59)
k Ai
.
.
.
!n
A
n−1 i
Equation (16.57) then defines the torsion stiffness C22 ∗ analogous to the term
similarly defined for the single-cell beam [Eq. (16.12f)]:
∗ Mx
C22 = (16.60)
φT
And the shear flows in the outer skin elements (both top and bottom) are given by
k
qk = qk∗ + q̃i∗ (16.62)
i=1
where both shear flows are proportional to the rate of twist, as per Eq. (16.56). We
next reorder the element shear flows as follows:
1 0 ···
------------- -------------
1 1 0 ···
I 1 1 1 0
n
.
..
" #
qk −−−−−−−−−−−− 1 −1− − · −1− −1
· ·− ∗
{Q} = = −1 Q
qk,k−1 1 0
0 −1
1 0
.. ..
0 . .
In−1
−1 1 0
−1 1
(16.63)
Note that here In refers to an n × n identity matrix. Within each element the
shear strain is then given by
where
------------- -------------
..
. $ %
1
0
K22 (1 − β) j
..
.
{λ} = − − − − − − − − − − − − − − − − − − − − − − −− {Q}
..
. $
%
1
0
K22 (1 − β) j,j−1
..
.
(16.65)
and where j = 1, . . . , n and ( j, j − 1) = (2, 1), . . . , (n, n − 1). The axial strain for
each element is then calculated with the realization that with only an applied
moment Mx , Nxx must be identically zero for all elements:
" #
K
{εxx }= − 12 λ φT
K11
-------------- --------------
..
.
$ %
K12
0
K11 K22 (1 − β) j
..
.
= − − − − − − − − − − − − − − − − − − − − − − − − − − − − {Q}φT
..
.
$ %
K12
0
K11 K22 (1 − β) j,j−1
. ..
(16.66)
This strain distribution can then be used to calculate the various section strains that
couple with the torsion moment.
Clearly, for determining the chordwise mass center location, all of the components
must be considered. For certain loading conditions some, but not necessarily all,
of these components can contribute to the stiffness characteristics. For bending
stiffness considerations, the leading-edge balance weight, aft skin, trailing-edge
stiffener, and the additional webs can all contribute and should be included. For
conditions relating to shear, the filler (if it is honeycomb) can also play, and its
contribution should be examined.
Mass properties. The calculations of the mass distribution and mass cen-
ter of the airfoil section are elementary exercises in statics. The mass densities
and cross-sectional areas of the various components ρj and Aj , respectively, and
(LE)
(from the geometry of the section) the component mass centers yCGj are typically
readily available quantities. From these quantities all of the pertinent mass-related
properties can be found:
Mass distribution:
m= ρj Aj (16.67)
j
Mass center:
(LE) 1 (LE)
yCG = yCGj ρj Aj (16.68)
j m
j
where (ky )j and (kz )j are, respectively, the radii of gyration of the component about
axes parallel to the y and z axes through its mass center:
Torsional inertia (about the section mass center):
(LE) (LE)
JM0 = IMY + IMZ − m(yCG )2 (16.71)
! (LE)
(LE) j yT (AT E)j
yNA = !j (16.72)
j (AT E)j
(LE)
where ATj , and [ yT ]j are, respectively, the area and centroid location of the jth
component, if it can take tension.
(Nonprincipal axis) flatwise bending stiffness:
EIyy = (EIy )j (16.73)
j
(LE)
EIzz = Ej [IAz − (yNA )2 AT ]j (16.74)
j
Product of inertia:
(LE)
EIyz = Ej z[y − yNA ] dA (16.75)
j j
&
2
EIyy + EIzz EIyy − EIzz
EIf = − 2
EIyz + (16.76)
2 2
&
2
EIyy + EIzz EIyy − EIzz
EIe = + 2 +
EIyz (16.77)
2 2
1 2EIyz
φ= tan−1 (16.78)
2 EIzz − EIyy
Shear center. Because other components other than just the spar can contribute
to the shear bearing capacity, the results of Section 16.4 can be applied in an
approximate manner. Figure 16.11 shows a possible approximation of the complete
blade section to a series of closed section cells.
Note that the shear bearing capacity of the honeycomb material has been approx-
imated as a series of thin-web septa, the number of which must be made on the
basis of the convergence of calculated values with variation of the assumed spac-
ing. The H value that would be needed for the approximate honeycomb septa
would be the product of the effective shear modulus G∗23 , as given by Eq. (16.21),
and the assumed spacing sHC .
∗
GJ = C22 + GJj (16.79)
Fig. 16.12 Typical effects of temperature and moisture concentration on tensile and
shear moduli (after Tsai and Hahn, data from Browning et al.).
NPG
[ · · · ]T = θi , (θi + 90◦ )s · · · T (16.81)
i=1
where all of the indices are for 1, 2, and 6. Ni and Mk are, respectively, the compo-
nents of stress resultant and moment per unit width, εjo are the average strains, and
km are the curvatures. The appropriate integrations over the thickness have already
been achieved with the definitions for the Aij , Bim , and Dkm coefficients. For sym-
metric laminates, the Bim terms are zero. However, Eq. (16.82) does not address
all of the terms required for Eq. (16.1) because, as developed in Appendix E, only
w displacements were considered. Also, the k2 and M2 curvature and moment,
respectively, relate to curvatures and moments on the sides of the beam. Addi-
tional development is therefore required for ascertaining the in-plane moment and
curvature characteristics. Also, the out-of-plane shear strain and stress resultant γxzo
and Vz , respectively, involve a careful consideration of interlaminate shear charac-
teristics, which are beyond the intended scope of the present text, but addressed by
Whitney. Furthermore, in practice the simplistic lay up shown in Fig. 16.14 would
be enhanced either by laminates normal to those shown in the figure and/or with
layers of spiral wound composite on the outside to give environmental protection
as well as a deterrent to interlaminate separation.
References
Section
16.3 Gibson, L. J., and Ashby, M. F., Cellular Solids—Structure and
Properties, 2nd ed. Cambridge Univ. Press, New York, 1997.
Hodges, R. V., Nixon, M. W., and Rehfield, L. W., “Comparison of
Composite Rotor Blade Models: A Coupled-Beam Analysis and an
MSC/NASTRAN Finite-Element Model,” NASA TM-89024, 1987.
Rehfield, L. W., “Design Analysis Methodology for Composite Rotor
Blades,” Proceedings of the Seventh DoD/NASA Conference on
Fibrous Composites in Structural Design, AFWAL-TR-85-3094,
U.S. Air Force, 1985, pp. V(a)-1–V(a)-144.
16.4 Rehfield, L. W., Atilgan, A. R., and Hodges, D. H., “Structural
Modeling for Multicell Composite Rotor Blades,” AIAA Paper
No. 88-2250, 1988.
16.6 Browning, C. E., Husman, G. E., and Whitney, J. M., “Moisture Effects
in Epoxy Matrix Composites,” Composite Masterials: Testing and
Design (Fourth conference), ASTM STP 617, American Society for
Testing and Materials, 1977, pp. 481–496.
Tsai, S. W., and Hahn, H. T., Introduction to Composite Materials,
Technomic Publishing, Lancaster, PA, 1980.
Whitney, J. M., Structural Analysis of Laminated Anisotropic Plates,
Technomic Publishing, Lancaster, PA, 1987.
Winckler, S. J., “Hygrothermally Curvature Stable Laminates with
Tension-Torsion Coupling,” Journal of the American Helicopter
Society, Vol. 30, No. 3, 1985, pp. 56–58.
As helicopter dynamics technology has advanced, so too has the size and scope
of the dynamist’s tool kit. Clear-cut structural-dynamics-related topics have been
addressed in the preceding chapters. However, in this chapter we examine topics
that are not strictly structural dynamics and/or aeroelasticity, but crossover into
and have their impact on these subjects:
Interactions of the rotor/drive system with the engine/fuel control system: The
requirements for higher-bandwidth engine control systems have introduced prob-
lems requiring a blending of those tools typically in the purview of the control
systems dynamist with those of the aeroelastician/structural dynamist.
Aeroelastic optimization: Optimization methodology has matured and is slowly
making its way into standard design methodology. The goal of optimization
methodology is to mechanize the design process, if even only partially, in order to
attack vibration and instability-related problems earlier in the design cycle.
Both of these topics are addressed in this chapter to at least a basic degree so that
the problem areas are well defined, and current solution methods are identified.
Where possible, additional reference material has been identified for more in-depth
study of these topics.
features are put into the context of the complete frequency spectrum of interest to
the helicopter dynamist in Fig. 17.1, as extracted from Kuczynski et al.
The design of modern helicopter engine/fuel control systems is driven by
two main considerations: maximizing control system responsiveness in the low-
frequency mode and stabilization of the higher-frequency drive system torsion
modes, which experience has shown to be potentially unstable. Although the
details of the low-frequency operations are important, we will focus on the
higher-frequency modes that impact on the dynamics of the drive train.
Control system. The up-front control system, which produces a fuel flow com-
mand in response to a throttle input, can be thought of as a classic regulator and can
typically be implemented in a variety of ways. At the heart of this subsystem would
be some form of servo-controlled flow valve, some form of transducer to convert
throttle position to a command signal, and a servo amplifier. The details of such
an arrangement is subject to the controller specifications and design assumptions
made by the control system designer. For present purposes it is sufficient to assume
that the flow valve could be represented by a first-order lag and that the remainder
of the controller could be assumed to be of the form of a standard position, integral,
derivative (PID) controller. Thus, let us denote the throttle position as X0 , the fuel
Fig. 17.2 Flow block diagram for the composite engine, drive, and control system.
flow valve command as X1 , and the commanded fuel flow as X2 . Then the Laplace
forms of these two portions of the control system could be written in the following
manner:
1 TD p
X1 = KP 1 + + X0 (17.1)
TI p 1 + TD p/N
KV
X2 = X1 (17.2)
1 + TV p
where KP is the proportional gain; TI is the integral time constant; and TD is the
derivative time constant. In the derivative portion of the controller, there is typically
a filter, with time constant TD /N, with N residing in a typical range of 3–10. For
the simple first lag KV and TV are the appropriate gain and time constants. Also,
as shown in Fig. 17.1 a notch filter is often included for system stabilization. The
characteristics and role of this filter are addressed in a subsequent section.
Engine/fuel metering system. The dynamics of the engine and its attendant
fuel control system are describable by means of the transfer function provided by
the engine manufacturer. In general, the engine/fuel control dynamic system is
excited by the pilot throttle control (both directly and indirectly through correlator
coupling with the collective pitch angle) and by the turbine speed feedback loop
(governor).
Drive system and rotor lead-lag dynamics. The drive system, as we have seen
earlier, consists of the engine power turbine spool, rotors and/or propellers, and
all interconnecting shafts, inertias, and gearing. Sources of damping that must
be considered include the turbine itself, all aerodynamic elements (rotors and
propellers), and, in the case of articulated rotors, the lead-lag dampers. Excitation
of the subsystem consists of the controls acting on the aerodynamic elements (blade
pitch angle, etc.) and the gas torque acting on the power turbine disk of the engine.
Fig. 17.3 Implementation of artificial damping in the torsion system through the
engine/fuel metering system.
Low-pass filter. The starting point for all filter design is the low-pass filter,
whose loss or attenuation specifications are shown in Fig. 17.4. As indicated in
Fig. 17.4, there is a range of frequencies from zero to ωp , where the loss A(ω) is
no greater than Ap , which defines the “pass” band. Similarly, there is a range of
frequencies from ωa to ∞ that defines an “attenuation” or “stop” band, wherein
the loss must be no less than Aa . The frequency band, between ωp and ωa defines
a transition band that would ideally be equal to zero, but for all practical filters
is finite and to be minimized. The transfer function that relates input to output,
the gain H(ω), is a complex-valued function that is the reciprocal of the loss {i.e.,
H(ω) = [A(ω)]−1 }.
Several basic low-pass filters have evolved for the low-pass filter: the
Butterworth, the Chebyshev, the Inverse Chebyshev, the Bessel, the Cauer, and the
elliptic. Of these, the Butterworth is the most popular design used in circuits. All
of these filters produce a low-pass filtering and can be defined mathematically with
transfer functions that are rational functions with numerators and denominators
that are each specific polynomials of the Laplace transform variable. The polyno-
mials for each of these transfer functions are variable in degree n, which thereby
defines the “order” of the filter. (See Baker for numerical values of the coeffi-
cients defining various orders of Buttersorth, Chebyshev, and Bessel filters.) Each
Fig. 17.5 (a) Typical loss and (b) amplitude response characteristics for the
Butterworth, Chebyshev, and elliptic low-pass filters.
these filter types. Each filter has been configured to have the same loss specifica-
tions and was selected because it can easily be configured to match the end of the
pass band condition (see Antoniou).
Satisfying the specification for the start of the attenuation band (i.e., ωa and
Aa ) is generally achieved by selecting an appropriate order of the filter. Antoniou
presents inequalities for determining the appropriate orders for these filters to meet
these conditions.
so that
and, therefore
eout p2 + ω02
= 2 (17.6)
ein p + 2ζD ω0 p + ω02
Fig. 17.7 Notch filter frequency responses a) amplitude, b) phase angle, normalized
with respect to the notch frequency ω0 .
that the sampling rate in the system is high enough to preclude aliasing problems
(see Chapter 8). We can assume here that a practical choice of sampling rate has
already been made for the digital control system and that adequate anti-aliasing
filters have been incorporated. The first problem in the conversion of an analog
filter to a digital one is the fact that the analog filter is mathematically defined
using the Laplace transform, which is not appropriate to a sampled time system.
For this case the z transform is the appropriate tool.
Some of the properties that make the z transform useful are given (without
proof) as follows:
1) It is a linear operation:
2) Translation:
3) Complex differentiation:
dF(z)
Z[nTf (nT )] = −Tz (17.11)
dz
4) Convolution:
∞
∞
Z f (kT )g(nT − kT ) = Z f (nT − kT )g(kT ) = F(z)G(z) (17.12)
k=−∞ k=−∞
Just as with Laplace transforms a set of standard z transforms exists for a selection
of typical functions (see Antoniou).
Digital filter elements. The most general description of a digital filter is that
it is recursive, which means that the response of the filter is a function of the
excitation as well as the response itself:
N
N
y(nT ) = ai x(nT − iT ) − bi y(nT − iT ) (17.13)
i=0 i=1
The nonrecursive filter is the special case of the recursive type, wherein the
response is not a function of itself. The value of N defines the order of the filter.
In the case of analog filters, the elements used to construct them include resisters,
capacitors, and inductors for passive systems and, for active filters, operational
amplifiers and selectable resistance and capacitance networks. In the case of digital
filters, there are only three (mathematically defined) elements: the unit delay, the
adder, and the multiplier. All of these are defined in terms of an input variable(s)
x(nT ) and an output variable y(nT ). The implementation of these can be performed
via a shift register and/or appropriate NAND or NOR gates. Mathematically, these
elements are defined as follows:
1) Unit delay:
y(nT ) = x(nT − T ) (17.14)
which can be generalized by means of the shift operator, E:
E −r x(nT ) = x(nT − rT ) (17.15)
The z transform of the unit delay is given as
Y (z) = z−1 X(z) (17.16)
2) Adder:
k
y(nT ) = xi (nT ) (17.17)
i=1
whose z transform is given by
k
Y (z) = Xi (z) (17.18)
i=1
3) Multiplier:
y(nT ) = mx(nT ) (17.19)
whose z transform is given by
Y (z) = mX(z) (17.20)
These basic elements are illustrated in Table 17.1. The transfer function relat-
ing x(nT ) to y(nT ) for the recursive filter [Eq. (17.13)] can be obtained by z
transforming this equation:
N
N
Zy(nT ) = ai z−i Zx(nT ) − bi z−i Zy(nT ) (17.21)
i=0 i=1
or
N
ai zN−i
Y (z) i=0
= H(z) = (17.22)
X(z)
N
zN + bi zN−i
i=1
T
y(nT ) − y(nT − T ) = [x(nT − T ) + x(nT )] (17.23)
2
which, upon applying the z transformation, yields the following transfer function:
Y (z) T z+1
HI (z) = = (17.24)
X(z) 2 z−1
2 z−1
s= (17.25)
T z+1
N
ai sN−i
i=0
HA (s) = (17.26)
N
sN + bi sN−i
i=1
Thus, the derived digital filter having this transfer function will have approxi-
mately the same time-domain response as the given analog filter. Discrepancies will
arise because of the choice of sampling rate and the effects of digital quantization
(effects of fixed digital word length).
where is the excitation frequency for the digital filter and ω is the corresponding
frequency for the analog filter. The mapping of jω on the imaginary axis to j on
the unit circle can be expressed with the help of Eq. (17.23) as
2 T
ω= tan (17.29)
T 2
For T sufficiently small (i.e., <0.3), the two frequencies are approximately
equal. However, when this term is not small, a warping of the frequencies will
result. That is, the filter performance of the analog and that of the derived digital
filter will be similar, but the frequency spectrum for the digital filter will be warped
relative to the analog filter at the higher frequencies. This can be seen by inverting
Eq. (17.29):
2 ωT
= tan−1 (17.30)
T 2
One method for overcoming this problem is by designing the analog filter to
have prewarps of the pass-band and stop-band edge frequencies ωi so that the
correct edge frequencies will occur, as specified, in the digital filter i . This is
accomplished with the use of Eq. (17.29):
2 i T
ωi = tan (17.31)
T 2
Filter realization. In the case of analog filter realization, the Laplace trans-
form transfer function can be directly used to define the appropriate differential
equations, which can then be implemented either passively or actively. In the case
of digital filters, the starting point must be the z-transformed transfer function,
which will result from the use of the bilinear transformation and prewarping, as
per Eqs. (17.27) and (17.31). The one difficulty is that the z-transformed transfer
function will generally be a rational function, with polynomials in z in both the
numerator and denominator. The final filter design must have a realization consist-
ing of the basic digital filter elements, that is, unit delays, adders, and multipliers.
As it turns out, there is no unique methodology for doing this (see Antoniou). What
is presented herein is a variant of the direct method. This method first decomposes
the transfer function into two simpler transfer functions representing, respectively,
the numerator and denominator. Thus,
Y (z) N(z) N(z)
H(z) = = = (17.32)
X(z) D(z) 1 + D (z)
where
N
N(z) = ai z−i (17.33)
i=0
and
N
D (z) = bi z−i (17.34)
i=1
where
and
This decomposition is illustrated in Fig. 17.8. Each of the simpler transfer functions
shown in Fig. 17.8 then represents a polynomial in (1/z). The second part of the
direct method is then to implement the realizations of each of the polynomials.
To this end, let us consider the numerator polynomial and rewrite Eq. (17.32) as
follows:
Fig. 17.8 Decomposition of H(z) into two simpler transfer functions representing the
numerator and denominator.
Fig. 17.9 Basic unit for realizing a polynomial transfer function (after Antoniou).
Fig. 17.10 One of several possible realizations of a polynomial transfer function N(z)
(after Antoniou).
Fig. 17.11 Realization of the complete transfer function H(z) (after Antoniou).
where
N
N1 (z) = ai z−i+1 = a1 + z−1 N2 (z)X(z) (17.39a)
i=1
N
N2 (z) = ai z−i+2 = · · · (17.39b)
i=2
The basic unit for realization, as defined by Eq. (17.38), is illustrated in Fig. 17.9.
The complete realization of the polynomial N(z) is inferred by Eqs. (17.39a)
and (17.39b), wherein the basic unit is cascaded. This is illustrated in Fig. 17.10.
The realization of the complete transfer function then first realizes the denominator
function −D (z) to realize the U2 (z) transfer function (using the technique outlined
previously) and then combines them, as shown in Fig. 17.11.
Figures 17.9–17.11 illustrate only one of several possible implementations
using unit delays, adders, and multipliers that can be programmed within a digital
environment to achieve a specified z transform for any given filter.
Fig. 17.12 Typical interfaces and functions of the FADEC computer (diagram
courtesy of Shawn Coyle).
including the principal features and practical benefits of the FADEC system. As
shown in Fig. 17.12, the FADEC is integrated into the engine/fuel control system
using a variety of sensors and performs several calculations regarding the meter-
ing of the fuel, as well as detection of imminent and actual engine failure and the
initiation of remedial action.
The principal benefit of the FADEC is that it makes the rotorcraft easier to fly,
as it relieves the pilot(s) from having to spend time monitoring the several engine
gauges. Equally important is that it keeps the wear and tear on the engine(s) to
a minimum and thereby reduces maintenance costs. FADEC systems are now
standard equipment on nearly all rotorcraft turbine engines and on at least one
helicopter reciprocating engine.
briefly on the optimization process itself and then show how it can be applied to
rotorcraft structural dynamics and aeroelasticity.
Fig. 17.13 Illustration of a hypothetical simple feasible design space and the
optimization process.
and provides the best value of the cost function. This defines the constrained opti-
mization (minimum) point. Generally, the constrained optimization process will
result in a solution wherein the design parameters reside somewhere on a constraint
boundary.
Some of the challenges in formulating a practical optimization program relate to
inherent numerical and discretization issues and to the need for trend information.
As relates to the hypothetical example presented in Fig. 17.12, depending on where
one starts in the design space, the problem is one of where to go to next and by how
much. To a large degree, this depends on knowing gradient information of both
the cost function and the constraints. Although the trend of information might be
available analytically, generally it is not, and numerical evaluations must be made.
Numerical evaluations of gradients (i.e., partial derivatives) will depend on the size
of the increments used for approximating the partial derivatives. Also, depending
on the starting point and the nonlinearity of the cost function and/or constraints,
the increments used by the optimization code for iteratively wending its way to the
optimization point can become erratic. Various techniques have been devised for
avoiding these erratic iterations and making the iteration converge more directly
to the optimization point. All of these issues have been addressed more or less by
several analysts. Several optimization resources are now available, and a few are
included in the list of references.
could be
3
J= (Qj2nc + Qj2ns )Wj where n = kb, k = 1, 2, . . . (17.40)
j=1
or
3
J= (|Qjnc | + |Qjns |)Wj again, where n = kb, k = 1, 2, . . . (17.41)
j=1
and where Wj is a set of arbitrary weighting factors that must be (user) selected
to reflect the priorities assigned to the various components. Ideally, one would
want to use some “bottom-line” description of the actual vibratory accelerations
at some critical point (or points) in the aircraft structure for the Fourier compo-
nents of Q. However, the ability to analyze the entire aircraft, including the effects
of rotor-fuselage coupling, would require a substantial formulation and computa-
tional chore. Furthermore, as we have seen, such a formulation is not presently
a reliably accurate basis. An alternate selection of quantity for the basis of the
cost function might be the various components of the hub shears and/or moments
H(x, y, z) and M(x, y, z) , respectively. For the vibration minimization problem several
appropriate cost functions could be defined, depending on the ability to accurately
formulate them quantitatively. An even simpler formulation might be to identify
and formulate appropriate dynamic magnification factors for a selection of blade
bending modes.
For the stabilization of aeroelastic phenomena, the cost function could take the
form of a combination of stability descriptors for any number of phenomena, again
with appropriate weighting functions. Thus, if we were to require that both ground
resonance and air resonance instabilities be stabilized or kept to stable levels, the
cost function could take the following form:
2
(GR)
(SC)
σm
J= Wm ζreq +
(SC) 2 (SC) 2
m [σ m ] + [ω ] m
GroundResonance
2
(AR)
(SC)
σn
+ Wm ζreq + (17.42)
(SC) 2
(SC) 2
n [σ n ] + [ω ] n
AirResonance
where ζreq would be a required damping ratio to be applied to all stability critical
modes and the various stability critical eigenvalue components are denoted as
(. . .) (SC) .
m ≤ pm ≤ pm
p(L) m = 1, 2, . . . , M
(U)
(17.43)
which specifies lower and upper bounds on each parameter’s range of variability
and
Gn ( pm ) ≥ 0 n = 1, 2, . . . , N (17.44)
which specifies that there shall be no interference between the design parameters.
which specifies the lower bounds on aeroelastic stability for all of the coupled
aeroelastic modes and
(L)
ωk2 ωk2 − 1 ≥ 0
(U) k = 1, 2, . . . , K (17.46)
1 − ωk2 ωk2 ≥ 0
which specifies the frequency placements within bounds and away from nearest
integral harmonics.
Iteration process. Beginning with some initial design (one whose design
parameters lie within the feasible design space, i.e., those that actually do satisfy the
constraints), the optimization process can then iterate to those design parameters
that minimize the cost function J. Typically, this minimization process ultimately
must be capable of generating and using “trend information” regarding both the
constraints and the cost function. This trend information takes the form of various
gradient or partial derivative matrices. These expressions can be calculated by any
of a number of methods available, as appropriate to the specific problem. The
choice of methodology for calculating these matrices, as well as using them to
effect iterative changes in the right direction to the optimum value, is subject to a
variety of schemes and implementations. A straightforward approach would be to
evaluate them using numerical differentiation (as is done in the CONMIN code).
There is a tradeoff to make, however, in that small increments needed for the
differentiation process are conducive to accuracy, but can lead to increased CPU
time to arrive at an optimization solution.
Fig. 17.14 Practical implementation scheme for rotorcraft design with aeroelastic
optimization.
References
Section
17.1 Alwang, J. R., and Skarvan, C. A., “Engine Control Stabilizing
Compensation - Testing & Optimization,” Journal of the American
Helicopter Society, Vol. 22, No. 3, 1977, pp. 13–18.
Antoniou, A., Digital Filters: Analysis and Design, McGraw–Hill
Book Company, New York, 1979.
Astrom, K. J., and Wittenmark, B., Computer Controlled Systems,
Theory and Design, (Information and System Sciences Series),
Prentice–Hall, Englewood Cliffs, NJ, 1984.
Baker, B. C., “Anti-Aliasing, Analog Filters for Data Acquisition
Systems,” Microchip Application Note AN699, 1999.
Corliss, L. D., “The Effects of Engine and Height-Control Character-
istics on Helicopter Handling Qualities,” Journal of the American
Helicopter Society, Vol. 28, No. 3, 1983, pp. 56–62.
Coyle, S., Cyclic & Collective: Second Edition of the Art and Sci-
ence of Flying Helicopters, Helobooks, a Division of Mojave Books
Limited, Mojave, CA, 2004.
Greensite, A. L., Elements of Modern Control Theory, Spartan Books,
New York, 1970.
Howlett, J. J., Morrison, T., and Zagranski, R. D., “Adaptive Fuel
Control for Helicopter Applications,” Journal of the American
Helicopter Society, Vol. 29, No. 4, 1984, pp. 43–54.
Johnson, D. E., Introduction to Filter Theory, Prentice–Hall Inc.,
Englewood Cliffs, NJ, 1976.
Kuczynski, W. A., Cooper, D. E., Twomey, W. J., and Howlett,
J. J., “The Influence of Engine/Fuel Control Design on Heli-
copter Dynamics and Handling Qualities,” Journal of the American
Helicopter Society, Vol. 25, No. 2, 1980, pp. 26–34.
17.2 Davis, M. W., and Weller, W. H., “Application of Design Optimization
Techniques to Rotor Dynamic Problems,” Journal of the American
Helicopter Society, Vol. 33, No. 3, 1988, pp. 42–50.
Kresselmeier, A., and Steinhauser, R., “Systematic Control Design
by Optimizing a Vector Performance Index,” Proceedings of IFAC
Symposioum on Computer Aided Design of Control Systems, 1979,
pp. 113–117.
References (continued)
Haftka, R. T., Gurdal, Z., and Kamat, M. P., Elements of Struc-
tural Optimization, Kluwer Academic Publishers, Dordrecht, the
Netherlands, 1990.
McCarthy, T. R., and Chattopadhyay, A., “A Coupled Rotor/Wing Opti-
mization Procedure for High Speed Tilt-Rotor Aircraft,” Journal of
the American Helicopter Society, Vol. 41, No. 4, 1996, pp. 360–369.
Vanderplaats, G. N., Numerical Optimization Techniques for Engi-
neering Design, with Applications, McGraw-Hill, New York,
1984.
Vanerplaats, G. N., “CONMIN - a Fortran Program for Constrained
Function Minimization, User’s Manual,” NASA TMX-62282, 1987.
Weller, W. H., and Davis, M. W., “Wind Tunnel Tests of Helicopter
Blade Designs Optimized for Minimum Vibration,” Journal of the
American Helicopter Society, Vol. 34, No. 3, 1989, pp. 40–50.
Young, D. K., and Tarzanin, R. J. (Jr.), “Structural Optimization and
Mach Scale Test Validation for a Low Vibration Rotor,” Proceed-
ings of the 47th Annual National Forum of the American Helicopter
Society, May, 1991.
obscurity. At that time, the instability was deemed to be improbable for the early
class of piston engine, propeller-driven aircraft then in existence. The instability
resurfaced in fact with the advent of turbine powered aircraft with high inertia
propellers. Identification of the instability was, of course, difficult with the break-
up of the stricken aircraft. It was the work of Houbolt and Reed in 1962 that
correctly inferred that this instability was the culprit and it was subsequently
reintroduced into the dynamics literature.
efficient solutions to the Floquet stability analysis problem as well as the devel-
opment of the multiple-degree-of-freedom frequency-domain stability analysis
methodology (as described in Chapter 9).
References
Section
18.1 Bellis, M., Inventors, The History of Computers, available online
at http://inventors.about.com/library/blcoindex.htm (cited June
2005).
Bellis, M., Inventors of the Modern Computer, FORTRAN—The First
Successful High Level Programming Language, available online
at http://inventors.about.com/library/weekly/aa072198.htm (cited
June 2005).
References (continued)
Coleman, R. P., and Feingold, A. M., “Theory of Self-Excited
Mechanical Oscillations of Helicopter Rotors with Hinged Blades,”
NACA Rept. 1351, 1958.
Cooley, J. W., and Tukey, J. W., “An Algorithm for the Machine
Calculation of Complex Fourier Series,” Mathematics of Compu-
tation, Vol. 19, 1965, pp. 297–301.
Hodges, D. H., and Dowell, E. H., “Nonlinear Equations of Motion
for the Elastic Bending and Torsion of Twisted Nonuniform Rotor
Blades,” NASA TN D-7818, 1974.
Houbolt, J. C., and Brooks, G. W., “Differential Equations of Motion
for Combined Flapwise Bending, Chordwise Bending and Torsion
of Twisted Non-uniform Rotor Blades,” NACA Rept. 1346, 1958.
Houbolt, J. C., and Reed, W. H., “Propeller-Nacelle Whirl Flutter,”
Journal of the Aeronautical Sciences, Vol. 29, No. 3, 1962, pp.
333–346.
Lawrence, C., and Kielb, R. E., “Nonlinear Displacement Analysis
of Advanced Propeller Structures Using NASTRAN,” NASA TM-
83737, 1984.
Lawrence, C., Aiello, R. A., Ernst, M. A., and McGee, O. G., “A
NASTRAN Primer for the Analysis of Rotating Flexible Blades,”
NASA TM-89861, 1987.
Smith, B. T. et al., Matrix Eigensystem Routines - EISPACK Guide,
2nd ed., Vol. 6, Lecture Notes in Computer Science, Springer–
Verlag, New York, 1976.
Stewart, C. D., Finite Element Analysis/Method Resources, available
online at http://www.steelynx.net/fea/html (cited July 2005).
Wikipedia, the free encyclopedia, Fast Fourier Transform, available
online at http://en.wikipedia.org/wiki/Fast_Fourier_transform
(cited June 2005).
Wilkinson, J. J., and Reinsch, C., Handbook for Automatic Com-
putation, Vol. II—Linear Algebra, Springer–Verlag, New York,
1971.
657
forward whirl = whirl motion in the same direction as the spin rotation
of the body
frequency response = complex-valued ratio of output response (amplitude
function and phase) to sinusoidal input force, as a function of
frequency (can be a scalar or matrix quantity)
Froude scaling = special type of model scaling wherein the velocities
are scaled to the square root of full-scale values
gyroscopic effect = tendency for particles in a spinning body to react
to a moment applied to the body in a nonspinning
coordinate frame so as to achieve a rotational velocity
a quarter of a revolution removed from the applied
moment
hingeless rotor = rotor with no actual mechanical hinges that
achieves flapping and lead-lag motion by elastically
flexing
hinges = mechanisms that hold the blades proper to the hub
and allow free angular motion with zero moment
transfer
hub = structure that holds the blades to the rotor shaft
hub moment = couple produced on the hub by the blades at their hub
attachment
impedance = ratio of force quantity(s) to response quantity(s), for
sinusoidally excited systems. The term is used gen-
erally to describe such a ratio wherein the response is
taken to be either displacement, velocity, or acceler-
ation, and used specifically wherein the response is
velocity (see mechanical impedance). Whether used
generally or specifically, impedance is the reciprocal
(inverse) of mobility as long as the response quanti-
ties are consistently defined. Impedance and mobility
can be scalar or matrix quantities.
indicial response = dynamic response caused by a unit input
induced velocity = downward air velocity generated by the rotor in
the process of developing (upward) rotor thrust, or
locally by a vortex
inflow = downward component of air velocity through and
perpendicular to the rotor disk
inflow ratio = ratio of total average velocity normal to the plane
of rotation to the tip speed, normally taken positive
when flowing downward through disk
isotropic = having the same properties in any direction
laminate = in a composite structure, the stacking of any number
of ply groups with dissimilar properties (fiber angle,
thickness, fiber type, etc.) to form a two-dimensional
sheet of finite thickness
lateral cyclic control = component of cyclic control used to roll the
helicopter
null matrix = any matrix whose elements are all zero; notation: [0]
matrix minor = In any square matrix the minor of the element Aij ,
Mij is the remaining determinant if the ith column
and jth row are deleted.
cofactor = cofactor of the element Aij , is the minor Mij multi-
plied by (−1)i + j
singular matrix = matrix is said to be singular if its determinant is zero
coefficient matrix = Given a set of linear equations in the unknowns {x},
the coefficients of {x} form the coefficient matrix.
augmented matrix = coefficient matrix to which the column of constants
has been added
rank of a matrix = order of the largest square array in the matrix of
which the determinant does not vanish
transposed matrix = matrix AT = [Aji ] is called the transpose of any given
matrix A(=[Aij ]); matrix that is equal to its transpose
is said to be symmetrical.
adjoint matrix = If A is a square matrix, the matrix obtained from A
by replacing each element by its cofactor and then
transposing the resulting matrix is called the adjoint
of A.
mechanical = for sinusoidally excited mechanical systems, the
admittance ratio of response velocity to excitation force;
mechanical admittance (mobility) is the reciprocal
or inverse of mechanical impedance
mechanical = for sinusoidally excited mechanical systems, the
impedance ratio of excitation force to response velocity;
impedance is the reciprocal of mechanical admit-
tance (mobility)
mobility = for sinusoidally excited systems, the ratio of response
quantity to force quantity; the term is used gener-
ally to describe such a ratio wherein the response
is taken to be displacement, velocity, or accelera-
tion and used specifically wherein the response is
velocity (see mechanical admittance). Whether used
generally or specifically, impedance is the reciprocal
(inverse) of mobility as long as the response quanti-
ties are consistently defined. Impedance and mobility
can be scalar or matrix quantities.
node point = point in a dynamic system, oscillating at one of its
natural frequencies, where the response is zero
orthotropic = having two dissimilar planes of symmetry at right
angles to each other
pitch = angle between a blade chord line and a plane perpen-
dicular to the rotor mast
pitch-flap coupling = automatic kinematic pitch change caused by flapping
motion, as results from the delta-three hinge
pitch-lag coupling = automatic kinematic pitch change caused by lead-lag
motion, as results from the alpha-two hinge
665
Fig. B.1 Bending frequency coefficients an for hinged beams with linear mass and
stiffness distributions.
Fig. B.2 Zero offset Southwell coefficient K0n for hinged beams with linear mass and
stiffness distributions.
Fig. B.3 Offset correction factors for Southwell coefficients K1n for hinged beams
with linear mass and stiffness distributions.
Fig. B.3 (continued) Offset correction factors for Southwell coefficients K1n for
hinged beams with linear mass and stiffness distributions.
Fig. B.4 Bending frequency coefficients an for cantilever beams with linear mass and
stiffness distributions.
Fig. B.5 Zero offset Southwell coefficient K0n for cantilever beams with linear mass
and stiffness distributions.
Fig. B.6 Offset correction factors for Southwell coefficients K1n for cantilever beams
with linear mass and stiffness distributions.
Fig. B.7 Bending frequency coefficients for nonrotating uniform hinged beams with
a mass at the tip, r = tip mass/beam mass.
Fig. B.8 Zero offset Southwell coefficients for a uniform hinged beam with a mass at
the tip, r = tip mass/beam mass.
Fig. B.9 Bending frequency coefficients for nonrotating uniform cantilever beams
with a mass at the tip, r = tip mass/beam mass.
Fig. B.10 Zero offset Southwell coefficients for a uniform cantilever beam with a
mass at the tip, r = tip mass/beam mass.
mobilities. The superscripts identify the substructures from which the partition
originates (i.e., the substructure into which we are looking when we define the
mobility). Accordingly,
a
YAA ≡ nA × nA , AA partition of the mobility matrix from substructure a as
defined by set A
a
YIA ≡ nI × nA , IA crosspartition of interface/internal mobilities as defined
by sets I and A
YIIa ≡ nI × nI , II partition of the mobilities, as defined by set I
qA , qI ≡ subsets of the internal and interface degrees of freedom
fA , fI ≡ subsets of internal and interface forces
T
a
YAI a
YAI
−1
a
− YIIa YIIa + YIIb YII (C.2)
−YBIb −YBI b
Note that in this equation the interface degrees of freedom are retained with system
A (and not B). The interface degrees of freedom could equally as well have been
carried with B (and not A), and an equivalent expression with terms rearranged
would be achieved.
By a redefinition of subscripts where à now represents the union of A and
I (Ã = A ∪ I), Eq. (C.2) can also be written in more compact form:
ab T
YÃÃ YÃB Ya 0 Ya −1 Ya
ÃÃ ÃI ÃI
= − YIIa + YIIb (C.3)
YBÃ YBB 0 b
YBB −YBI
b −YBI
b
where the special à partitioning for the substructure carrying the interface is to be
noted. The important characteristic of these two relationships is that the coupling
process, wherein the total coupled mobilities are found, utilizes (additionally) only
interface information from the two unconnected subsystems. Also, one should
note that the algebra involved is quite efficient since the inverse, representing
the most time-consuming calculation, is only of order nI × nI , and is generally
much smaller than the system matrices themselves. These equations form the
fundamental foundation for the generalization of the coupling process for any
number of substructures. This generalization is presented next.
foresection through interface set (M), and to the transmission (D) through interface
set (K). And last, the transmission is connected to the turbine through interface
set (K) and to the fuselage through interface set (L). After separating out the sets
of common interfaces for each substructure, the coupling process of Fig. C.2 can
be pictured symbolically in Fig. C.3.
By labeling the sets of interfaces, the coupling process is now fully described.
The utilization of graph theory for the assembly process at hand will assure proper
sign selection of common coordinates. In Fig. C.3 a “tail” is identified by a (+), and
an “arrow” is identified by a (−). Although this convention represents an arbitrary
selection of arrows for common connection points, it will make the coupling of
the subsystems consistent. In accordance with Fig. C.3 a Boolean mapping matrix
[M] can be constructed:
1 0 0 1 1 A
−1 1 0 0 0 B
[M] =
0
(C.4)
−1
Bodies
−1 1 0 C
0 0 −1 −1 0 D
I J K L M
Interface sets
This matrix defines which “bodies” are connected and through which “interface
sets” the connections are to be accomplished. It also reflects the arbitrary sign
convention chosen for Fig. C.3. For example, the first column of Eq. (C.4) can be
constructed in the following manner: According to Fig. C.3, the interface set I is
given a +1 on A and consequently a −1 on B. This corresponds directly to the
first column values given in Eq. (C.4).
In order to complete the coupling process, another Boolean map is needed. This
matrix organizes the interface degrees of freedom and their interrelations. Such a
matrix is formed from the rows of the previously defined [M] mapping matrix:
where [Mi ] is the ith row of the [M] matrix. Consequently there exist as many
of these matrices as there are subcomponents to assemble. Consider, for instance,
substructure C, the third substructure in the coupling sequence:
0 0 0 0 0 I
0 1 −1 0 1 J
[N (3) ] = [Mi ]T [Mi ] =
0 −1 1 0 −1 K (C.6)
(i=3) 0 0 0 0 0 L
0 1 −1 0 1 M
I J K L M
where [M3 ] is the third row of [M], the mapping matrix defined by Eq. (C.4).
The individual terms in this matrix correspond to 1) the partition of the point
and the cross mobilities that are present for a particular substructure and 2) the
sign convention for the same. Similar matrices can be derived for the A, B and D
substructures.
With all of these tools for the assemblage process in mind, the generalized
mobility coupling equation for N subassemblies can then be written as
[Y ] = [Yαα ] − [M] ⊗ [Yαγ ]
N
−1 T
· [N ] ⊗ Yγ γ
(i) (i)
[M] ⊗ [Yαγ ] (C.7)
i=1
where
α ≡ A, B, . . . , K, substructure identification in addition to set definition of
internal degrees of freedom
γ ≡ I, J, . . . , L, (interface set definition)
Y ≡ total synthesized mobility matrix
Yαα ≡ diagonal mobility matrix of the uncoupled subsystems as defined by
sets α
M ≡ mapping matrix, as defined by Eq. (C.4)
Yαγ ≡ mobility matrix for the cross partitionings between the internal and
interface degrees of freedom, as defined by sets α and γ
Y(i)
γγ ≡ mobility matrix for the cross and point relations between the interface
degrees of freedom, as defined by set γ for substructure i
N(i) ≡ ith auxiliary mapping matrix, formed from the ith row of [M] as defined
in Eq. (C.5)
N = number of substructures
where {qγ+ } and {aγ− } are displacements in correspondence with the sign convention
chosen in the map of Eq. (C.4).
A new general equation, similar to Eq. (C.7) for the synthesis of N substructures,
can then be written:
N −1
(i) T
[Y ] = [Yαα ] − [M] ⊗ [Yαγ ] · [N ] ⊗ [Y γ γ ]
(i)
[M] ⊗ [Yαγ ]
i=1
(C.10)
where all terms, except [Ȳγ(i)γ ], remain the same as those in Eq. (C.7). [Ȳγ(i)γ ] is the
mobility matrix of the interface relations with the modification that the diagonal
terms have the connection impedance added in for the interface set in question.
For example, for interface I with an “impedance connection” between common
interface set I for substructures A and B, the diagonal term would read:
Assumptions
The principal assumption made in this analytical formulation is that both the
ground- and air-resonance phenomena can be adequately modeled using linear
constant coefficient differential equations of motion. That ground resonance can
be so modeled has been adequately verified experimentally [see Bielawa, 1962],
and the use of linear analyses is thus well-established. However, the assumption
that air resonance can also be modeled using only linear differential equations is
less secure, and certain nonlinearities must be addressed and linearized for inclu-
sion when appropriate. The predominance of linearity is recognized, however, on
the basis that air resonance is basically a phenomenon closely akin to ground res-
onance. Like ground resonance, air resonance involves first-order blade elastic
in-plane bending effects (reasonably and adequately describable using linear
dynamics) as a principal dynamic ingredient. Finally, although not yet experi-
mentally verified, the linear analysis presented herein has been shown to be capable
of predicting “air resonance-like” instabilities.
Another principal assumption is that, air resonance can be modeled with the use
of rotor mode descriptions of the blade in both flapwise and inplane bending. This
assumption is made on the basis that the rotor is isotropic (i.e., having three or more
blades) and is thereby subject to the same basic approach followed in the work of
685
Coleman and Feingold for ground resonance. To make the formulation sufficiently
general and yet particularly applicable to hingeless and bearingless rotor types,
the equations are developed using a modal formulation for the blade motions. The
equations of motion are therefore formulated using a Galerkin technique based on
the following detailed assumptions:
1) The transverse deflection of the blade is elastomechanically represented by
a single (uncoupled, but rotating) spanwise variable normal bending mode each in
the flatwise and edgewise directions. These modes are “uncoupled” in the sense
that they are calculated assuming that there is zero inertia or elastic coupling caused
by pitch, twist, or precone angles.
2) The blades respond quasi-statically in torsion. The principal nonlinear
elastomechanical coupling between the flatwise and edgewise modes is retained
and linearized. This linearization is achieved using appropriate trimmed flight
conditions.
3) The blade has a constant value of built-in coning angle βB maintained by
the aerodynamic lifting loads.
4) The rotor hub is attached to a general pylon substructure that does not rotate
at the rotor speed. This pylon substructure is defined in terms of the five rigid-body
degrees of freedom of the hub: longitudinal, lateral, and vertical translations, and
pitch and roll rotations. For explicit ground- and/or air-resonance applications the
pylon substructure is taken to be the fuselage, which is assumed to be a rigid body.
5) The effects of control angles, as determined by appropriate trimmed flight
conditions, are retained in the aerodynamic formulation.
6) The hub loads are equilibrated in the hub-fixed, nonrotating coordinate sys-
tem. Thus, longitudinal and lateral (gravity) forces arise in the formulation as a
result of rotations of the fuselage in pitch and roll, respectively.
7) For free-flight conditions the equation governing the vertical translation
pitch-angle degrees of freedom will contain load factor augmenting terms propor-
tional to velocity and pitch rate. For constrained flight (wind-tunnel) conditions
these terms are omitted.
8) Basic quasi-steady aerodynamic theory is used to define the unsteady air-
loads on the blade. Low angle-of-attack (unstalled) conditions are assumed: The
lift coefficient cl is represented by a lift-curve slope a(= clα ) times an angle of
attack α, and the drag coefficient cd is represented by a uniform constant drag
coefficient cd0 together with even-powered variations in angle of attack δn α n . The
moment coefficient cm is assumed to be zero.
Kinematic Descriptions
Pertinent Degrees of Freedom
For both the ground resonance and air resonance phenomena, the basic mathe-
matical description is provided by 11 differential equations that respectively model
the responses in the hub x, y, and z translations, hub roll and pitch rotations, and
rotor mode descriptions for the blade elastic responses: blade cyclic in-plane (edge-
wise) bending in the x and y directions, blade cyclic flapwise (flatwise) bending in
the roll and pitch directions, and finally, cosine and sine components of the torsion
I cos ψ − sin ψ −βB cos ψ i i
J = sin ψ cos ψ −βB sin ψ j = [U] j (D.2)
K βB 0 1 k k
Fig. D.1 Depiction of the sets of unit vectors in the two moving coordinate systems.
r = ri + y5 j + z5 k (D.3)
where the elastic in-plane and out-of-plane deformations are, respectively, given
by the first bending modal responses:
where qv0 and qw0 are the steady in-plane and out-of-plane modal deformations,
respectively, as determined by the trimmed flight condition. The derivation of the
dynamic terms follows from a straightforward application of Coriolis theorem.
The theorem accurately gives the components of inertial acceleration as measured
in each of the two moving coordinate systems. Upon denoting acceleration with
respect to inertial space as [ p2I ] and simple differentiation with respect to time as
[p], the Coriolis theorem can be used to write the inertial acceleration vector for
the position vector r for either of the two moving coordinate systems (ν = 1, 2):
where ωivv is the angular velocity of the vth coordinate system (v = 1 for the
hub-centered coordinate system and v = 2 for the rotating, coned blade axis coor-
dinate system) relative to an inertial coordinate system. The angular velocity vector
of the hub-centered axis system (relative to the inertial frame) is expressed as
On the other hand, the angular velocity vector of the rotating, coned-blade axis
system is expressed as
Additionally in the former case of the hub-centered coordinate system, the hub
velocity v0 is composed of a finite steady-state value in the negative longitudinal
x direction V and perturbational values in each of the three component axes:
After evaluation and combination of the various terms in the preceding equation,
the inertial acceleration vector can then be expressed using the rotating, preconed
where the κ factor is equal to 1 or 0, depending on whether the pylon (fuselage) is,
respectively, in a free-flight or gimballed (constrained) configuration. The preced-
ing expression for inertial acceleration can then be represented in the following
abbreviated form:
i
p2I r = ARx ARy ARz j (D.9)
k
where the ARx , ARy , and ARz vector components represent the components of the
inertial acceleration vector, as measured in the number 2 (coned and rotating)
moving coordinate system.
The dynamic terms in the rotor degree-of-freedom equations resulting from
rotor mass elements can then be obtained from the preceding equation using
D’Alembert’s principle together with an application of the Galerkin technique.
A separate application of the Galerkin technique is made for each of the appro-
priate Fourier rotor mode components of the blade elastic modal variables. To
obtain the dynamic inertia terms caused by rotor blade mass that are required
for the nonrotating degree-of-freedom equations, it is much more appropriate and
convenient to take inertial acceleration vector components in the directions of the
unit vectors in the number 1 (unconed and nonrotating) moving coordinate system:
I
p2I r = [ARx ARy ARz ] [U]−1 J (D.10a)
K
I
pI r = [ANRx ANRy ANRz ] J
2
(D.10b)
K
The three components of inertia loads and the rolling and pitching components
of inertia moment at the hub caused by rotor mass (as expressed in the nonrotating
coordinate system) can then be expressed in terms of the ANRα coefficients resulting
from the operations defined by Eqs. (D.2), (D.8), and (D.10a).
Dynamic terms caused by rotor stiffness and damping. One great advan-
tage of the use of a modal description for the blade bending equations is the ease
with which the stiffness and (viscous equivalent) structural damping can be intro-
duced into the blade bending equations. These generalized stiffness and damping
terms for the edgewise bending excitation can be written in terms of the generalized
mass as follows:
R
(s)
+ (d)
= R2 m γv2 dr [ωv2 + 2ζv ωv0 ]
˙
0
In a similar manner the stiffness and structural damping terms in the flapwise
bending equations can be introduced:
R
m γw2 dr [ωw2 θR + 2ζw ωw0 θ̇R ]
(s) (d)
θR + θR = R2
0
The Kw factor is now the Southwell coefficient for the first flatwise bending mode,
and the viscous equivalent structural damping is formulated the same way.
Torsion equation. The appropriate starting point for formulating the quasi-
static torsion response is Eq. (3.114), written in abbreviated form as
After the usual application of Galerkin’s method, the simplified equation for
blade torsion can then be written as
S77 θ̈ blade root + S78 (q̈θ + 2ζθ ωθ q̇θ + ωθ2 qθ ) + 2 (S80 cos 2θ0 − S78 )qθ
(a)
= φ − S90 (qw0 + qv0 θR ) (D.18)
where the angular acceleration at the blade root arises from the angular motion
(a)
of the hub in pitch and roll. The aerodynamic torsional excitation φ can be
expressed as a combination of cosine and sine components, using the hub angular
rates and the rotor-mode description for the torsional response. This aerodynamic
excitation is developed in detail in a subsequent subsection.
Dynamic terms resulting from rotor and fuselage gravity loads. One of
the ramifications of choosing a body-fixed coordinate system is that the orientation
of the axes relative to the horizon is variable. Hence, components of the gravity
force in the x and y directions will vary in accordance with the fuselage pitch and
roll angles. Because the gravity force of the total aircraft can be expressed using
the rotor thrust T , the hub force terms caused by gravity can be written as
(g) (g) (g)
FH = FHR + FHR = g(mR + mF )[θy I − θx J − K]
= T [θy I − θx J − K] (D.21)
where, by considerations of equilibrium, T is the total rotor thrust, and where
only the first two components are considered because they are perturbational. In a
manner similar to that followed for the inertial loads, the gravitational hub moments
caused by gravity loads can then be similarly defined.
b b
+ βB S16 θ̈yR + S48 ¨x − T θyF = Fx(a) + Fx(h) (D.22)
2 2
Hub lateral force Fy :
b b
− βB S16 θ̈xR + S48 ¨y − T θxF = Fy(a) + Fy(h) (D.23)
2 2
(mFz + mR )z̈ − mFz xcg θ̈yF + κ(mFz + mR )V θ̇yF = Fz(a) + Fz(h) (D.24)
b b
+ S2 (1 + βB2 ) θ̈yF + S12 θ̈yR
2 2
b b
− 2 S2 θ̇xF + (1 − κ) cFθy θ̇yF − 2 S12 θ̇xR
2 2
b
+ βB S46 ¨x + [mFx gh1 + (1 − κ)kFθy ]θyF
2
S48 ẍ + βB S46 θ̈yF + S49 [¨x + 2ζv ωv0 ˙x + (ωv2 − 2 )x ]
S48 ÿ − βB S46 θ̈xF + S49 [¨y + 2ζv ωv0 ˙y + (ωv2 − 2 )y ]
S77 (θ̈xF + θ̇yF )+ S78 φ̈x + 2ζθ ωθ φ̇x +(ωθ2 − 2 )φx + 2 (S80 cos 2θ0 − S78 )φx
(a)
+ 2S78 (φ̇y + ζθ ωθ φy ) + S90 (qw0 y − qv0 θyR ) = φx (D.31)
S77 (θ̈yF + θ̇xF )+ S78 φ̈y + 2ζθ ωθ φ̇y +(ωθ2 − 2 )φy + 2 (S80 cos 2θ0 − S78 )φy
(a)
−2S78 (φ̇x + ζθ ωθ φx ) + S90 (−qw0 εx + qv0 θxR ) = φy (D.32)
The ( )(h) terms represent arbitrary explicit forces and/or moments applied
directly to the hub from some other source. For some applications these ( )(h) terms
can be expressed in terms of existing degrees of freedom, or they can be used more
generally to define rotor aeroelastic operators, that is, aeromechanical transfer
functions. Note that the preceding equation set is intended for the dual purpose of
modeling both ground resonance and air resonance characteristics. For simplified
(minimal sophistication) ground-resonance applications, only Eqs. (D.22), (D.23),
(D.27), and (D.28) need to be used, with the ( ) terms suppressed. A more
rigorous modeling of ground resonance is affordable by including the fuselage
and rotor degrees of freedom. In any given application of the preceding equations,
the control over which equations and degrees of freedom are to be included in the
equation set can be accomplished using the formalized constraint matrix described
in a subsection to follow.
For air resonance applications all of the equations are generally used, but with
the ( ) terms suppressed. Additionally, the ( ) terms serve a dual
purpose; in the case of a more rigorous ground resonance modeling, wherein the
roll and pitch degrees of freedom must be included, they allow a more direct
modeling of the landing-gear stiffnesses and damping characteristics. In the case
of air resonance, they can provide the modeling of the restraint moment about
some gimbal point (as with wind-tunnel installations) for the cases that are so
constrained.
The various (inertia) integration constants appearing in Eqs. (D.23–D.33) are
all functions of the spanwise mass distribution m (having the units of lb-s2 /in2 )
and are defined as follows:
R
S1 = m r dr (D.33a)
0
R
S2 = m r 2 dr (D.33b)
0
R
S10 = R2 m γw2 dr (D.33c)
0
R
S12 = R m rγw dr (D.33d)
0
R
S16 = R m γw dr (D.33e)
0
R
S25 = R2 m γv γw dr (D.33f)
0
R
S46 = R m rγv dr (D.33g)
0
R
S48 = R m γv dr (D.33h)
0
R
S49 = R2 m γv2 dr (D.33i)
0
R
S77 = R 2 2
m γθ k̄m2 + k̄m1
2
dr (D.33j)
0
R
S78 = R2 m γθ2 k̄m2
2
+ k̄m1
2
dr (D.33k)
0
R
S80 = R2 2
m γθ2 k̄m2 − k̄m1
2
dr (D.33l)
0
R
−2
S90 = R {γθ γw γv (EIE − EIF )} dr (D.33m)
0
the blade radius R. Using this approach, the preceding (dimensional) integration
constants can be expressed as follows:
R 1
m
S1 = m r dr = m0 R 2
r̄ dr̄ = m0 R2 S̄1 (D.34a)
0 0 m0
Similarly,
S2 = m0 R3 S̄2 (D.34b)
S10 = m0 R3 S̄10 (D.34c)
S12 = m0 R3 S̄12 (D.34d)
S16 = m0 R S̄16 2
(D.34e)
S25 = m0 R3 S̄25 (D.34f)
S46 = m0 R S̄46 3
(D.34g)
S48 = m0 R2 S̄48 (D.34h)
S49 = m0 R3 S̄49 (D.34i)
S77 = m0 R S̄77 3
(D.34j)
S78 = m0 R3 S̄78 (D.34k)
S80 = m0 R S̄80 3
(D.34l)
S90 = R−1 S̄90 (D.34m)
Note that S̄90 still has the units of bending stiffness (i.e., lb-in2 ).
the following expressions for the tangential and perpendicular components of air-
foil sectional velocity and the airfoil pitch angle and pitch-angle rate θ and θ,θ̇ ,
respectively:
Tangential section air velocity component UT :
UP = RŪP
= R φ + µ[(1 − κ)θyF − cos ψ[βB + γw (θxR sin ψ − θyR cos ψ)]]
1
+ βB (ẋ cos ψ + ẏ sin ψ)/R + r̄(θ̇yF cos ψ − θ̇xF sin ψ) − θyR cos ψ
+ γw [(θ̇yR − θxR ) cos ψ − (θ̇xR + θyR ) sin ψ] (D.36)
Note that in Eq. (D.36) and elsewhere in this appendix, the prime symbol
denotes differentiation with respect to nondimensional span r̄.
The rotor inflow is represented by the inflow angle φ based on the inflow
and advance ratios, both of which have distinct radial distributions depending
on whether the rotor is in hovering or forward flight. The “0” subscript denotes
that dynamic (perturbational) variables are omitted:
ŪP0 −1 −λ(r̄)
φ = arctan = tan (D.39)
ŪT0 r̄ + µ sin ψ
The kx and ky coefficients are functions of both advance ratio and inflow ratio and
are defined in detail in Chapter 10.
The perturbational in-plane and out-of-plane force and pitching-moment airload
distributions from which all of the required aerodynamic coefficients are obtained
are given by
(a)
d[δfy ] 1
= δfy(a)
dr R
ρac2 R2 cd δn
= ŪT0 −2 + φ θ0 + n α (n−1) δ ŪT
2 a a
δn (n−1) δn (n−1)
+ θ0 + 2φ − n α δ ŪP + ŪT0 φ − n α δθ
a a
ρac2 R2 ¯ (a)
= δ fy (D.41)
2
d[δfz(a) ] 1 ρac2 R2
= δfz(a) = ŪT0 {[2θ0 + φ]δ ŪT
dr R 2
ρac2 R2 ¯ (a)
+ δ ŪP + ŪT0 δθ } = δ fz (D.42)
2
d[δmx(a) ] 1 ρπ c3 R
= δmx(a) = {−ŪT0 δ θ̇ }
dr R 8
ρπc3 R (a)
= δ m̄x (D.43)
8
The angle of attack α in Eq. (D.41) is the sum of the pitch angle θ0 and the inflow
angle φ.
= (a)
w0 − (gS16 + βB 2 S12 ) cos θ0 (D.58)
where the steady aerodynamic generalized modal loads are determined by any
standard blade element theory [see Leishman and Prouty]:
R
dFy0 dFz0
v0 = R cos θ0 +
(a)
γv (r) sin θ0 dr (D.59)
0 dr dr
R
dFy0 dFz0
w0 = R γw (r) − sin θ0 +
(a)
cos θ0 dr (D.60)
0 dr dr
Equations (D.59) and (D.60) require the steady in-plane and out-of-plane airload
distributions, which in turn require a suitable trim calculation to determine the
aircraft equilibrium, and more immediate, the inflow ratio and blade (control)
pitch angles. The development and presentation of a suitable trim calculation is
beyond the intent of this text, and the reader is referred to Leishmann, Johnson,
and Prouty for excellent treatments of the equations for rotor trim.
where the total vector of degrees of freedom {Z} is defined by Eq. (D.1), and
{Lξ(h) } is the five-element vector of additional excitation loads at the hub. For
some potential applications of these equations, it might arise that some one or
several of the degrees of freedom can either be considered to be identically zero
or are to be arbitrarily constrained to zero (perhaps to investigate their importance
to an instability). In either case, it would be desirable to be able to simplify the
equations to the resulting smaller size in a convenient, accurate manner. One
method for doing this is to define the constraint on the degrees of freedom using
a constraint matrix. Thus, after neglecting {Lξ(h) } for the moment, we define a
reduced set of independent degrees of freedom as denoted by {Z̃}, which is then
related to the vector of 11 original degrees of freedom {Z} as a linear combination.
Such a relationship can be described in a straightforward manner using a matrix
equation, as follows:
where the [T ] matrix will be 11 × n matrix, where n will generally be <11. The
constraining out of selected degrees of freedom is then achieved by first substituting
Eq. (D.62) into Eq. (D.61) and then premultiplying the resulting matrix equation
by [T ]T . This then produces a requisite square matrix formulation of the dynamic
equations:
This equation is then in a form suitable for an application of any standard eigen-
value solution.
constrained to zero:
xF /R 0 zfoc /R
−zfoc /R
yF /R 0 θ
xF
zF /R = 0 0 (D.64a)
θyF
θxF
1
0
θyF 0 1
θ
= [Tfoc ] xF (D.64b)
θyF
where Zi are the rotor aeroelastic degrees of freedom defined in Eq. (D.1). These
incremental hub loads Lξ(h) essentially define rotor aeroelastic operators that can
be included in the dynamic equations of motions of other aeroelastic systems as
concentrated forces or moments, as appropriate. Thus, as shown in Fig. D.2, these
hub loads can then be applied to the nonrotor substructure in a direction opposite
to the sense they are defined for the rotor.
In the following material, equations of motion are considered for the hovering
mode configuration of a wing with a rotor-nacelle substructure attached to it at a
spanwise station sR and with the rotor vertically located relative to this attachment
point by the distance hR . With the implementation of a Galerkin formulation for
the elastic motion (bending and torsion) of the wing, we can write the equations
of motion for these wing degrees of freedom as integrals of the wing loadings
Lw and Lθ , respectively, multiplied by the bending and torsion mode shapes ηw
and ηθ over the semispan of the wing sW . Note that the Lw and Lθ loadings
represent dynamic loads resulting only from the explicitly wing motion degrees of
freedom qWw and qWθ , that is, the wing bending and torsion generalized (modal)
variables, respectively. Stated another way, these loadings represent the dynamic
loads on the wing in the absence of the rotor and nacelle. The actual inclusion of
the incremental hub loads in this integral formulation can then be accomplished
in two basic steps. First, the hub position variables x, y, z, θxF , and θyF must be
expressed in terms of the wing bending and torsion motion variables qWW and qWθ :
x/R 0 ηθ (sR )hR /R
−η (sR )hR /R
y/R w 0 q
Ww
z/R = ηw (SR )/R 0 (D.66)
(s ) qWθ
xF
θ
η 0
w R
θyF 0 ηθ (sR )
qWW
= [TR/W ]hov (D.67)
qWθ
Note that for the forward-flight condition a similar transformation matrix can be
defined.
Then, the incremental hub loads (as seen by the wing) are treated mathematically
as negative delta functions in the wing integral (Galerkin) formulations. Thus, the
resulting wing bending equation would be of the following form:
sW
ηw Lw ds + ηw (sR ) −Fz(h) + ηw (sR ) −Mx(h)
F
0
Equations (D.68) and (D.69) can be combined with the [TR/W ]hov matrix defined
by Eqs. (D.66) and (D.67) to yield a unified formulation of the coupled equations
of motion that contain the rotor-wing coupling and are expressed only in terms
of the wing bending degrees of freedom. This is accomplished by first expressing
εxx a11 a12 a13 σxx
ε
a σ
yy a22 a23 0
yy
21
εzz a a32 a33 σzz
= 31 (E.1)
ε
a44 0 σyz
yz
εxz
σ
0 a55
xz
εxy 0 a66 σxy
[It should be kept in mind that although strains are quantities that can be mea-
sured, stresses are really mathematical abstractions (albeit good ones) that can
only be “measured” by measuring strains and then using them through Hooke’s
law to obtain stresses.]
707
εxx 1 −ν −ν
σxx
ε
σ
yy
−ν −ν
yy
1 0
1
εzz −ν −ν 1 σzz
= (E.2)
ε
2(1 + ν) 0 σyz
yz E
εxz
2(1 + ν)
σ
0
xz
εxy 0 2(1 + ν) σxy
1 ν
− 0
E E σx
εx ν
1
εy =
− E 0 σy (E.3a)
γxy E τxy
1
0 0
G
The inverse matrix representation of Eq. (E.3a), generally defining the stiffness
characteristics, is given as
E νE
0
(1 − ν 2 ) (1 − ν 2 )
σx εx
σy = νE E εy (E.3b)
0
τxy (1 − ν 2 ) (1 − ν 2 ) γxy
0 0 G
Figure E.1a shows how the stresses and strains are oriented on a two-
dimensional element of isotropic material. In the case of uniaxial composite
materials, where there is a distinct orientation of the fibers, the stress-strain
relationship now must use four basic properties, Ex , Ey , Es , and νx (or, alter-
natively, νy ), and take the fiber orientation into account. Figure E.1b identifies
the corresponding stress-strain field for the uniaxial composite two-dimensional
element.
where
Q11 m4 n4 2m2 n2 4m2 n2
n4
Q22
m4 2m2 n2 4m2 n2 Qxx
Q 2 2
m n m 2 n2 m 4 + n4 −4m n2 2
Qyy
=
12
(E.8)
Q66
2 2
m n m 2 n2 −2m2 n2 (m2 − n2 )2
Qxy
Q16 m3 n −mn3 mn3 − m3 n 2(mn3 − m3 n) Qss
Q26 mn3 −m3 n m3 n − mn3 2(m3 n − mn3 )
Fig. E.2 Determination of the off-axis stiffnesses (after Tsai and Hahn).
and where m = cos θ , n = sin θ , and Qxx , Qyy , Qxy , and Qss are as defined by
Eq. (E.6).
Multidirectional Laminates
A practical composite structure will use uniaxial composite plies and ply groups
(plies all sharing the same properties and fiber orientations) with a variety of arbi-
trary fiber orientations and sequences. Such a built-up configuration is denoted a
laminate, which can be either symmetrical about some midplane or unsymmetrical.
Laminate Codes
To identify the laminate configuration, a stacking sequence must be used. Fig-
ure E.3a shows a typical stacking sequence for a symmetric laminate. The code
used to define the stacking sequence for a multidirectional composite identifies not
only the order of the orientations (in an ascending order from the bottom ply), but
also the number of plies within each ply group, written as a subscript. Thus, the
appropriate code for the laminate shown in Fig. E.3a, written with the symmetric
notation, would be
Alternatively, the stacking sequence could be codified using the total notation,
with the subscript T :
[03 /902 / − 45/456 / − 45/902 /03 ]T (E.10)
Because the flexural stiffness and strength characteristics of a laminate are
greatly enhanced by placing the tensile-carrying fibers as far from the neutral
plane as possible, the use of sandwich configurations has become quite practical.
Figure E.3b shows a typical stacking configuration with the inclusion of a
nonstructural core between two load carrying laminates.
The core material is usually referred to as a cellular solid and can be any of
a wide variety of metallic, polymer, or ceramic (foams and/or two-dimensional
honeycomb). Gibson and Ashby present a comprehensive exposition of such mate-
rials. For the most part such materials are used in composite structures in a
nonstructural role, principally for holding the load-carrying laminates in place.
With regard to codifying the ply sequencing with a core insert, the configuration
shown in Fig. E.3b is representative and would be codified as
[03 /902 / − 45/453 /zc ]S (E.11)
where the individual σ1 , σ2 , and σ6 stress components for each of the plies are
orientated to the same reference direction using Eqs. (E.7) and (E.8). Also, using
Fig. E.4 In-plane stiffness of a typical symmetric laminate plate, showing corre-
sponding components of the actual ply stresses and the average across the laminate.
N1 = hσ̄1 (E.13a)
N2 = hσ̄2 (E.13b)
N6 = hσ̄6 (E.13c)
Quasi-Isotropic Laminates
For some applications it might be required that the composite laminate behave
exactly as a metallic (isotropic) material. This can be accomplished if the following
conditions are met:
[0/60/ − 60]S
For a symmetric laminate this defines a minimum of six plies. Similarly, for
the Pi/4 lay up we must have equal numbers of plies with orientations in the 0-,
90-, 45-, and −45-deg directions, for a minimum of eight plies. For this case the
laminate designation is
[0/90/45/ − 45]S
w = w(x, y) (E.16)
and where the “1” and “2” axes are taken to be the local x and y axes respectively.
Three applied moments will produce flexing of the plate, the M2 moments shown
in Fig. E.5, similar opposing moment M1 about the “2” axis, and M6 , co-linear
twisting moments about the “1” axis. These moments can be expressed in terms
Fig. E.5 Flexing and stress-strain conditions on the sides of a symmetric laminate
plate caused by the sustained moments on the 2 axis faces.
Because σ6 is a shear stress, M6 will take the form of moments applied along
the “1” axis, whereas M1 and M2 are applied at the edges of the plate parallel
to the “2” and “1” axis, respectively. To complete the stress-strain description,
the strains must be defined in terms of the displacements. Thus, in terms of the
out-of-plane displacement w, the displacements in the “1” and “2” directions, u
and v, respectively, are expressed as
∂w
u = −z (E.18a)
∂x
∂w
v = −z (E.18b)
∂y
∂u ∂ 2w
ε1 = = −z 2 (E.19a)
∂x ∂x
∂v ∂ 2w
ε2 = = −z 2 (E.19b)
∂y ∂y
∂u ∂v ∂ 2w
ε6 = + = −2z (E.19c)
∂y ∂x ∂x∂y
The second derivatives in the expressions for the ε1 and ε2 strains can be related
to curvatures k1 and k2 , respectively, and the ε6 strain can be related to a twisting
rate k6 (curvature):
∂ w 2
k1 ∼
=− 2 (E.20a)
∂x
∂ 2w
k2 ∼
=− 2 (E.20b)
∂y
∂ 2w ∂ ∂w ∂φ1
k6 ∼
= −2 = −2 = −2 (E.20c)
∂x∂y ∂x ∂y ∂x
where φ1 is the angle of twist about the “1” axis. Thus, the strains can be written as
ε1 (z) k1
ε2 (z) = z k2 (E.21)
k6
ε6 (z)
where
D11 = Q11 z2 dz (E.23a)
D12 = Q12 z2 dz (E.23b)
D16 = Q16 z2 dz (E.23c)
D22 = Q22 z2 dz (E.23d)
D26 = Q26 z2 dz (E.23e)
D66 = Q66 z2 dz (E.23f)
and where the Q11 thru Q66 terms are defined by Eq. (E.8).
References
Books
Antoniou, A., Digital Filters: Analysis and Design, McGraw–Hill Book Company,
New York, 1979.
Astrom, K. J., and Wittenmark, B., Computer Controlled Systems, Theory and Design,
(Information and System Sciences Series), Prentice–Hall, Englewood Cliffs, NJ, 1984.
Bathe, K.-J., Finite Element Procedures in Engineering Analysis, Prentice–Hall,
Englewood Cliffs, NJ, 1982.
Bisplinghoff, R. L., Ashley, H., and Halfman, R. L., Aeroelasticity, Addison–Wesley,
Cambridge, MA, 1955.
Bramwell, A. R. S., Helicopter Dynamics, Wiley, New York, 1976.
Broch, J. T., Mechanical Vibration and Shock Measurements, Larsen & Son, Denmark
(for Brüel and Kjaer), 1984.
Churchill, R. V., Operational Mathematics, McGraw–Hill, New York, 1958.
Coyle, S., Cyclic & Collective: Second Edition of the Art and Science of Flying
Helicopters, Helobooks, a Division of Mojave Books Limited, Mojave, CA, 2004.
Den Hartog, J. P., Mechanical Vibrations, McGraw–Hill, New York, 1940.
Doughtie, V. L., and James, W. H., Elements of Mechanism, Wiley, New York, 1954.
Dowell, E. H., Clark, R. C., Cox, D., Curtiss, H. C. Jr., Edwards, J. W., Hall, K. C.,
Peters, D. A., Scanlan, R., Simiu, E., Sisto, F., and Strganac, T. W., A Modern Course in
Aeroelasticity, Fourth Edition, Kluwer Academic Publishers, Norwell, MA, 2004.
Ehrich, F. F., Handbook of Rotordynamics, McGraw–Hill, New York, 1992.
Ehrich, F. F., “Self-Excited Vibration,” Shock and Vibration Handbook, Third Edition,
edited by C. M. Harris, McGraw–Hill, New York, 1988, Chap. 5.
Ewins, D. J., Modal Testing: Theory and Practice, Research Studies Press, Herts,
England, U.K. 1984.
Frazer, R. A., Duncan, W. T., and Collar, A. R., Elementary Matrices, Cambridge
Univ. Press, Cambridge, England, 1950.
Fung, Y. C., An Introduction to the Theory of Aeroelasticity, Wiley, New York, 1955.
Gessow, A., and Myers, G. C., Aerodynamics of the Helicopter, Ungar, New York,
1967.
Gibson, L. J., and Ashby, M. F., Cellular Solids - Structure and Properties, 2nd ed.,
Cambridge Univ. Press, New York, 1997.
Greensite, A. L., Elements of Modern Control Theory, Spartan Books, New York,
1970.
Haftka, R. T., Gurdal, Z., and Kamat, M. P., Elements of Structural Optimization,
Kluwer Academic Publishers, Dordrecht, the Netherlands, 1990.
Hildebrand, F. B., Methods of Applied Mathematics, Prentice–Hall, New York, 1952.
Hildebrand, F. B., Introduction to Numerical Analysis, McGraw–Hill, NewYork, 1956.
719
Huelsman, L. P., Active and Passive Analog Filter Design, An Introduction, McGraw–
Hill, New York, 1993.
Hurty, W. C., and Rubinstein, M. F., Dynamics of Structures, Prentice–Hall,
Englewood Cliffs, NJ, 1964.
Johnson, D. E., Introduction to Filter Theory, Prentice–Hall Inc., Englewood Cliffs,
NJ, 1976.
Johnson, W., Helicopter Theory, Princeton Univ. Press, Princeton, NJ, 1980.
Jones, D. I. G., “Application of Damping Treatments,” Shock and Vibration Handbook,
Third Edition, edited by C. M. Harris, McGraw–Hill, New York, 1988, Chapter 37.
Jones, R. M., Mechanics of Composite Materials, McGraw–Hill, New York, 1975.
Lamb, H., Hydrodynamics, Dover, New York, 1932.
Leishman, J. G., Principles of Helicopter Aerodynamics, rev. ed., Cambridge Univ.
Press, New York, 2005.
Myklestad, N. O., Vibration Analysis, McGraw–Hill, New York, 1944.
NATO Advisory Group for Aeronautical R & D (AGARD), Structures and Materials
Panel, Manual on Aeroelasticity, edited by W. P. Jones, London, 1960.
Nikolsky, A. A., Helicopter Analysis, Wiley, New York, 1951.
Payne, P. R., Helicopter Dynamics and Aerodynamics, Macmillan, New York, 1959.
Press, W. H., Flannery, B. P., Teukolsky, S. A., and Vetterling, W. T., Numerical
Recipies in FORTRAN, the Art of Scientific Computing, 2nd ed., Cambridge Univ. Press,
New York, 1992.
Prouty, R. W., Practical Helicopter Aerodynamics, P. J. S. Publications, Peoria,
IL, 1982.
Ralston, A., and Wilf, H. S. (eds.), Mathematical Methods for Digital Computers,
Wiley, New York, 1960.
Randall, R. B., Frequency Analysis, Larsen & Son, Denmark (for Brüel and Kjaer),
1984.
Rao, S. S., Mechanical Vibrations, Addison–Wesley, Reading, MA, 1995.
Reed, F. E., “Dynamic Vibration Absorbers and Auxiliary Mass Dampers,” Shock
and Vibration Handbook, 3rd ed., edited by C. M. Harris, McGraw–Hill, New York, 1988,
Chap. 6.
Saaty, T. L., and Bram, J., Nonlinear Mathematics, McGraw–Hill, New York, 1964.
Savant, C. J., Jr., Basic Feedback Control System Design, McGraw–Hill, New York,
1958.
Scanlan, R. H., and Rosenbaum, R., Introduction to the Study of Aircraft Vibration
and Flutter, Macmillan, New York, 1951.
Scarborough, J. B., The Gyroscope, Theory and Applications, Interscience, New York,
1958.
Schwamb, P., Merrill, A. L., and James, W. H., Elements of Mechanisms, Wiley,
New York, 1938.
Smith, B. T., Boyle, J. M., Dongarra, J. J., Garbow, B. S., Ikebe, Y., Klema, V. C.,
and Moler, C. B., Matrix Eigensystem Routines - EISPACK Guide, 2nd ed., Vol. 6, Lecture
Notes in Computer Science, Springer-Verlag, New York, 1976.
Stewart, H. L., Pneumatics and Hydraulics, Audel, (Div. of Bobbs-Merrill).
Indianapolis, IN, 1966.
Timoshenko, S., Vibration Problems in Engineering, Van Nostrand, New York, 1937.
Tong, K. N., Theory of Mechanical Vibrations, Wiley, New York, 1960.
Internet
Bellis, M., Inventors, The History of Computers, available online at
http://inventors.about.com/library/blcoindex.htm (cited June 2005).
Bellis, M., Inventors of the Modern Computer, FORTRAN - The First Successful High
Level Programming Language, available online at http://inventors.about.com/library/wee-
kly/aa072198.htm (cited June 2005).
Stewart, C. D., Finite Element Analysis/Method Resources, available online at
http://www.steelynx.net/fea/html (cited July 2005).
Wikipedia, the free encyclopedia, Fast Fourier Transform, available online at
http://en.wikipedia.org/wiki/Fast_Fourier_transform (cited June 2005).
Journal Articles
Alwang, J. R., and Skarvan, C. A., “Engine Control Stabilizing Compensation - Test-
ing & Optimization,” Journal of the American Helicopter Society, Vol. 22, No. 3, 1977,
pp. 13–18.
Bailey, C. D., “Exact and Direct Analytical Solutions to Vibrating Systems with
Discontinuities,” Journal of Sound and Vibration, Vol. 44, No. 1, 1976, pp. 15–25.
Beddoes, T. S., “A Synthesis of Unsteady Aerodynamic Effects Including Stall
Hysteresis,” Vertica, Vol. 1, 1976, pp. 113–123.
Bielawa, R. L., “An Improved Technique for Testing Helicopter Rotor-Pylon Aerome-
chanical Stability Using Measured Rotor Dynamic Impedance Characteristics,” Vertica,
Vol. 9, No. 2, 1985, pp. 181–197.
Bielawa, R. L., “Notes Regarding Fundamental Understandings of Rotorcraft Aeroe-
lastic Instability,” Journal of the American Helicopter Society, Vol. 32, No. 4, 1987,
pp. 4–15.
Friedmann, P. P., and Straub, F., “Application of the Finite Element Method to Rotary-
Wing Aeroelasticity,” Journal of the American Helicopter Society, Vol. 25, No. 1, 1980,
pp. 36–44.
Gabel, R., and Capurso, V., “Exact Mechanical Instability Boundaries as Determined
from the Coleman Equation,” Journal of the American Helicopter Society, Vol. 7, No. 1,
1962, pp. 17–23.
Gandhi, F., and Chopra, I., “Analysis of Bearingless Main Rotor Aeroelasticity Using
an Improved Time Domain Nonlinear Elastomeric Damper Model,”Journal of the American
Helicopter Society, Vol. 41, No. 3, 1996, pp. 267–277.
Gangwani, S. T., “Calculation of Rotor Wake Induced Empennage Airloads,” Journal
of the American Helicopter Society, Vol. 28, No. 2, 1983, pp. 37–46.
Gangwani, S. T., “Prediction of Dynamic Stall and Unsteady Airloads for Rotor
Blades,” Journal of the American Helicopter Society, Vol. 27, No. 4, 1982, pp. 57–64.
Gaonkar, G. H., Sastry, V. V. S. S., Reddy, T. S. R., Nagabhushanam, J., and Peters,
D. A., “The Use of Actuator-Disc Dynamic Inflow for Helicopter Flap-Lag Stability,”
Journal of the American Helicopter Society, Vol. 28, No. 3, 1983, pp. 79–88.
Gaukroger, D. R., Skingle, C. W., and Heron, K. H., “Numerical Analysis of Vector
Response Loci,” Journal of Sound and Vibration, Vol. 29, No. 3, 1973, pp. 341–353.
Gladwell, G. M., and Stammers, C. W., “On the Stability of an Unsymmetrical Rigid
Rotor Supported in Unsymmetrical Bearings,” Journal of Sound and Vibration, Vol. 3,
No. 3, 1966, pp. 221–232.
Hodges, D. H., “Vibration and Response of Nonuniform Rotating Beams with
Discontinuities,” Journal of the American Helicopter Society, Vol. 24, No. 5, 1979,
pp. 43–50.
Hohenemser, K. H., and Yin, S.-K., “Some Applications of the Method of Multiblade
Coordinates,” Journal of the American Helicopter Society, Vol. 17, No. 3, 1972, pp. 3–12.
Horvay, G., and Yuan, S. W., “Stability of Rotor Blade Flapping Motion When the
Hinges are Tilted: Generalization of the ‘Rectangular Ripple’ Method of Solution,” Journal
of the Aeronautical Sciences, Vol. 14, No. 10, 1947, pp. 583–593.
Houbolt, J. C., and Reed, W. H., “Propeller-Nacelle Whirl Flutter,” Journal of the
Aeronautical Sciences, Vol. 29, No. 3, 1962, pp. 333–346.
Howlett, J. J., Morrison, T., and Zagranski, R. D., “Adaptive Fuel Control for Heli-
copter Applications,” Journal of the American Helicopter Society, Vol. 29, No. 4, 1984,
pp. 43–54.
Ibrahim, S. R., and Mikulcik, E. C., “A Method for the Direct Identification of Vibration
Parameters from the Free Response,” Shock and Vibration Bulletin, No. 47, Pt. 4, 1977,
pp. 183–198.
Isaacs, R., “Airfoil Theory for Flows of Variable Velocity,” Journal of the Aeronautical
Sciences, Vol. 12, No. 1, 1945, pp. 113–117.
Isaacs, R., “Airfoil Theory for Rotary Wing Aircraft,” Journal of the Aeronautical
Sciences, Vol. 13, No. 4, 1946, pp. 218–220.
Jetmundsen, B., Bielawa, R. L., and Flannelly, W. G., “Generalized Frequency Domain
Substructure Synthesis,” Journal of the American Helicopter Society, Vol. 33, No. 1, 1988,
pp. 55–64.
Kemp, N. H., “On the Lift and Circulation of Airfoils in Some Unsteady-Flow
Problems,” Journal of the Aeronautical Sciences, Vol. 19, No. 10, 1952, pp. 713, 714.
Kennedy, C. C., and Pancu, C. D. P., “Use of Vectors in Vibration Measurement and
Analysis,” Journal of the Aeronautical Sciences, Vol. 14, No. 11, 1947, pp. 603–625.
Kuczynski, W. A., Cooper, D. E., Twomey, W. J., and Howlett, J. J., “The Influence of
Engine/Fuel Control Design on Helicopter Dynamics and Handling Qualities,” Journal of
the American Helicopter Society, Vol. 25, No. 2, 1980, pp. 26–34.
Leishman, J. G. and Beddoes, T. S., “A Semi-Empirical Model for Dynamic Stall,”
Journal of the American Helicopter Society, Vol. 34, No. 3, 1989, pp. 3–17.
Loewy, R. G., “Helicopter Vibrations: A Technological Perspective (The AHS
Alexander A. Nikolsky Honorary Lecture),” Journal of the American Helicopter Society,
Vol. 29, No. 4, 1984, pp. 4–30.
Loewy, R. G., “A Two-Dimensional Approximation to the Unsteady Aerodynam-
ics of Rotary Wings,” Journal of the Aeronautical Sciences, Vol. 24, No. 2, 1957,
pp. 81–92.
Loewy, R. G., Rosen, A., and Mathew, M. B., “Application of the Principal Curva-
ture Transformation to Nonlinear Rotor Blade Analysis,” Vertica, Vol. 11, No. 1/2, 1987,
pp. 263–296.
Luke, Y. L., and Dengler, M. A., “Tables of the Theodorsen Circulation Function
for Generalized Motion,” Journal of the Aeronautical Sciences, Vol. 18, No. 7, 1951,
pp. 478–483.
Lytwyn, R. T., Miao, W.-L., and Woitsch, W., “Airborne and Ground Resonance of
Helicopter Rotors,” Journal of the American Helicopter Society, Vol. 16, No. 2, 1971,
pp. 2–9.
MacFarlane, A. G. J., and Postlethwaite, I., “The General Nyquist Stability Criterion
and Multivariable Root Loci,” International Journal of Control, Vol. 25, 1977, pp. 81–127.
McCarthy, T. R., and Chattopadhyay, A., “A Coupled Rotor/Wing Optimization Pro-
cedure for High Speed Tilt-Rotor Aircraft,” Journal of the American Helicopter Society,
Vol. 41, No. 4, 1996, pp. 360–369.
Miller, R. H., and Ellis, C. W., “Blade Vibration and Flutter,” Journal of the American
Helicopter Society, Vol. 1, No. 3, 1956, pp. 19–38.
Nagy, E. J., “Improved Methods in Ground Vibration Testing,” Journal of the American
Helicopter Society, Vol. 28, No. 2, 1983, pp. 24–29.
Nelson, R. B., “Simplified Calculation of Eigenvector Derivatives,” AIAA Journal,
Vol. 14, No. 9, 1976, pp. 1201–1205.
Ormiston, R. A., and Hodges, D. H., “Linear Flap-Lag Dynamics of Hingeless Heli-
copter Rotor Blades in Hover,” Journal of the American Helicopter Society, Vol. 17, No. 2,
1972, pp. 2–14.
Paul, W. F., “A Self-Excited Rotor Blade Oscillation at High Subsonic Mach Numbers,”
Journal of the American Helicopter Society, Vol. 14, No. 1, 1969, pp. 38–48.
Penrose, R., “On the Generalized Inverse of a Matrix,” Proceedings of the Cambridge
Philosophical Society, 1955, pp. 406–413.
Peters, D. A., “Flap-Lag Stability of Helicopter Rotor Blades in Forward Flight,”
Journal of the American Helicopter Society, Vol. 20, No. 4, 1975, pp. 2–13.
Pitt, D. M., and Peters, D. A., “Theoretical Prediction of Dynamic-Inflow Derivatives,”
Vertica, Vol. 5, No. 1, 1981, pp. 21–34.
Popelka, D. A., Sheffler, M. W., and Bilger, J. “Correlation of Test and Analysis for the
1/5-Scale V-22 Aeroelastic Model,” Journal of the American Helicopter Society, Vol. 32,
No. 2, 1987, pp. 21–33.
Proceedings
Beddoes, T. S., “Onset of Leading Edge Separation Effects Under Dynamic Conditions
and Low Mach Number,” Proceedings of the 34th Annual National Forum of the American
Helicopter Society, May 1978.
Bielawa, R. L., “Techniques for Stability Analysis and Design Optimization with
Dynamic Constraints of Nonconservative Linear Systems,” Proceedings of the 12th
AIAA/ASME/AHS Structures, Structural Dynamics and Materials Conference, AIAA Paper
71-388, 1971.
Bielawa, R. L., “Dynamic Analysis of Multi-Degree-of-Freedom Systems Using Phas-
ing Matrices,” Proceedings of Specialists’ Meeting on Rotorcraft Dynamics, NASA Ames
Research Center, 1974.
Rehfield, L. W., “Design Analysis Methodology for Composite Rotor Blades,” Pro-
ceedings of the Seventh DoD/NASA Conference on Fibrous Composites in Structural Design,
AFWAL-TR-85-3094, U.S. Air Force, 1985, pp. V(a)-1–V(a)-144.
Young, D. K., and Tarzanin, R. J. (Jr.), “Structural Optimization and Mach Scale Test
Validation for a Low Vibration Rotor,” Proceedings of the 47th Annual National Forum of
the American Helicopter Society, May 1991.
Reports
Baker, B. C., “Anti-Aliasing, Analog Filters for Data Acquisition Systems,” Microchip
Application Note AN699, 1999.
Bielawa, R. L., Johnson, S. A., Chi, R. M., and Gangwani, S. T., “Aeroelastic Analysis
for Propellers,” NASA CR-3729, 1983.
Bielawa, R. L., “Simplified Dynamic Equations for Ground and Air Resonance,”
Rensselaer Polytechnic Inst., Rotorcraft Technology Center, Troy, NY, Rept. D-90-1, 1990.
Bielawa, R. L., “An Experimental and Analytical Study of the Mechanical Instability
of Rotors on Multiple-Degree-of-Freedom Supports,” Princeton Univ., Princeton, NJ, Dept.
of Aeronautical Engineering Rept. 612, 1962.
Cansdale, R., Gaukroger, D. L., and Skingle, C. W., “A Technique for Measuring
Impedance of a Spinning Model Rotor,” Royal Aircraft Establishment TR-71092, London,
1971.
Castles, W. Jr., and Gray, R. B., “Empirical Relation Between Induced Velocity, Thrust
and Rate of Descent of a Helicopter Rotor as Determined by Wind-Tunnel Tests on Four
Model Rotors,” NACA TN 2474, 1951.
Coleman, R. P., and Feingold, A. M., “Theory of Self-Excited Mechanical Oscillations
of Helicopter Rotors with Hinged Blades,” NACA Rept. 1351, 1958.
Flannelly, W. G., Fabunmi, J. A., and Nagy, E. J., “Analytical Testing,” NASA CR-
3429, 1981.
Gangwani, S. T., “Synthesized Airfoil Data Method for Predicition of Dynamic Stall
and Unsteady Airloads,” NASA CR-3672, 1983.
Greenberg, J. M., “Airfoil in Sinusoidal Motion in a Pulsating Stream,” NACA TN-
1326, 1947.
Halfman, R. L., Johnson, H. C., and Haley, S. M., “Evaluation of High Angle-of-Attack
Aerodynamic—Derivative Data and Stall-Flutter Prediction Techniques,” NACA TN-2533,
1951.
Ham, H. D., “Helicopter Blade Flutter,” AGARD Rept. 607 (revision of Pt. III, Chap.
10 of the AGARD Manual of Aeroelasticity), 1973.
Hammond, C. E., and Doggett, R., “Demonstration of Subcritical Damping by
Moving-Block/Randomdec Applications,” NASA SP-415, 1976, pp. 59–76.
Hodges, D. H., and Dowell, E. H., “Nonlinear Equations of Motion for the Elastic
Bending and Torsion of Twisted Nonuniform Rotor Blades,” NASA TN D-7818, 1974.
Hodges, R. V., Nixon, M. W., and Rehfield, L. W., “Comparison of Composite Rotor
Blade Models: A Coupled-BeamAnalysis and an MSC/NASTRAN Finite-Element Model,”
NASA TM-89024, 1987.
Houbolt, J. C., and Brooks, G. W., “Differential Equations of Motion for Combined
Flapwise Bending, Chordwise Bending and Torsion of Twisted Non-uniform Rotor Blades,”
NACA Rept. 1346, 1958.
729
fixed points, 261, 264 forward flight, 399, 447–448, 496, 705
fixed wing aerodynamics, 316
flutter calculations, 305–306 forward translational flight, 439
vs rotary wing, 1–2 forward whirl, 397, 399
flap-damping coefficient vs frequency Fourier amplitudes, 388
ratio, 468 Fourier analysis, 349, 417
flap-lag stability characteristics in Fourier coefficients, 15–17
hingeless rotors, 378–379 Fourier methods, 14–19
forward flight, 378–379 Fourier series, 17, 145, 349–350, 389
hover condition, 374 Fourier transform, 18–19, 287, 470
nonlinear effects, 379 free-free systems, 31
flapping excitation of rotor blade, 214 frequency and period relationship, 10
flatwise bending stiffness, 618 frequency domain
flatwise mass moment of inertia, 617 analysis, 2
flexibility ratio, 153 formulation, 403
flexible root constraint, 112 methods, 305–309
flexible rotor blade, flutter, 488–492 substructure synthesis, 677–683
flexible shaft elastically deformed in frequency placements, 240
transverse direction, 150 frequency ratio, 262–264
flexible supports with mass vs flap-damping coefficient, 468
eccentricity, 152 vs pitch-damping coefficient, 468
flight acceleration data, 249 frequency response
Floquet stability analysis, 653 characteristics, 285
Floquet theory, 318–319, 323, 332 damped case, 29
Fluidlastic, 588–589 undamped case, 28–29
flutter instability characteristics, frequency response function (FRF), 19,
rigid-blade rotors, 394 226, 230, 275, 286–287, 327, 332,
flywheels, flexible shaft connected, 31 417
focused pylon vibration-alleviation alternate expressions, 287–288
device, 270 calculation, 286–288
folded Abbé prism, 571–572 linear single-degree-of-freedom
force balance for forward whirling or system, 13
whipping shaft, 379 frequency scaling parameter, 428
force boundary conditions, 115 frequency spectrum of rotorcraft dynamic
force displacement for connecting related phenomena, 626
beam, 158 frequency sweep, 170
force integration method, 189–205, 209, frequency sweep excitation, 291
211 frequency-temperature correspondence
force load distribution, 106, 120 principle, 581
force phasing matrices, 297, 310–314, Froude scaling, 560, 563–564
496 fuel control system see combined
energy flow paths, 313–314 rotor/drive/engine/fuel control
mathematical implementation, 312–313 system
force-summation method, 190 functions, tabulations, 168
force transducers, 565 fuselage
forced motion dynamic equations, 151 dynamic characteristics, 226, 275
forced response equation, 192 dynamic modifications, 246–257
forced responses ground resonance analysis, 405
FEM, 230 impedance, 230
single-degree-of-freedom equations, 205 inertia, dynamic terms caused by, 692
forcing frequency and damping, 13 mobility, 224, 230, 232, 234, 239,
FORTRAN, 651 576
rotor lateral edgewise excitation, 406, 693 effect of chordwise mass offset, 360
rotor lift, 563 effect of collective angle, 359
rotor lift coefficients, 563 effect of mass ratio (Lock number), 359
rotor longitudinal edgewise excitation, effect of rotor type, 358
406, 693 effect of solidity, 360
rotor mass elements, dynamic terms caused rotor whirl trend studies, 395
by, 687 rotor-wing combination, 704–706
rotor modes, 138–139, 387–389 rotors
rotor-nacelle system center of gravity, 388–389
dynamic components defining flutter, cosine torsion response, 694
402 unsteady aerodynamics and flutter,
dynamic subsystems, 401 447–527
elastomechanical description, 390–391 see also combined
multiple-degrees-of-freedom rotor/drive/engine/fuel control
frequency-domain method, 400–404 system; two-bladed rotor
substructure, 705 Routh array, 303–304, 411
whirl flutter, 389–404 very lightly damped systems, 305
wing flexibility effect, 396–397 Routh criterion, 303, 356, 358, 367, 408,
rotor pitching moment, 436 493
rotor pitchwise flatwise excitation, 694 Routh–Hurwitz criterion, 303–305, 394
rotor-pylon systems Runge–Kutta method, 324–325, 417,
aerodynamic description using 538–539
quasisteady theory, 391–393
comprehensive analyses, 400–404 St. Venant torsion theory, 606
mechanical and aeromechanical scalar formulation, 499
instabilities, 387–446 scalar matrix, 43
passive elastomechanical coupling, 706 scalar multiplication, 46, 54
stability characteristics, 394–396 scalar triple product, 56
rotor rolling moment, 436 scale factors, 429, 560, 564
rotor rollwise flatwise excitation, 694 scaling formulas for aeroelastic quantities,
rotor sine torsion response, 694 561
rotor speed, 316 scaling laws, 557–560
characteristics, 73 scatter factors for aerospace metals, 295
limiting case, 74 Sears function, 455–457
vs initial curvature, 152 second-order backward-differencing, 537
vs mass unbalance, 151 secondary aeroelastic responses, 244
zero, 76 secondary mass, 239
rotor stiffness and damping, dynamic terms secondary section components, 616–617
caused by, 690–670 section centers, 617
rotor stiffness ratio, 154 section moment, pitch rate effects, 347
rotor systems, elastomechanical semianalytical method, 193
properties, 317 semicanonical form, 309
rotor testing see model rotor testing semipositive-definite matrices, 315
rotor thrust, 436 shafts
rotor unsteady airloads theories, configuration with significant distributed
classification, 450–451 inertia, 167
rotor wake, 502–503 critical speeds, 149–162
division into upper and lower parts, 221 density, 167
geometry, 496 with significant mass, 167–168
rotor weaving instability, 352–362, 386 shake testing, 275–292, 677
effect of chordwise aerodynamic offset, advanced measuring devices, 282
361 basic single-point, 275–276