Sie sind auf Seite 1von 377

Erwin Kramer

Dynamics of Rotors
and Foundations

With 304 Figures

Springer-Verlag
Berlin Heidelberg GmbH
Prof. Dr. rer; nat. Erwin Krii.mer
IsselstraJ3e 12
64297 Dannstadt-Eberstadt
Gennany

ISBN 978-3-662-02800-1 ISBN 978-3-662-02798-1 (eBook)


DOI 10.1007/978-3-662-02798-1

This work is subject to copyright. AII rights are reserved, wether the whole or part of the
material is concemed, specifically the rights oftranslation, reprinting, reuse ofillustrations,
recitation, broadcasting, reproduction on microfilm or in any other way, and storage in data
banks. Duplication ofthis publication or parts thereofis permitted only under the provisions
ofthe German Copyright Law of September 9, 1965, in its current version, and permission
for use must always be obtained from Springer-Verlag Berlin Heidelberg GmbH .
Violations are liable for prosecution under the German Copyright law.

Springer-Verlag Berlin Heidelberg 1993


Originally published by Springer-Verlag Berlin Heidelberg New York in 1993
Softcover reprint ofthe hardcover Ist edition 1993

The use of general descriptive names, registered names, trademarks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore frec for general use.

Typesetting: Carnera-ready by author;

6113020 543210 Printed on acid-frec paper


Foreword

My acquaintance with the Author began some 30 years ago while he was with
Brown Boveri in Mannheim. Since then we have become close colleagues and it
was a natural consequence to cooperate in writing a book on rotordynamics -
he as an expert with a wealth of industrial experience and I as translator from
German into English, with a background in industry and in university teaching
and research into the vibration of rotating systems.
The outcome is a book with a traditionally German emphasis on a wide and
detailed coverage of the dynamics of rotors and foundations. The bibliography
is extensive and the author is not content to simply refer to this, but to give his
own very lucid explanations of the phenomena involved. This is backed up by
detailed calculations and graphs which give the engineer a thorough appreciation
of the different factors affecting the rotordynamic performance of machines.
In addition, tools of the trade such as finite element modelling and stability
analysis are covered in detail and the result is a book which I am sure will be
regarded as an essential volume on the book shelf of any student or engineer
who has a need to consider machine vibration.

Southampton, England Roy Holmes


Preface

In the last decades rotordynamics has developed into a wide area of speciality.
Numerous investigations have appeared in the literature and even for the spe-
cialist it is difficult to appreciate the whole picture. A choice of the essential
results and their correlation in as understandable a form as possible is therefore
a prime requirement. It is hoped that this will be accomplished in this book
through the knowledge and experience gained the author, over his working life.
The book is written for anyone who is concerned with rotordynamics,
whether by choice or necessity. It is recommended for lecturers and students
and for the manufacturers and users of rotating machines and their founda-
tions. On the one hand the mechanics and mathematical derivations are dealt
with in detail and, on the other, an attempt is made to simplify the under-
standing of the results by presenting numerous figures and corresponding text
for readers with little expert knowledge.
The contents of this book are described in detail in Chapt. 1. They consist
of the background literature, work in the specialist area of machine dynamics
in the Technical University of Darmstadt and of the author's own personal
experiences in industry, at university and as a specialist in the field. The cited
literature has been chosen from personal knowledge and judgement and the
reader's indulgence is asked if the choice in some cases appears inadequate.
The object of writing this book in English could only be realized because
Professor R. Holmes, a well-known researcher in rotor dynamics, offered to
translate the German manuscript. I would like to thank Professor Holmes espe-
cially for his unselfish readiness, for his critical appraisal and translation of the
manuscript, as well as for many useful pieces of advice.
Dr. Ing. H. D. Klement has also contributed extensively to the contents of
this book. He conveyed to the author in many discussions his knowledge and
experience on the subject of writing and using computer program MADYN,
described in Chapt. 25. For this and for the calculation of most of the examples,
the author would like to thank Dr. Klement most warmly. Furthermore, thanks
are due everyone who read parts of the manuscript and gave advice, especially
Dr. Ing. L. Eckert, Prof. H. Freund, Prof. J. Glienicke, Prof. H. J. Thomas and
Dr. Ing. J. Schmied.
Many questions and acknowledgements originate from the author's earlier
employment at BBC, now ABB Zurich, and from discussions with experts of
this company. The author wishes to thank former colleagues for their assistance
and ABB for information on some of the given examples and for authorization
to publish the results.
VIII

Gratitude is expressed to Mrs. M. Mayer for the careful typing of the book,
to Dipl. Ing. D. Glaser for drawing the figures and to Mrs. Dipl. Ing. E. Grunau
for final reviewing of the manuscript. Finally, the author would like to thank the
representatives of Springer- Verlag for their pleasant and efficient cooperation.

Darmstadt, Germany Erwin Kramer


Table of Contents

I Fundamentals 1
1 Introduction ........ 3

2 The one Degree of Freedom System 7


2.1 Equation of Motion . 7
2.2 Natural Vibration .. 8
2.3 Harmonic Excitation 11
2.4 Random Excitation . 17

3 Shaft with Central Disc 19


3.1 Shaft with Rigid Bearings, Jeffcott Rotor. 19
3.2 Startup, Run Down . . . . . . . . . . 24
3.3 Jeffcott Rotor with Flexible Bearings 29
3.4 Jeffcott Rotor with Bearing Damping 33

4 Shaft with Non Central Disc 37


4.1 Equations of Motion 37
4.2 Natural Vibrations ... 41
4.3 Unbalance Vibrations .. 52
4.4 Harmonic Excitation in one Direction . 56
4.5 Shaft with Flexible Bearings 59

5 The Short and the Long Rotor 65


5.1 Unbalance Moment . . . 66
5.2 Natural Frequencies . . . 68
5.3 Unbalance Vibrations .. 71

6 Oil-Film Bearings ...... 77


6.1 Hydrodynamic Bearing Theory 77
6.2 Short Bearing Theory. . . . . . 83
6.3 Static and Dynamic Properties 89
7 Rotors with Oil-film Bearings 97
7.1 Equations of Motion .. 97
X Table of Contents

7.2 Stability . . . . . . . . . . 99
7.2.1 Rigid Rotor . . . . 99
7.2.2 The Jeffcott Rotor 104
7.3 Unbalance Vibrations. 108
7.4 Summary 113

8 Vertical Rotors 115


8.1 Tilt-pad Bearings . 115
8.2 Vibrations..... 123

9 Rolling-element Bearings 129


9.1 Deformation of the Rolling-element 130
9.2 Stiffness of Rolling-element Bearings 132
9.3 Lateral Stiffness, Critical Speeds . . . 135
9.4 Consequences of Element Rotation . 138
9.5 Damping of Rolling-element Bearings . 140

10 Shaft Seals . . . . . . . . . . 143


10.1 Restoring Force . . . . . 144
10.2 Jeffcott Rotor with Seals 146
10.3 Smooth Seals . . . . 152
10.4 Labyrinth Seals . . . 156
10.5 Floating Ring Seals. 157

11 Steam Whirl . . . . . . . 161


11.1 The Steam Whirl Force. 161
11.2 Jeffcott Rotor with Steam Whirl . 164

12 Internal Damping . . . . . . . . . . . 169


12.1 Jeffcott Rotor with Internal Viscous Damping 169
12.2 Damping due to the Shaft Material . . 174
12.3 Damping from Assembly Components . 177
12.4 Conclusion and an Example 184

13 Non-circular Shafts . . . . . . . 191


13.1 Non-circular Jeffcott Rotor 192
13.2 Supplementary Comments . 199

II Rotors with Several Bearings 201


14 Computer Model . . 205

15 Influence Coefficients 207


Table of Contents XI

16 Equation of Motion 211


16.1 Statics . . ... . 211
16.2 Dynamics . . . 217
16.2.1 lligid Masses 217
16.2.2 Beam Element . 221
16.2.3 Flexible Couplings 224
16.2.4 Bearing Supports . 227
16.2.5 Shaft Seals, Steam Whirl . 228
16.3 Reduction in the Number of Coordinates 228
16.4 Discussion of the Equation of Motion 230

17 Results . . . . . . . . . . . . . . . . . . . 233
17.1 General Discussion of Natural Vibrations. 233
17.2 Rotors with two Bearings. . . 236
17.3 Rotors with Several Bearings. 247

18 Static Bearing Loads . . . . . . . . 255

III Rotor-foundation Coupling 257


19 Foundation . . . . . . . . . . . 261
19.1 Calculations . . . . . . . . 262
19.1.1 Ground Dynamics. 262
19.1.2 lligid Foundation Block 266
19.1.3 Flexible Foundation. 269
19.2 Results . . . . . . . . . . . 277
19.2.1 lligid Foundation . 277
19.2.2 Frame Foundation 283
19.2.3 Table Foundation . 283
19.2.4 Spring Foundation 287
19.2.5 Reinforced Concrete and Steel .. 287

20 Rotor and Foundation 293


20.1 Calculations . . . 294
20.2 Results. . . . . . 296
20.2.1 Small Machine Sets with Frame Foundations 298
20.2.2 Large Turbine Sets 305
20.2.3 Conclusion. . . . . . . . . . . . . . . . . . . . 311

IV Further Features 315


21 Rough Calculations . . . . . . . . . . 317
21.1 Natural Frequencies and Critical Speeds 317
XII Table of Contents

21.2 Resonance Amplitudes 322


22 Bending Stresses in Rotors 325

23 Cracked Rotors . . . . . . . 329


23.1 Vibrations of a Rotor with a Transverse Crack . 329
23.1.1 Jeffcott Rotor . 329
23.1.2 General Rotor . . 334
23.1.3 Results. . 335
23.2 Case Studies . 342
23.3 Summary . . . . 345

24 Solution of the Equation of Motion 349


24.1 Homogeneous Equation, Natural Frequencies. 349
24.2 Harmonic Excitation . . . . . 353
24.3 General Excitation . . . . . . 353
24.4 Modal Method of Calculation 353
24.4.1 Symmetric Matrices . 354
24.4.2 Harmonic Excitation . 356
24.4.3 Reduced Modal Calculation 358
24.4.4 Transfer Function . . . 360
24.4.5 Unsymmetric Matrices 360
24.5 Rayleigh Quotient . 362

25 Program MADYN . . . 365


25.1 Review. . . . . . . 365
25.2 Building the Model 367
25.3 Eigenvalue Algorithm. 368

References 371

Index . . . 381
Part I

Fundamentals
1 Introduction

Rotors of machines have, while in operation, a great deal of rotational energy,


and a small amount of vibrational energy. The purpose of rotordynamics as
a subject is to ascertain how this latter can be kept as small as possible. Its
significance can be underlined by the following example.
The rotational energy of a rotor turning with angular velocity n is the
same as the kinetic energy of a non-rotating rotor with translational motion of
velocity v = nip where ip is its radius of gyration. For example, for a rotor with
radius of gyration 200 mm and 3000 rpm, the velocity v is 63 m/s or 226 km/h.
The typical velocity of translational vibration, on the other hand is measured
in mm/s, a factor of about 10- 4 smaller than the velocity v.
Engineering components concerned with the subject of rotordynamics are
the rotors of machines, especially of turbines, generators, motors, compressors,
blowers and the like. In operation the rotor undergoes bending and torsional
vibration. In this book only bending vibrations will be considered, and these
will be assumed to be linear. The shaft will be assumed straight, and so, for
example, crankshafts of piston machines will not be considered.
Similarly, the subject of balancing will not be covered. However, this book
gives the essential groundwork relevant to that subject.
The vibration of a rotor depends upon its geometry and the type of support,
as well as on the excitation forces. The vibrating rotor also excites its foundation.
Similarly the vibrations of a rotor are influenced more or less by the foundation.
The interplay between the two is illustrated in this book, to elucidate whether a
rotor and its foundation should be considered separately or as a coupled system.
The book consists of four parts:

Fundamentals

- Rotors with several bearings

Rotor-foundation coupling

Further features.

In rotordynamics a remarkable amount can be explained by the dynamics of a


one degree of freedom oscillator and so the basic relationships governing this
model are covered at the outset. The following eleven chapters of the first part
are concerned with the dynamics of a leffcott rotor. This model, introduced in
4 1 Introduction

1895 by Fappl [1], was named after Jeffcott, because in 1919 he explained the
science of rotordynamics in a graphic and illuminating way, still used today [2].
Developments since the earliest days can be summarized as follows:
1869 Rankine [3] remarked on the existence of critical speeds. He described a
critical speed as " ... the limit of speed below which centrifugal whirling
is impossible ... " which is obviously incomplete.

1883 Laval built the turbine named after him and solved the problem of quiet
operation - his operating speed ranged up to 40,000 rpm - by using an
extremely flexible shaft. He aimed at the self-centring of the disc above
the critical speed, a phenomenon which he intuitively recognized.

1895 The fundamental investigations of Dunkerley [4] and Fappl [1] appeared.
It thus became recognized that a shaft has several critical speeds and that
under certain circumstances these were the same as the natural frequencies
of a non-rotating shaft. In order to calculate the critical speeds of cylindri-
cal shafts with several discs and bearings the general theory of Reynolds
[4] was applied. The gyroscopic effect was also considered, together with
its dependence on speed. The required solution of the frequency equation
was at that time only possible for simple models. Dunkerley found, as a
result of numerous measurements, the relationship known today by that
of Southwell, by which the first critical speed can be calculated, even for
complicated cases.
Even with the general knowledge of critical speeds, the shaft behaviour at any
general speed was still unclear but more was learnt from the calculations of
unbalance vibrations, as given by Fappl [1]. Also the behaviour of Laval rotors
at high speed was confirmed by his theory.
1910 Stodola presented in [5] a graphical procedure to calculate critical speeds,
which was widely used until, with the advent of the digital computer, it
was replaced by the transfer matrix method of Myklestad and Prohl [6],
[7].
The gyroscopic moment has the effect of making the natural frequencies de-
pendent on rotor speed, while at the same time doubling their number. This is
shown in Chap. 4, using the model of a Jeffcott rotor with a single non-central
disc. Along with other parameters, the ratio of diametral to polar moment of
inertia plays an important role. This is elucidated in Chap. 5 for a Hlindrical
rotor, whose length can be varied.
The oil-film of a fluid-film bearing acts like a spring-damper system and
it influences the critical speeds and limits the vibration amplitude at resonance
by virtue of this damping. More important, however, is the fact that the oil-
film can cause an instability known as oil-film whirl. The fundamentals of the
dynamics of fluid-film bearings are described in Chap. 6, while in Chap. 7, the
5

behaviour of a Jeffcott rotor with fluid-film bearings is explained. In Chap. 8 a


vertical rotor is considered.
Rolling element bearings act like springs. They are essentially stiffer than
fluid-film bearings and have almost no damping. Neither do they lead to insta-
bility (Chap. 9).
There is an extremely comprehensive literature on the role of oil-film bear-
ings in rotordynamics, which can only be touched on here. The developments up
to 1957 were largely due to Newkirk [8] who explained them in a very detailed
and graphic way. Lund [9] gives an overview of the field up to 1987 and the
latest results can be found in ref. [10], amongst others.
Shaft seals have a similar effect as fluid-film bearings. They influence the
critical speeds, can provide damping or on the other hand cause instability.
These are described in Chap. 10. Some of the first investigations were carried
out in 1955 by Lomakin [11]. Since then shaft seals have acquired a significant
role in their effect on rotordynamics, as the extensive literature shows. This will
only be referred to briefly in this book but recent publications and bibliography
can be found in reference [12].
Instability from fluid-film bearings and shaft seals arises from the fact that,
during radial displacement of a rotor, a restoring force is produced, which has
a component at right angles to this displacement. Such a mechanism is also
possible for a rotor with blades, as a result of variable leakage around the blade
tips. This was found by Thomas [13], when instabilities ofturbo-machines could
not be explained by contemporary theory. Further details and bibliography on
this subject can be found in Chap. 11.
The phenomenon of instability was described in detail by Newkirk [14] in
1924, whose interest was in turbomachines, in particular blast-furnace blowers.
At first it was thought that the cause was internal friction due to shrink fits
on the shaft and a theory was developed to explain this also, together with ex-
perimental verification (Newkirk, Taylor, Kimball [15], [16]). The actual cause
of instability, in fact, lay in the oil~film bearings [8]. The error in understanding
arose due to the fact that rolling-element bearings were used in the test facility
designed to investigate the problem. Notwithstanding, in the following years it
was established that in a few cases, internal friction or damping could indeed
be a cause of instability. The designer must thus be aware of these possibilities
and design machines accordingly and for this reason the fundamentals of this
subject are described in Chap. 12. Investigation of a recent case (example 12/1)
underlines the importance of these causes.
The last chapter of the first part describes the vibration of a shaft with un-
equal bending stiffnesses, a so-called non-circular shaft. The special phenomena
encountered here are particularly important for two-pole rotors of turbogener-
ators.
The basic model used for the first part of this book is the J effcott rotor and
with this simple model most of the important results can be achieved analyti-
6 1 Introduction

cally. With a complicated model as subsequently used this is no longer possible.


For calculation purposes we then have the choice of the transfer-matrix method
[6], [7] or the finite-element method. The transfer-matrix method is the simpler,
but is not so widely applicable as the finite-element method.
The second part of this book is concerned with shafts on several bearings
and a simple model for the bearing pedestal, for which the transfer-matri;x:
method suffices. The finite-element method is then introduced, since, with the
use of today's digital computers, the advantages of the transfer-matrix method
have declined in significance.
Part two begins with a short description of mathematical models. In the
fifteenth chapter the fundamental ideas of nodes and their forces and displace-
ments (coordinates) are explained by the use of a simple beam model and two
kinds of influence coefficients are defined.
The finite-element method consists in the determination of the equation of
motion (Chap. 16) and its solution (Part IV, Chap. 24). The equation of motion
is a matrix equation. The formulation of the matrices of the model from the
element matrices (Sect. 16.2) is shown in Sect. 16.1 for a static system. The
dynamic matrices are then obtained in a similar way. The vibration behaviour
of shafts with several bearings is dealt with in Chap. 17, using some typical
examples.
Part three deals with coupled rotor-foundation systems. First, in Chap. 19,
the foundation is considered without and with a rigid machine mounted on it.
By comparison in Subsect. 19.2.5 examples of a reinforced concrete foundation
and a steel foundation are taken, together with the dynamics of the rotor.
For the coupled system (Chap. 20) it is appropriate to distinguish between
small machines and large turbo-sets. At the end of this part some comments
are made on the basis of results taken from several examples, as to whether it
is better to calculate the individual systems separately or as coupled together.
Part four begins with a short chapter on procedures involving rough calcu-
lations. Although it may appear to be rather a bold step to carry out rough
calculations with the models considered here, there are several reasons why it
is necessary to acquire such experience. Central to this is the understanding
achieved from the leffcott rotor and the proportionality factors relating such
results with more detailed and exact calculations.
A significant point in the judgement of the importance of vibrations is
the resulting stressing of a shaft whose kinetic fundamentals are compiled in
Chap. 22. In this connection the dynamics of cracked rotors is also of interest
and up-to-date knowledge is presented in Chap. 23.
Finally, Chap. 24 is concerned with the solution of the equations of motion
and the book closes with a description of computer program MADYN, with
which most of the examples have been calculated.
2 The one Degree of Freedom System

The simplest rotordynamic model is the Jeifcott rotor discussed in Chap. 3. Its
behaviour allows it to be considered essentially as having one degree of freedom.
The properties of a one degree of freedom model will therefore be considered
first.

2.1 Equation of Motion

The model shown in Fig. 2.1 represents this simple system. It consists of a mass,
a spring and a viscous damper. The spring force is proportional to displacement
z and the damping force to velocity a:. The spring and damper are therefore
termed linear.
The equation of motion of this model is

mz + da: + kz = F(t) (2.1)

where m is the mass


d the damping coefficient
k the stiffness coefficient
F(t) the excitation force
t time
z displacement of the mass from the static equilibrium position.
The complete solution z(t) consists of the general solution of the equa-
tion for F(t) = 0, the homogeneous equation, and the particular solution for
F(t) f:. O. In the following the general solution is considered and then the parti-
cular solution with harmonic excitation. Finally a procedure will be given for
integrating equation (2.1) when the excitation force F(t) is random in nature.

t F(t)
x(t)

Fig. 2.1. One degree of freedom system


8 2 The one Degree of Freedom System

2.2 Natural Vibration

Putting F(t) = 0 in equation (2.1) gives


m:il+d:i:+kz = 0

Dividing by m, gives
(2.2)
where
d
(2.3)

(2.4)

Assuming a solution of the form

z(t) = Ce>.t (2.5)


one obtains the equation

and, for C f:. 0, the characteristic equation

This has the roots (eigenvalues)

A,A* = -5 J5 2 - w~ (2.6)
The quantity 5 in equation (2.:l) is a measure of the damping and is sometimes
called the decay constant. It has the dimension of wn For 0 < 5 < Wn one
speaks of underdamping, in which case

(2.7)
where
Wd = Jw~ - 52 (2.8)
With underdamping the general solution of the homogeneous equation is

(2.9)

where C1 and C2 are constants. These must be of such values that both the left
and right hand sides of equation (2.9) are real.
2.2 Natural Vibration 9

This is achieved by putting

(2.10)

in which A and B are real constants.


Inserting equations (2.10) into equation (2.9) gives finally

:c(t) = e-6t(Acoswdt + B sinwdt) = ze- 6t sin(wdt + r,o) (2.11)

where
z = v'A2 + B2 (2.12)
A
r,o = arctan B . (2.13)

For the case without damping, 5 = 0, equation (2.11) becomes

(2.14)

This is the function which is plotted in Fig. 2.2, and gives the so-called natu-
ral vibration, consisting of a harmonic displacement :c. The value of Wn is the
natural circular frequency of the simple system without damping. With period
211" and periodic time Tn then wnTn = 211". Hence

T.n_- 211" (2.15)


Wn

and
1 Wn
fn ,- -Tn--- 211"
- (2.16)

t
x

t--
Wn

Fig. 2.2. Natural vibration without damping


10 2 The one Degree of Freedom System

is the simple natural frequency.! The constants A and B in equation (2.14) are
determined from the initial conditions. With z(t = 0) = Zo and z(t = 0) = zo,

:Co
A=zo , B=- . (2.17)
Wn

The natural vibration with damping is described by equation (2.11). This


equation consists of a term e- 6t decreasing with time and a harmonic term.
The whole can be shown as a damped harmonic vibration (Fig. 2.3). Its natural
circular frequency is Wd from equation (2.8).
Given the decay constant 5, one can express the value of damping as the
damping ratio D or as the logarithmic decrement {}. The damping ratio is defined
as
(2.18)

It can also take the forms


d d dwn
D-------- (2.19)
- 2mwn - 2Vkffl - 2k

The logarithmic decrement is defined as

(2.20)

where Zn is the value of the displacement from equation (2.11) at time to and
:en+! is the value at time to + T d , where

(2.21)

t-

Fig. 2.3. Natural vibration with damping

lWhile the natural circular frequency Wn is preferred by theorists, the simple natural
frequency In is mostly used by practitioners.
2.3 Harmonic Excitation 11

Td is the periodic time of the damped vibration. Using the above equations, one
obtains
(2.22)
and, for D < 1,
(2.23)

2.3 Harmonic Excitation

After the natural vibration, forced vibrations under harmonic excitation will
now be considered. First the amplitude of the exciting force will be regarded as
constant and later it will be assumed to be proportional to the square of the
exciting frequency, as is the case for unbalance excitation. Finally, the case of
harmonic motion of the support will be considered.
For constant force-amplitude

F(t) = F coswt .

In order to solve the equation of motion (2.1) the displacement is represented


by the complex quantity
z = z + jy
and the corresponding F(t) by

F(t) = F (coswt + j sinwt) = Feiwt


Hence, from equation (2.1)

mz + di + kz = Feiwt (2.24)

with the particular solution


z(t) = Ce iwt (2.25)
and
C= F (2.26)
k -mw 2 +jdw
The displacement z(t) = Re{z(t)} leads to,

z(t) = $ cos (wt - e) (2.27)


with
(2.28)
12 2 The one Degree of Freedom System

and
dw
= arctan k -mw 2 (2.29)

The result is that, with harmonic excitation of frequency w, the mass moves with
harmonic motion of the same frequency. The displacement follows the force by
the phase angle .
For w = 0, one has a static force of value F. Here the displacement is given
by
F
x. = - (2.30)
k
The amplitude z divided by x. is called the magnification factor V. With the
frequency ratio
w
"1=- (2.31)
Wn

and the damping ratio D from equation (2.18) then using equation (2.28), the
magnification function

(2.32)

and from equation (2.29), the phase angle

2D"I
= arctan-- (2.33)
I _1J2

V and are shown in Figs. 2.4 and 2.5 as functions of forcing frequency ratio
and with different damping ratios. The magnification function is also known as
the resonance curve. Each of these resonance curves starts at unity and proceeds
towards zero for large values of "I. For D :s; 1/../2 = 0.707 the curves have a
maximum of
v. _ 1 (2.34)
max - 2DVI _ D2
at a frequency ratio
1Jmax = VI - 2D2 (2.35)
For small damping D <K 1 then

1
Vrnax ~
2D (2.36)
1/max ~ 1 .

The expression 1/2D is called the resonance sharpness or Q-factor.


Hence
1 zmax
Q = - ~ Vmax = - - (2.37)
2D x.
2.3 Harmonic Excitation 13

Fig. 2.4. Magnification function V(1], D)

o0 2
71-
Fig. 2.5. Phase angle e( 1], D)
14 2 The one Degree of Freedom System

From Fig. 2.5 the phase angle lies between 0 and 7r and for 1/ = 1, that is
W W n , e is equal to 7r /2, independent of damping ratio.
=
Fig. 2.4 shows that the resonance curves become broader with increase in
damping. This breadth thus represents the damping value. It can be shown that,
for D ~ 1, the expression

D~W2-WI = Llw
(2.38)
2wmax 2wmax

holds, with WI and W2 the forcing frequencies at which Z = zmax/ v'2 ~ 0.707zmax
(Fig. 2.6). The values WI and W2 are referred to as the 'half power points'. For
D = 0.1, the exact value of the right-hand side of expression (2.38) is 0.103.
There is thus an error of 3 % and so this expression is sufficiently accurate in
practice for low damping ratios. Expression (2.38) can also be used to find the
damping ratio of an actual system from a measured response curve.
In the case of unbalance excitation, one can complete the system of Fig. 2.1
with an unbalance exciter, as shown in Fig. 2.7. It consists of rotating masses
mo/2, having angular velocity wand produces an excitation force

With the total mass m = mI + mo the equation of motion is


m:e + di; + k;e = moew 2 cos wt . (2.39)

The particular solution follows directly from the solution with constant force
amplitude, if one replaces F by moew 2 One then obtains using equations (2.27)
and (2.28) the solution
;e(t) = z' cos(wt - e) (2.40)

t
x
~-+--+- 0.707 xmax

w---
Fig. 2.8. Resonance curve and half value width ..1w
2.3 Harmonic Excitation 15

m,

k k
2" 2"
Fig. 2.7. Simple system with unbalance excitation

where

or
~I
x = mO Ve I
(2.41 )
ml +mO
where 2
V'- 1/ (2.42)
- V(1 _1/ 2)2 + (2D1/)2
One thus has a different magnification function while the phase angle e is the
same as for constant force amplitude. V'is shown in Fig. 2.8. It can be seen
that the function V' has a reverse character compared with V. VI begins at the
origin and ends at a value of unity. The maximum value of V'is the same Vmax
as appears in equation (2.34). It occurs, however, at

I 1
1/max = VI _ 2D2 (2.43)

that is for 1/ > 1, while Vmax occurs for 1/ < 1. Expressions (2.36), (2.37) and
(2.38) are also valid for D <t: 1.
Finally consider the excitation of a system by means of motion z(t) at its
support (Fig. 2.9). In this case the equation of motion is

mx + d:i: + kx = kz + di (2.44)

Putting
y=x-z (2.45)
gives
my + diJ + ky = -m.i(t) (2.46)
16 2 The one Degree of Freedom System

0= 01

015
V'
___ / / 0.2

31--- - _ - - - / - 1 0 .25
/ . 03
/
/
o
/ 05

/1-12
1
5\

OI~O~~==--~-----2~---~-

7/-
Fig. 2.8. Magnification function V'(fJ. D)

k d

Fig. 2.9. Simple system with excitation at support


2.4 Random Excitation 17

For harmonic motion z( t) = zcos wt one obtains


my + diJ + ky = mzw 2 cos wt (2.47)

The right-hand side of equation (2.47) is of the same type as equation (2.39),
so the corresponding particular solution is

y(t) = V'zcos(wt - e) (2.48)

where V' is obtained from equation (2.42).

2.4 Random Excitation

When the excitation force F(t) or the support displacement z(t) is a random
function the solution to the equation of motion is obtained by numerical inte-
gration, for which there are many established procedures (see amongst others
reference [171]). The procedure described below is employed in program MA-
DYN (see Chap. 25) and has proved very useful.
The procedure consists in representing F(t) by individual values Fi and
replacing the function between the two values Fi and Fi+1 by a straight line (Fig.
2.10). By taking relatively small time steps therefore F(t) can be represented
as a random function by the resulting series of polygons (Polygon procedure).
For a random time interval ti ~ t ~ ti+lI that is 0 ~ t* ~ L1ti the following
equation of motion is valid for the linear force characteristic

m:c.. + d:c. + k :c = F.i L1Fi *


+ ~t (2.49)
Liti
and the complete solution is

(2.50)

t
F

ti ti+1

~t--
~ t* Fig. 2.10. The polygon procedure
18 2 The one Degree of Freedom System

where
1 LlFi
v=---
k Llti
With the initial values z;,:i:i for t = ti, that is t* = 0

A = Zi - U , B = .!.. [:i:i + 5(z; - u) - v)


Wd

With known initial values Zo, Zo for t = 0 the solution can be calculated up to t l
Using equation (2.50) and the derivative z(t*) gives the end condition ofthe first
interval, that is the initial condition of the second interval. Hence calculations
can proceed over many time steps. The procedure is mathematically exact for
the series of polygons. The solution agrees with the true solution, in as much as
the series of polygons agrees with F(t). As can be seen, this is only a question
of step size, which one can correspondingly choose.
One can obtain especially simple relations with a constant step size and by
the use of non-dimensional values,

Z Z
~
z.
WnZ.
~ =
(2.51)
F
Fr
Zs = F = -
k Fr
Fr is an assumed random reference force and one obtains the equation,

{ :::~ } = [ :~~:~~ ] { :: } + 1T [~~:~~~ ]{~:+1 } (2.52)

with the elements


all = 0 + DS , a12 = S = -a2l , a22 = 0 - DS
111 2D - (2D + LlT)O + (1 - 2D2 - LlTD)S
112 = LlT + (0 - 1)2D - (1 - 2D2)S (2.53)
121 0 - 1 + (D + LlT)S
122 1- 0 - DS

and
D damping ratio, equation (2.18)
o e- D4T cos ( v'1 - D2 LlT)
S = 1 e- D4T sin (.vI
r;---n;;
- D2 LlT) (2.54)
v'1- D2
Llt
LlT = wnLlt = 211"-
Tn
It can be seen that, for a certain model with parameters m, d, k the elements of
the matrices of equation (2.52) are the same for all steps in the ca:Iculation, so
that it becomes very easy to perform.
3 Shaft with Central Disc

After the consideration of the simple system in Chap. 2 we now turn to ro-
tors. Thus in the next chapters their special phenomena will be investigated in
the simplest models possible to clarify their understanding. Later more com-
plicated models will be considered which represent adequately real rotors and
their bearings. The reader will then become conversant with more or all of the
known phenomena.

3.1 Shaft with Rigid Bearings, Jeffcott Rotor

The most simple rotor model consists of a massless shaft, at whose centre is
a fixed rigid circular disc and which is supported in rigid bearings. The model
is called a Laval shaft or leffcott rotor.l The shaft has a circular cross-section,
either with constant diameter over its whole length or having a symmetrical
distribution of stepped portions.
The disc is mounted with its plane perpendicular to the shaft axis. Its centre
of mass S can be coincident with the shaft centre W or can have a radial offset,
or eccentricity e.
The disc is assumed to move only in its own plane, or more precisely in the
plane defined by axes 1,2 in Fig. 3.1.
The movement of the shaft centre W relative to the unloaded position is
given by coordinates Yt, Y2 and the angle turned through by the disc is given
by cpo The position of the mid point S is determined from the equations

Zl Yl + e cos cp (3.1)
Z2 = Y2 + e sin cp
The system has therefore three degrees of freedom given by Yt, Y2 and cp or
by and cpo
Zt, Z2
The equations of motion can be obtained by using Newton's second law in
translation and rotation. The restoring force of the shaft and the disc weight
act on the disc. Further, one assumes a damping force from the surrounding
medium. For ease of calculation it is assumed to be viscous in character, pro-

lGustave de Laval, *1845, Orsa/Lan Kopparberg, t1913, Stockholm.


Henry H. Jefcott, *1887, Donegal, Ireland, t1937 Walton-on Thames, Surrey.
20 3 Shaft with Central Disc

I
I 5
-----::=--, X W
+----
I ~

Fig. 3.1. leffcott rotor

portional to velocity, with its resultant passing through W. These assumptions


do not correspond to reality, but are necessary for ease of solution.
Thus for the mass-centre,

-kYl - diJI
(3.2)
-kY2 - dY2 - G

where m is the mass of the disc


k the stiffness coefficient
d the damping coefficient
G mg, the disc weight.

Taking moments

where I" is the polar moment of inertia of the disc


T(t) the resultant torque.
The resultant torque is the difference between the driving torque and the sum
of the counter-moment and the resistive torque. The last arise mainly from
drag in the bearings and on the shaft which usually depend on the speed. This
dependence will, however, be neglected here.
In equations (3.2), Zl and Z2 can be substituted by the second derivatives
of equations (3.1). Using equation (3.3) also, one has the following system of
equations of motion for Yl, Y2 and cp:

mih + dYl + kYl - me (<,0 sin cp + cp2 cos cp) o


mih + dY2 + kY2 + me (<,0 cos cp - cp2 sin cp) -G (3.4)
1,,<,0 + e (dYl + kyJ) sin cp - e (dY2 + kY2) cos cp T(t) .
The equations are coupled and non-linear. One obtains essentially simpler equa-
tions if a known time variation of angle cp(t) is used instead of a known resultant
torque T(t). Thus one obtains the following equations from the first two equa-
tions of (3.4).
3.1 Shaft with Rigid Bearings, Jeffcott Rotor 21

mih + dih + kYl = me ( cp sin cp + <jJ2 cos cp)


(3.5)
mih + dil2 + kY2 me ( -cp cos cp + <jJ2 sin cp) - G .
These are linear uncoupled equations in Yl and Y2 with a known right-hand side,
which can be solved numerically (for example by using the polygon procedure
of Sect. 2.4). The resultant torque T(t) corresponding to the given cp(t) can
then be found immediately from the third equation of (3.4) by inserting the
now known values of cp(t), Yl(t) and Y2(t) and the requisite derivatives.
The case of constant angular velocity <jJ = {} has special significance. In this
case cp = {}t and from equations (3.5) the following equations are obtained

mih + dYl + kYl = me {}2 cos {}t


(3.6)
mY2 + dY2 + kY2 = me {}2 sin (}t - G .
Their form is similar to the case of unbalance excitation applied to the simple
system in Sect. 2.3. Following the solution of those equations, the particular
solutions are
Yl(t) r cos (m - e)
(3.7)
Y2(t) = rsin({}t-e)-y.
where
r = me{}2 _ V'e
V(k - m {}2)2 + (d {})2 -
V' is from equation (2.42) with", = !!..
d{} Wn

= arctan k _ m{}2

Y. ~ is the static displacement due to weight.


From equations (3.7) it is seen that the shaft centre W moves with angular
velocity {} in a circle of radius r and centre 0', the point of static equilibrium.
The angle between rand e remains constant (Fig. 3.2).
The characteristic of the relative radius r / e (= V') versus excitation fre-
quency ratio", = {} /w n is shown in Fig. 2.8. For the maximum, equations (2.34)
and (2.43) are valid or (2.36) for D ~ 1. The angular velocity at the maxi-
mum value is called critical and is usually assumed to be simplified to {}c = W n
Correspondingly Wn
nc = - , (3.8)
211"
the critical speed of the Jeffcott rotor.
The form of angle e versus", is shown in Fig. 2.5. Typical positions of points
Sand Ware shown in Fig. 3.3. Thus,

r' > r for {} < Wn


e ..l r for {} = Wn
r' < r for {} > Wn
22 3 Shaft with Central Disc

Fig. 3.2. Paths of W and S with constant angular velocity cp =n

0' 0'
s

Fig. 3.3. Relative positions of S and W for different angular velocities

For il ----t 00, S tends towards the centre of rotation a', that is it becomes
self-centring.
In order to understand further the behaviour of a rotor with constant an-
gular velocity, the forces and moments applied to the disc are considered. Sub-
stituting the solutions (3.7) into equations (3.6), and putting Q = ilt - e gives
the forces in directions 1 and 2.

mril 2 cos Q + dilr sin Q - kr cos Q +meil 2 cos ilt = 0


(3.9)
mril 2 sin Q - dilr cos Q - kr sin Q + kys +meil 2 sin ilt - G = 0 .
The corresponding resultant forces are shown in Fig. 3.4. Acting on the disc
are the centrifugal force m r il 2 and the unbalance force m e il 2 , the damping
force d il r, the spring forces kr and kys and the weight G.
It can be concluded from the figure that because of equilibrium

meil 2 sine dilr


(3.10)
meil 2 cos e + mril 2 kr
3.1 Shaft with Rigid Bearings, Jeffcott Rotor 23

Fig. 3.4. Forces and resultant torque on the JeJJcott rotor

and the resultant torque

T(t) = kresine - dnrecose + G ecosnt. (3.11)


These equations can also be confirmed by computation after transformation.
From equation (3.11) the required torque consists of a constant part and a part
oscillating with frequency n. The constant part emanates from the damping
(without damping sine is always zero). The oscillating part originates from the
weight (being zero for a vertical shaft). For well-balanced rotors with small
eccentricity T(t) is very small in comparison to the driving torque.
Finally consider the free vibration of the Jeffcott rotor. The equations in
question follow by putting G, e, T(t) = 0 in equations (3.4). Thus,

mjit + dill + kYI 0


(3.12)
mih + dil2 + kY2 = o.
These equations show that the Jeffcott rotor exhibits vibration in directions 1
and 2 as does the simple system in Fig. 2.1 in the z direction.
The corresponding model is shown in Fig. 3.5.
The solutions of equations (3.12) are, from equations (2.11),

YI(t) = e- 6t (AI coswdt + BI sinwdt)


(3.13)
Y2(t) = e- 6t (A2 cos wdt + B2 sinwdt) .
These two equations describe the path of the shaft centre W in the 1-2 plane
during free vibration, which is called natural motion. With zero damping this
natural motion consists solely of harmonic vibrations in directions 1 and 2 with
natural frequency Wn = .Jk/m. Together these give an ellipse or the special
motions shown in Fig. 3.6.
24 3 Shaft with Central Disc

- Y1

Fig. 3.5. leffcott rotor. Model for natural vibration

Fig. 3.8. leffcott rotor. Possible natural motions of the shaft centre with zero damping

The angular velocity n of the shaft can be of any value. With damping the
natural motion is similar to that without damping, except that the amplitudes
decrease with time by the factor e- 6t

3.2 Startup, Run Down

Startup or run down ofthe Jeffcott rotor is determined by the non-linear equa-
tions (3.4). For an ideally balanced disc, that is for e = 0, the problem becomes
very simple, as then the angle of rotation of the disc is independent of its dis-
placement. The angle of rotation is obtained from the equation,
Jpt(; = T(t) (3.14)
in which T(t) is the resultant torque defined in Sect. 3.1. For constant turning
moment Tc and initial angular velocity rpo, the instantaneous angular velocity
after time t is

(3.15)

In order to reach the angular velocity rp = Wn from rest, the system requires a
time,
(3.16)
3.2 Startup, Run Down 25

When an eccentricity is present, the three equations (3.4) must in general be


solved simultaneously. This can be avoided, if one assumes rp(t) as known,
whence the system (3.4) reduces to the two uncoupled linear equations (3.5).
We will allow this possibility and investigate the case of constant angular ac-
celeration with
rp=a, cjJ=at, (3.17)
From equations (3.5) it is sufficient to solve only the first equation, as the
second equation gives the same solution for Y2 as for Yl but with a constant part
Y. = -G/k together with a phase shifting. We thus investigate the solution of

mjiI+dilI+ k Yl =me[asin(~e) +(at)2cos(~t2)] . (3.18)

In order to consider the minimum number of parameters, the following variables


will be used, Yl
x

r (3.19)

a =
Putting D as the damping ratio and ( )' as d( )/dr, equation (3.18) gives

x" + 2Dx' + x = a sin rp + (ar)2 cos rp


a- r2 (3.20)
rp =
2
The essential parameters are thus only the non-dimensional acceleration a and
the damping ratio. Further, it can be observed that, in equations (3.20), the
first term on the right-hand side is the only one of significance. The ratio of
the factors for sin rp and cos rp is ar2 = 2rp and so the second factor (ar y very
soon surpasses the first, a.
Solutions to the above problems are discussed in several publications, for
example [17J, [18J, and [19J, from which further literature can be extracted. The
essential results are shown in the following figures, which are produced with the
help of the polygon procedure described in Sect. 2.4.
Figs. 3.7 and 3.8 use as an example a Jeffcott rotor with damping ratio
D = 0.01. They show the character of the exciting force as a function of time
and of the displacement in direction 1 for an angular acceleration a = O.Olw~,
that is a = 0.01. From equations (3.19) and Tn = 27r/wn , the non-dimensional
time t /Tn = r /27r. Also the non-dimensional angular velocity cjJ / Wn = a rand
the number of revolutions z = rp/27r = a r 2 /47r. Thus these figures can be con-
sidered as showing time-dependent angular velocity and number of revolutions.
In the example thirty revolutions are covered, very high starting acceleration
being chosen.
26 3 Shaft with Central Disc

, , , ,
10 20 30 .l.._
Tn
, , ,
0.5 1.5 2 .i_
!.Un
, , ,
5 10 20 30 z -
Fig. 3.7. Excitation force for startup with a = 0.01 w~. leJJcott rotor with D = 0.01
The exciting force oscillates with circular frequency cj; and the maxima fol-
low an approximately parabolic envelope. The displacement becomes distinctive
after a while, its maxima growing in a roughly parabolic fashion decreasing af-
ter an absolute maximum in an oscillatory way with time. In this example, the
absolute maximum value occurs at cj; = 1.2wn and has the value 14.4e. For a
non-accelerating, quasi-stationary drive (dashed curve), the maximum occurs
at cj; = Wn with a value Ylmax = e/2D = 50e. The behaviour in this example is
typical and is also generally valid:

For accelerating startup the envelope of displacement amplitudes has an


absolute maximum and is followed by lower maxima.

The absolute maximum occurs at an angular velocity cj; > Wn and is smaller
than for a non-accelerating drive.

For run down, that is for slowing down of angular velocity, the envelope
curve is approximately symmetric to cj; = wn

The absolute maximum amplitude occurs for cj; < wn It is likewise smaller
than for a non-decelerating drive and it displays smaller maxima with
reduction in angular velocity.
3.2 Startup, Run Down 27

rI
SO

-10

1j1= wn --1 LIj1=1.2w n

$ i
10
i
20
i
30 ..1.. -
Tn
$ O.S 1.S 2 .t_
Wn

$ S 10
i
20 30
i

z -
Fig. 3.8. Amplitude at startup with a = O.Olw~, JefJcott rotor with D = 0.01

The position and value of the absolute maximum displacement depend upon
the non-dimensional acceleration 0 = a/w~ and upon the damping ratio. The
upper graph of Fig. 3.9 shows the position of the maximum, characterised by
CPmax/W n The lower graph shows the value of the maximum as a ratio of the
maximum for non-accelerating drive (a = 0).
The number of revolutions Zn turned by the rotor up to an angular velocity
CP = Wn is a better measure than is the non-dimensional angular acceleration
o = a/w~. From the above equations, Zn = (1/471")0. The scale assigned to Zn in
Fig. 3.9 shows that the effect displayed is stronger where fewer rotor revolutions
are necessary up to cp = w n Turbomachines are as a rule started up or shut
down so slowly that they display resonance as if with a non-accelerating drive.
The required resultant torque T{t) for constant angular acceleration follows
immediately by inserting the results into the left-hand side of the third equation
of (3.4). The more difficult case of a given torque has been investigated by
Gasch, Markert and Pjiitzner in references [19] and [20]. These authors assumed
constant torque and found the following essential results:
28 3 Shaft with Central Disc

0= 0.01
0.02
0.05

1.1

I f I I Iii I I f I
0.001 0.002 0.005 0.010

----
...!!.-
Wn1

0= 0.05
1
Ylmax hx >O)
Y1max (a=O)

0.5

0 i i I I
, I I
,,, I
0.001 0.002 0.005 0.010
a
w~-
I i i i
I I I I
100 50 20 10 5
-zn
Fig. 3.9. Position and value of the maximum displacement of a Jeffcott rotor as a
function of starting acceleration. Zn = number of revolutions up to ~ = Wn

For a sufficiently high torque the rotor behaves in a similar way to one
having constant acceleration. This is shown in the character of amplitude
and of angular velocity with time, in the example of Fig. 3.10. The latter
transpires to be roughly linear, corresponding to roughly constant angular
acceleration.

For an insufficient torque the rotor does not emerge from the critical region,
it "stays suspended" (Fig. 3.11).

Reference [19] contains charts, from which the behaviour can be numerically
portrayed. In [20], it is shown that one can calculate the behaviour of a realistic
rotor, using the results for a leffcott rotor found in reference [19].
3.3 Jeffcott Rotor with Flexible Bearings 29

Angular velocity

Fig. 3.10. leffcott rotor. Amplitudes and angular velocity for startup with constant
torque [19]

Angu lar ve locity

/ Displacement

Time -

Fig. 3.11. As in Fig. 3.10, but with insufficient torque [20]

3.3 J effcott Rotor with Flexible Bearings

Up to now in this chapter the bearings supporting the rotor have been assumed
to be rigid . The bearings of real shafts are more or less flexible and have their
special dynamic characteristics. These will be investigated in more detail in
Chapters 6 and 7.
30 3 Shaft with Central Disc

In this section it will simply be assumed that a bearing can be replaced


by massless springs in two mutually perpendicular radial directions (preferably
horizontal and vertical). Thus, for the Jeffcott rotor, one obtains the model
shown in Fig. 3.12. The two bearings have equal pairs of stiffnesses kll k2 in
directions 1 and 2 respectively. When kl i- k2 the bearing is referred to as
anisotropic. Using k as the stiffness of the shaft with stiff bearings, the total
stiffness of the system of shaft and bearings is

in direction 1 k' k
I - I + k/2k l
(3.21)
and in direction 2 k' k
2 - 1 + k/2 k2

A drive is assumed giving constant angular velocity cp = [J for which equations


of motion (3.6) hold for stiff bearings. The first of these equations relates to
direction 1, the second to direction 2. From these, the equations of motion
follow for the Jeffcott rotor with flexible bearings in which k is replaced by k~
and k~, respectively:

mjit+ dYI + k~YI m e [J2 cos [Jt


(3.22)
mih + dY2 + k~Y2 m e [J2 sin [Jt - G .

By analogy with equations (3.7), the particular solutions of equations (3.22)


are
fit cos (at - el)
(3.23)
fh sin ([Jt - e2) - ~
2

2 2

5
D,W
.... o~==~~ ....

i i
~===l0t-

Fig. 3.12. leffcott rotor with flexible bearings


3.3 Jeffcott Rotor with Flexible Bearings 31

with
iii V;'e

"I;
V'
)(1- "In + (2Di "1;)2
2
{}
"Ii
Wi (3.24)

Wi
~R
+ m = k 2ki Wn wn=~
d
Di
2m Wi
2Di"li
ei arctan---2 , i = 1,2.
1- "Ii
Accordingly the disc centre undergoes harmonic motion in directions 1 and 2,
under the action of unbalance excitation, with a frequency equal to the shaft
frequency and amplitudes iit and ih, whose character corresponds to V' in Fig.
2.8. Because of the different stiffnesses the shaft has two natural frequencies
WI, W2 that is two critical speeds nIl n2. Fig. 3.13 shows as an example the
character of the amplitudes versus n = {}/27r. The maximum amplitudes are,
from Sect. 2.3, with i = 1,2:
~
Yi,max =
2D
yh-Di
e
1- D2
: : : ; -D
e
2;
.hD
WIt i ~ 1 (3.25)

9,
y,

n-

Fig. 3.13. Amplitudes of leffcott rotor with anisotropic flexible bearings.


nI ~ wI/27r, n2 ~ w2/27r, nn = wn /27r.
32 3 Shaft with Central Disc

With rigid bearings the disc centre describes a circle in plane 1,2, as dis-
cussed in Sect. 3.1. This is also the case for isotropic flexible bearings with
k2 = k1' where f12 = fit and e2 = e1' With anisotropic flexible bearings, that is
k2 i- k1 the path of the disc centre is, in general, an ellipse, whose shape and
major axis direction depend on the shaft speed. Such ellipses are shown in Fig.
3.14. In this example, the elliptical orbit is roughly horizontal in the region of
nl! and roughly vertical in the region of n2. For 1.04 nI, and 0.96 n2 the orbit
degenerates to a straight line. Below and above these speeds the direction of
rotation of the orbit is the same as for the rotational speed of the shaft, that is
forward whirl. Between these values, it is in backward whirl.
The expressions for the natural vibrations of the Jelfcott rotor with flexi-
ble bearings are obtained from equations (3.13) if one replaces Wd in the first
equation by

and in the second equation by (3.26)

where 5 = d/2m.

+=F+~efl-&-
0.9n, n, 1.04n, 1.13n, 0.96n, n, 1.1n,
Forward whirl Backward whirl Forward whirl

Fig. 3.14. Jeffcott rotor with 1:1 = 0.51:, 1:2 = 21:, D = 0.1. Orbits of disc centre W

The system thus vibrates in directions 1 and 2 with a damped natural


frequency of equal decay constant and in general/different natural frequency.
The natural motion in plane 1,2 consists of the two natural vibrations together
and depends on the initial condition. An example is given in Fig. 3.15. It shows
the orbit of the disc centre W, if initially the radius r and the velocity rW1
lie perpendicular to each other. The orbit path is somewhat confused. This is
clarified for the corresponding natural frequencies in directions 1 and 2 (Fig.
3.16), which have the well-known characteristics of damped vibration of a simple
system.
3.4 JefFcott Rotor with Bearing Damping 33

Fig. 3.15. JeJJcott rotor with k1 = 0.5k,


k2 = 2k, D = 0.1. Natural motion of disc
centre W

Yl
O.Sr

-O.Sr

Y2
O.Sr

0
3T1

-O.Sr

Fig. 3.18. Natural vibrations in directions 1 and 2 for the example of Fig. 3.15

3.4 Jeffcott Rotor with Bearing Damping

In the sections of Chap. 3 up to now it has been assumed that the surrounding
medium acts as a damper on the disc of the shaft. Hence one arrives at simple
equations and a fundamental understanding. Mostly, however, the damping at
the bearings is much greater than that at the disc. This is especially true of
a shaft which sleeve bearings, but also the bearing housings and the rest of
34 3 Shaft with Central Disc

the bearing structure give a large contribution to the damping of the system.
With the objective of investigating relevant phenomena for the simplest possible
model, the model shown in Fig. 3.17 will now be considered. This and similar
models were investigated in references [21) to [25). The following procedure is
extracted from ref. [26).
Isotropic bearings are assumed and constant angular velocity of the shaft.
Using the procedure of Sect. 3.1, it is sufficient to investigate only one direction
of motion. The displacements ZI and Z2 given in Fig. 3.17, which relate to the
disc and the bearings, respectively are valid for an arbitrary direction in the
plane 1,2 of Fig. 3.1. The equations of motion of the model are

mZI + kSZI - kSZ2 = F(t)


(3.27)
2dB z2 - kSZI + (ks + 2kB ) Z2 = 0

where m is the disc mass


ks the shaft stiffness
kB the bearing stiffness
dB the bearing damping.

Natural vibrations are first considered, that is the solutions of equations (3.27)
for F(t) = O. The solutions

ZI(t) = 'PI eAt , Z2(t) = 'P2 eAt

yield, as usual, the eigenvalues and the general solution. The model has one
degree of freedom and a half and not two degrees of freedom, because Z2 only
occurs up to the first derivative. Hence there are three eigenvalues. In fact, as
a rule, a complex conjugate pair and a real eigenvalue exist:

(3.28)

For certain parameters three real eigenvalues can also be obtained,

(3.29)

!x,.F(t)
m

Fig. 3.17. leJJcott rotor with bearing


/ damping
3.4 Jeffcott Rotor with Bearing Damping 35

Accordingly, the model usually has one natural frequency w, but not with all
parameters. The real part of the eigenvalue is always negative, so that the
natural motion usually consists of one damped natural vibration and one slowly
decaying displacement. In special cases there are three such displacements with
different decay constants.
Fig. 3.18 shows the non-dimensional natural frequencies as functions of
non-dimensional bearing stiffness, for different bearing damping. The reference
value for natural frequency 1 is the natural frequency for stiff bearings

lo=~fi
2'/1' m
(3.30)

where
ks = 48EI (3.31)
P
For a shaft with constant cross-section, one would expect 1 to be always smaller
than 10' There is no natural frequency in a certain region where kB < 0.0625 ks
and dB < 0.3248 Jks m (see [26]).
For unbalance excitation with F( t) = men 2 cos nt, the forced vibration
:1:1 (t) = Z cos( nt - e) results in known resonance behaviour, if the excitation
frequency is roughly equal to the natural frequency w. The maximum ampli-
tude zmax of the disc is shown in Fig. 3.19 as a function of bearing damping
for different bearing stiffnesses. For increasing bearing damping the amplitude
decreases as expected. After a minimum however, it increases again, which is
obvious, since for infinite damping, the bearing journal can no longer move and
the damping becomes inoperative. It can be seen that, in general, for efficacy

0.5

O+-------.--------r-------r----~~
a 0.5 2 5 00
ks/ks -
Fig. 3.18. Natural frequency of Jeffcott rotor with bearing damping
36 3 Shaft with Central Disc

t
e

10-1 -I---r-rTTITrrr-"'rTTT1lrr-lrTTTTITn---Y--rTlTTTTr---'-TTTl~--'--'rTTTrnr----

---
10- 3 10 2 10 3
dB
---
Vksm'
Fig. 3.19. Maximum amplitude of the disc of a Jeffcott rotor with bearing damping

of damping, a certain amount of movement is required. The optimal bearing


damping to ensure minimum vibration amplitude is shown in Fig. 3.20.
In ref. [26J an example is given where the results of this simple model are
well suited to the determination of the severity of resonance of a real rotor.

t 6

({k:~' 4 L
2

2 3 4 5
~---
ks
Fig. 3.20. Optimum bearing damping
4 Shaft with Non Central Disc

In the model of Chap. 3 the disc was assumed for simplicity to be symmetrically
placed between the bearings in the centre of the shaft. Hence consideration could
be confined to plane motion ofthe disc. Now consider the three models shown in
Fig. 4.1, with either a noncentral or an outboard disc. Besides the displacement
of the disc, its rotation around a radial axis must also be considered. As a result
its moment of inertia and gyroscopic effect are called into play. The theory of
such a model has been given in many publications, for example [27], ... , [31].
Consideration will first be given to the equations of motion, then the natural
frequencies and later forced vibrations as a result of unbalance and of a har-
monic force with fixed radial direction. Finally, the special features of anisotropic
flexible bearings will be investigated.

4.1 Equations of Motion

The models of Fig. 4.1 should have the following properties:

- Massless, flexible shaft with circular cross-section and constant or varying


diameter along its axis.

Stiff circular disc with mass m and polar and diametral moments of inertia
i p , i d , respectively.
- Rigid bearings

- Drive, giving constant angular velocity {}.

Model I Model IT Model m

Fig. 4.1. Models of this chapter


38 4 Shaft with Non Central Disc

It will be first assumed that the centre of mass of the disc S is coincident with
the centre of rotation, the point W on the shaft. It will also be assumed that
the eccentricity e = 0, and so no unbalance force will exist. Further, the plane of
the disc is assumed to be perpendicular to the shaft axis, so that no unbalance
moment exists. No other excitation forces or moments exist, in particular the
gravity force is zero.
The equations of motion for free vibration will first be considered. The disc
can move in the radial direction and, as well as rotating around its axis can
turn about a random radial axis. The displacement and rotation should take
place in the space determined by axes 1 and 2 in Fig. 4.2 and are assumed
S in the directions of the respective axes. CP3 and CP4 are rotations of the disc
around axes 1 and 2, respectively, or around a parallel axis. Fh F2 and M 3 , M4
are corresponding forces and moments.
The equations of motion are

mZ1 EF1
mZ2 = EF2
(4.1)
D3 = E M3
D4 EM4 ,

in which D3 and D4 are the time derivatives of the angular momenta. The terms
on the right-hand side of equations (4.1) are the forces and moments applied
to the disc from the shaft. In order to determine them, we consider now the
displacements and rotations of the shaft at S as a result of forces F1, F2 and of
moments M 3 , M 4 Using displacement influence coefficients hik'

:1:1 =hllF1 + h14M4


:1:2 h22F2 + h23 M3
(4.2)
CP3 = h32F2 + h33M3
CP4 = . h41F1 + h44 M4

Fig. 4.2. Displacements, rotations,


forces and moments
4.1 Equations of Motion 39

By inversion one obtains the forces and moments required for the displacements
and rotations
F1 k l1 :i:1 + k 14cp4
F2 k 22 :i:2 + k 23cp3
(4.3)
M3 = k 32 :i:2 + k 33cp3
M4 k 41 :i:1 + k44cp4 ,

where kik are force influence coefficients or stiffnesses. Corresponding reactions


act on the disc as a result of displacements :i:1, :i:2 and rotations cp3, cp4. Thus,
from equations (4.1) and (4.3),

mZ1 -kn:i:1 - k14 cp4


mZ2 -k22:i:2 - k23 cp3
(4.4)
D3 -k32:i:2 - k33cp3
D4 -k41:i:1 - k44 cp4

In order to obtain D3 and D4 consider Fig. 4.3. This shows on the left the torque
components and on the right their time derivatives.
Accordingly,
D3 = Idr{;3 + I pncjJ4
(4.5)
D4 = I dr{;4 - IpncjJ3
Using the displacement influence coefficients the following simplifications hold.
From symmetry,
h n = h22 , h44 = h33 , h14 = -h23 (4.6a)
and from Maxwell's reciprocal theorem,

(4.6b)

t IpQIP3

J I dlji4

:$:;, I p QIP4
Fig. 4.3. Torque components and their time derivatives
40 4 Shaft with Non Central Disc

Thus the force influence coefficients are,

k22 = h
kll =~
-h23
k23 k32 = -k14 = -k4l = - -
Ll (4.7)
h
k33 = k44=~
Ll = h22h33 - h~3 .

Only three force influence coefficients are thus of significance. Choosing k22' k23
and k33 gives from equation (4.4) the ultimate system of equations,

mZl + k22 :1)l - k23I(J4 = 0


mZ2 + k22 :1)2 + k23I(J3 = 0
(4.8)
ldCP3 + lpOl{;4 + k23:1)2 + k33I(J3 = 0
ldCP4 - lpOl{;3 k23 :1)l + k33I(J4 = 0.

These are the equations of motion for free vibration of the models of Fig. 4.1.
The required displacement influence coefficients for a shaft with constant cross-
section (El constant) are given in tables 4.1 and 4.2. For a stepped shaft, one
would first obtain the appropriate displacement influence coefficients h22' h23
and h33 by static considerations and hence, from equations (4.7), the force
influence coefficients.

Table 4.1. Displacement influence coefficients for the models of Fig. 4.1

Model I Model II Model III

a=Jb a = b = 1/2
h22 a2b2 [3
c2~'iI c) c3
3EY1 48EJ 3EJ
ab(b - a) c(21 + 3c) c2
h23 0 6El 2EI
3Ell
a3 + ~3 1 c
h33
3Ell I2E1 '3"l;lc E1
Ll a3b3 c3( 41 + 3c) c4
9(El)212 36(El)2 12(El)2
4.2 Natural Vibrations 41

Table 4.2. Force influence coefficients for the models of Fig. 4.1

Model I Model II Model III

a=/:b a = b = 1/2
b3
1 ac
3 48El 12El
k22 3El!LJt.. 12El
a3b3 13 c3(41 + 3c) ~
3El'(a - b) El 21 3c 6El
k23
a 2b2
0
-6 c2(41 + 3c) -7
12El 4El
k33 3El (1)
I
-r 12El ccl,~ c3c ) c

In matrix notation, one uses the displacement vector

the mass matrix

the gyroscopic matrix

where 934 = lpn, 943 = -934, other elements = 0


and the stiffness matrix.

K = (kik) with kik from equations (4.7).

One thus obtains the equation corresponding to equations (4.8),

M:i:+ Gx+Kx = o. (4.9)

4.2 Natural Vibrations

The free or natural vibrations of the rotating shaft are obtained from the solu-
tions of equations (4.8) or (4.9), using appropriate initial conditions. With the
substitution

one obtains
(4.10)
For cp =/: 0, the determinant of ( ) must be zero. It follows that the characteristic
equation is
(4.11)
42 4 Shaft with Non Central Disc

In the coefficients of equation ( 4.11) there occurs the angular velocity of the shaft
to the power two, and so the eigenvalues AI: are independent of the direction of
shaft rotation. It is thus sufficient to calculate eigenvalues for positive n values
only.
It transpires that four pairs of complex conjugate eigenvalues exist, of the
form AI: = j WI:, Ak = - j WI: (k = 1, ... 4). The typical behaviour of these
eigenvalues versus n is given for a sample case in Fig. 4.4.
The kinematic interpretation of the natural vibrations is not simple with
this type of solution. On the other hand there is a very evident interpretation
obtained by the introduction of complex variables for the displacements and
rotations. Hence, we find that the essential natural frequencies are the positive
values of W2 and W4 and the negative values of W1 and W3. Although in general,
one does not speak of negative natural frequencies, here this terminology is
suitable.
Thus introduce the complex variables
z = Z1 + jZ2
(4.12)
1j; = CP4 - jCP3

whose geometric meaning is shown in Fig. 4.5. z and 1j; lie in different mutually
perpendicular planes. Using the equations (4.12), equations (4.8) reduce to

mz + k 22 Z - k23 -rP = 0
(4.13)
h; - j Ip n~ - k23Z + k331j; 0
whence only four eigenvalues are expected. Substitutions

z(t) = r eiwt 1j;(t) = a eiwt (4.14)

Wo -----2wo W,
W2 Q-
W3
-2wo
t -3wo
Im'A."k -4wo

Fig. 4.4. Eigenvalues of a shaft as in model I of Fig. 4.1, with a = 0.351, ip = 0.41,
Id= 0.5Ip - - - Asymptote
4.2 Natural Vibrations 43

Im z

x, Rez
Fig. 4.5. Geometric dispositions of z and ""

with real parts r and Cl lead to equations,

(_mw 2 + k22 ) r k23 Cl 0


(4.15)
- k23 r + (-IrJW2 + Ipnw + k33) Cl = 0

and to the characteristic equation

As usual, we reduce the number of parameters to an absolute minimum by the


introduction of non-dimensional quantities. These are

II =!:!....
Wo
, ", = ~
Wo
, wo = Vh 1 22m
, (4.17)

where Wo is the natural circular frequency for the models of Fig. 4.1 when the
disc is replaced by a point mass. Substitution into equation (4.16) gives, after
some manipulation, the dimensionless characteristic equation,

(4.18)

where

a4 = (A - B2)p~ a3 = (A - B2)2 p;
a2 =
1 +Ap~ al Ap; ( 4.19)
h33 12 h231
A B
h22 h22
and the non-dimensional radii of gyration
id
Pd = ,pp

where (4.20)

f.
44 4 Shaft with Non Central Disc

The non-dimensional eigenvalues Vk = Wk/WO thus depend only on the ratios


id/l and ip/l, the radii of gyration and the influence numbers A and B. For a
shaft with constant cross-section, A and B depend only on the position of the
disc, defined by a/lor c/l (see table 4.3). For a thin disc, only one radius of
gyration is needed as ip = V2i d For a constant cross-section and thin disc the
non-dimensional eigenvalues thus depend only on two parameters.
For a thin circular disc of radius R
R2 R
1d = m-
4
id = 2
R2 (4.21)
R
1p m- ip =
2 v'2'
and thus 1d = O.51p
Table 4.3 summarizes the expressions for reference frequency Wo and for the
values of A and B for our models.

Table 4.3. Reference frequency Wo and A, B from equations (4.19) for the models of
Fig. 4.1.

Ii = a/I, b = b/l = 1 - Ii, e = c/l .


Model I Model II Model III

Wo
l3 E1 I
m a 2 b2
J: 3 E1
mel + c)c2
/3 E1
mc3
li3 + -3
b 1 +3 e
A 3
7i2T/ (1 + e) e2
b-Ii 1 + 1.5 e
B 1.5
bli (1 + e)e

The characteristic equation (4.18) has four real roots, two negative and two
positive. Their numerical evaluation reveals that Vi and V3 are negative and V2
and V4 positive. Fig. 4.6 shows the typical character of the natural frequency
Wk = VkWO versus angular velocity n of the shaft for the example of Fig. 4.4.
The complete solution using all four natural frequencies is

k = 1, .. .4. (4.22)

For a certain value of k the solution is


jWkt jWkt
Zk(t) = Tke Wk(t) = ak e , (4.23)
4.2 Natural Vibrations 45

t 4wo
Wk

2wo
:::::::----
0
0
Q-

Fig. 4.8. Natural frequencies from equation (4.18) for the example of Fig. 4.4.

which fr:om Figs. 4.5 and 4.2 describes the following motion, the kth natural
motion:
The point S moves in a circle of radius Tk with constant tangential
velocity TkWk. The plane of the disc makes a constant angle O:k rela-
tive to the line 0 S (Fig. 4.7). For positive Wk there is forward whirl,
for negative Wk, backward whirl.
The natural motion of the model considered here differs fundamentally from
the natural motion of the Jeffcott rotor considered in Chap. 3. In the latter
the natural motion takes the form of an ellipse or in special cases a circle or a
straight line. In the present case the gyroscopic effect produces a circular natural
motion. In a random direction through 0 in the plane 1,2 the point S describes
a harmonic natural motion with frequency Wk. It should be emphasised that
the Jeffcott rotor with rigid or flexible bearings has only one natural frequency,
which is independent of the shaft speed, while the models of Fig. 4.1 have four
natural frequencies, which depend on the shaft speed.
The deflection form of the shaft at a natural frequency is knQwn as the mode
shape. It is determined by Tk and O:k, for which the first of equations (4.15) is
valid
(4.24)

View in direction A Fig. 4.7. kth natural motion.


46 4 Shaft with Non Central Disc

Four mode shapes exist, which depend on the angular velocity n, that is on
shaft speed, as do the natural frequencies.
As an example consider the mode shapes of model I with a = 0.35 I, ip =
0.4 I and Id = 0.5 Ip (as in Figs. 4.4 and 4.6) first for a stationary shaft, then
for n = 0.8 woo Using the given formulae one obtains
for n = 0:
- Wl = W2 = 0.914 Wo , Cll = Cl2 = -1.868 rd1
- W3 = W4 = 1.845 Wo , Cl3 = Cl4 = 6.694 r3/1

and for n= 0.8 Wo:

Wl -0.698 Wo Cll= -3.028 rdl


W2 1.033 Wo Cl2 = -1.095 r2/1
W3 -1.451 Wo Cl3 = 2.365 r3/1
W4 2.716 Wo Cl4 = 19.930 r4/1

The corresponding mode shapes are shown in Fig. 4.8. With one exception, it
was assumed that rk = 0.1 1. For the fourth mode shape at n = 0.8 Wo it was
assumed that r4 = 0.05 I and so,

Cl4 = 19.93 x 0.05 = 0.997 == 57.1 .

In fact the displacements are naturally much smaller but the mode shapes have
the character shown.
To achieve a better understanding, we consider the deflection of the shaft
due to bending under static load F 2 Here

Forward whirl

Backward whirl

Forward whirl

Backward whirl

Statics Q: 0 Q:O. 8w o

Fig. 4.8. Natural mode shapes of model I with a = 0.35 I, ip = 0.4 1 and Id = 0.5 Ip.
4.2 Natural Vibrations 47

l'

h23 1 l'
a ---.-
h22 1
and with the data of the example,

a = -1.3191'/1 , that is a = -7.55 for l' = 0.11 .

Thus one can ascertain the following. For the first two mode shapes the disc is
inclined as for a static load. For the third and fourth mode shapes the inclination
is in the opposite direction.
The complete solution (4.22) depends on the initial conditions. They give
a system of linear equations for 1'1, 1'4. The associated angles a1, ... a4 are
found using equation (4.24). Using the model considered here as an example,
with initial conditions,

z(O) l' z(O) j l' a


'Ij!(0) -1.0764 1'/1 -(O) j 'Ij!(o)a ,

Fig. 4.9 shows the motion of point S. It consists of all four natural motions and
has the equation,

z( t) = 0.0802 l' e
jW1t
+ 0.8789 l' ejW2t + 0.0392 l' ejW3t + 0.0018 l' ej W 4 t
Having clarified the natural motions, turn again to the natural frequencies and
consider the special case of a central disc, that is model I with a = 1/2. Here
k23 = 0, so that equation (4.16) has the following solutions.

W1, W2
~
=f-

Ipa
m
( 4.25)
Cpa) k33 2
W3, W4
2Id =f 2Id + Id .

X2

Xl

Fig. 4.9. Natural motion of the disc centre S from t = 0 to t = 3 To (To = 211" /wo).
Data as in Fig. 4.8, n = 0.8 Woi see text for starting conditions.
48 4 Shaft with Non Central Disc

WI and W2 represent the natural frequency Wn of the Jeffcott rotor, for displace-
ment only and no inclination of the disc. With inclination of the disc, natural
frequencies W3 and W4 are added, in which W3 represents backward and W4 for-
ward whirling. For [J = 0

W30 , W40 = ~~ . (4.26)

These natural frequencies are equal to WI, W2 if k33 = k22 i~ or when a shaft of
constant section has a thin circular disc, whose radius of gyration is id = 1/2.
Fig. 4.10 shows the natural frequencies of this model with different radii of
gyration.

-==========} -
O~--~.----'-----r-----.----
W3

Fig. 4.10. Natural frequencies of model I for a = 1/2 and Id = 0.5 Ip.

Figs. 4.11 and 4.12 show the natural frequencies of models II and III for
ip = 0.4 I and 0.4 c respectively, for which the shafts are always assumed to be
of constant cross-section.
The natural frequencies of models I, II and III are seen to have similar
trends when plotted against [J. For [J = 0, each pair of natural frequencies are
equal in value and using equation (4.18) with", = 0,

WlO, W20 = ~WoJ~ (a2 - Ja~ - 4a4)


(4.27)
W30, W40 = ~WoJ 2~4 (a2 + Ja~ - 4a4) ,

with a2 and a4 obtained from equations (4.19). These natural frequencies are
shown in Figs. 4.13, 4.14 and 4.15 for models I, II and III, respectively. The
4.2 Natural Vibrations 49

/l zs ~
I Ic I __ /
J-======----___
4~ I

-=-"'::..:::::::~-= - W3
__ - W2
.....-::--

Q-
Fig. 4.11. Natural frequencies of model II for c = 0.2 I, ip = 0.4 I, Id = 0.5 Ip

.,-/-/
__ - -

------
_ _ -;:::;-..L _ _ _ _ :::.:::;:::..__

.// W2
-w,
O~---.----~-----.----~----~

Q-
Fig. 4.12. Natural frequencies of model ill for ip = 0.4 c, Id = 0.5 Ip.

f
W40
Wo
w20
3

2
1t1 all
0.2 } W40
/0.35 Wo
Wo 1 /0.5
/0.5 } w20
0.35 Wo
0.2
2

Fig. 4.13. Natural frequencies of model I, with Id


it
I --= 0.5 Ip, for n = o.
50 4 Shaft with Non Central Disc

5
ell
W40 4
Wo
01 } W40
w20 3
0.2 Wo
Wo
0.3
2
lS. ZS ~
I I c.1
t
I

3 } w2
zD-2
yO.1 w:
0

--
0 2
ip
I
Fig. 4.14. Natural frequencies of model II, with Id = 0.5 Ip, for n = o.

0.5 1.5
.!.e._
e
Fig. 4.15. Natural frequencies of model III, with Id = 0.5 Ip, for n = o.

asymptotes of these curves are of interest. They are obtained from equation
(4.16) with [} ~ 00 and w ~ 00. In the first case the following equation
applies,

with the solutions,

w = 0 and ( 4.28)
4.2 Natural Vibrations 51

In the second case, it follows from

that
( 4.29)

Equations (4.28) relate to the asymptotes for -WI, W2 and -W3 and equation
(4.29) describes the asymptote for W4. This is a straight line through the origin
with slope lp/ la, that is a slope of two for a thin circular disc.
The frequency Wi. from equation (4.28) is the natural frequency of the model
when inclination of the disc is prevented. This is shown in Figs. 4.16 and 4.17
for models I and II. For model III

_ J12 EI ( 4.30)
Wi. - 3
me
The behaviour of the natural frequencies versus {} for models with other pa-
rameters can thus be easily estimated.

O~--~--,---.---~---.-
o 0.1 0.2 0.3 0.4 0.5 Fig. 4.16. Natural frequency
..9.. of model I when inclination of the disc
I is prevented

W.I. 3
Wo

O~--,---.---~---.--~---
o 0.1 0.2 0.3 0.4 0.5 Fig. 4.17. Natural frequency of
c model II when inclination of the disc
I is prevented
52 4 Shaft with Non Central Disc

4.3 Unbalance Vibrations

Unbalance vibrations result from an unbalance force or unbalance moment.


An unbalance force arises from eccentricity of the mass centre. An unbalance
moment arises when the disc is not mounted perpendicular to the axis of rotation
of the shaft, that is, it is mounted obliquely. We assume a small eccentricity
and small inclination of the disc. The equations of motion thus result to a
good approximation by adding corresponding terms on the right-hand sides of
equations (4.8). A mass-eccentricity e generates the unbalance force

( 4.31)

where unbalance
U=me. (4.32)
A moment results from an inclination, which is derived in Sect. 5.1, and is
described as a centrifugal moment. From equation (5.2),

(4.33)

where
(4.34)
which is the unbalance moment.
Fig. 4.18 shows the unbalance force vector and the centrifugal moment vec-
tor, when the eccentricity and the axis of inclination make an angle fJ. The shaft
rotates which constant angular velocity {}, whence the vectors also rotate with
this angular velocity. Finally, one again introduces viscous damping on the disc,

";)

Fig. 4.18. Unbalance force and centrifugal moment


4.3 Unbalance Vibrations 53

with coefficients d for displacement and d for rotation. Hence the equations of
motion for unbalance vibration are

mil + dZ l + k22:1:l - k23CP4 Fucos {It


mi2 + dZ 2 + k22:1:2 + k23CP3 Fusin{lt - G
(4.35)
IdtP3+ d rp3 + I p{lrp4 + k23:1:2 + k33CP3 Mu cos( {It + {3)
IdtP4 + d rp4 - I p{lrp3 k23:1:l + k33CP4 Mu sin( {It + {3) .

As in the section on natural frequencies, a complex displacement z = :1:1 + j:l:2


and a complex rotation "p = CP4 - j CP3 are introduced, whence from equations
(4.35), the following system of equations results.

mz +dz +k22Z - k 23"p = Fu ejOt - j G


(4.36)
h~ + (d -j Ip {l).,j; -k23Z + k33"p = -j Mu ej{ot+~)

Particular solutions of equations (4.36) are

z(t) Zo + Z e jOt (4.37)


"p( t) "po + !Ii e jOt
The terms Zo and "po arise from the gravity force. The unbalance force and the
centrifugal moment produce a circular motion of the disc-centre W with radius
r = IZI and slope 1!Ii1 of the disc. The circular motion has the angular velocity
{l of the shaft and occurs in the same direction. Hence it is to be expected
that resonance occurs when {l is equal to a natural frequency and is forward
in direction. We obtain these frequencies to a good approximation from the
characteristic equation (4.16) of the model without damping.
For w = {l, we have
m (Ip - I d) {l4 - [(Ip - I d) k22 - m k 33 ] {l2 - k22k33 + k~3 = 0
with solutions

the critical angular velocities.


For {l = {lc equality results with the first natural frequency of synchronous
vibration W2 and for {l = {l~, with W4. The last case is only possible for Id > Ip
as the asymptote to W4 has a slope Ip/ld and hence for Id < Ip there is no
intersection with the straight line of slope w = fl.
We confine ourselves, in this chapter, to models with thin discs, for which
only the critical angular velocity {lc exists. The effect of gyroscopic action in
such models is shown in Fig. 4.19. Without gyroscopic action two critical angular
velocities exist, namely {l = W20 and {l = W40, while with gyroscopic action,
54 4 Shaft with Non Central Disc

u}~1..51
?
/'
/'
W /' W,,_Q
40 /' _ -W3
W20 - - - -----=--- W2
-:;::::.._- -W,
Q-
Fig. 4.19. Critical angular velocity for Id < Ip

only for Dc > W20 is there a critical angular velocity. The gyroscopic effect thus
raises the lower critical angular velocity and prevents the higher. This is valid,
as mentioned, for Id < Ip. The behaviour for Id > Ip will be investigated in
Chap.5.
In Figs. 4.20, 4.21 and 4.22 are shown the critical angular velocities of
the three models, each with a thin circular disc. It is found that due to the
aforementioned influence of gyroscopic action, the value of the critical angular
velocity increases strongly with ip/l or ip/c.
To investigate the behaviour at resonance consider an example, namely
the model I already investigated in which a = 0.35 l, ip = 0.4 land Id =
0.5 Ip. Assume an unbalance force with U = m e and introduce damping d to
limit the resonance. Any damping affecting inclination d is assumed to be zero.
Coefficient d used in Do = d/2mwo, is assumed to give Do a value of 0.05. For
radius T of the forced circular motion the response curve shown in Fig. 4.23 is
obtained, having a maximum Tmax = 10.6 e.
The peak value Tmax can be estimated by using either Do or the correspond-
ing modal damping ratio Dk (equation (24.23)).

3
..Q.
I
_----0.2
+-:::::.....==========0.35
0.5

O~---.----.---~.---

0.5
Fig. 4.20. Critical angular
velocity for model I with Id = 0.5 Ip
4.3 Unbalance Vibrations 55
c
I
0.1
f 3
Q e 0.2
Wo
0.3
2

ZS ~
I I.e I
0

,-
0 0.5 1 1.5
ip Fig. 4.21. Critical angular
velocity for model II with Id = 0.5 Ip

1.J...--~

0.5 1 1.5

----
ip
c
Fig. 4.22. Critical angular
velocity for model ill with Id = 0.5 Ip

10.6
10
.r.
e

Oo---~~-----r-----.----
o .n. _ Fig. 4.23. Orbit radius for
Q, unbalance excitation. For data see text
56 4 Shaft with Non Central Disc

Thus
1 1
Tmax ~ 2Do e or Tmax ~ 2Dk e (4.39)
The first gives Tmax ~ e/2 x 0.05 = 10e and the error is -6 %. For the second, the
eigenvalues and modal damping ratios of the model were calculated (Fig. 4.24).
Of these D2 is significant at resonance. For {} = (}q D2 = 0.047 and, using
equation (4.39) Tmax ~ 10.6 e, exactly the same as the value obtained by lengthy
calculation although such accuracy, is a coincidence. Further use of equation
(4.39) is discussed in Chap. 21.
The behaviour for an unbalance moment will be considered in Chap. 5 for
short and long rotors, and by so doing the rotor length will be shown to playa
decisive role.

0.047
~----f----- D2

0.02

Fig. 4.24. Modal damping ratios of the


1"1_ example.

4.4 Harmonic Excitation in one Direction

With unbalance excitation the vector of the exciting force or moment rotates
in synchronism with the shaft. Consider now the case that a harmonic force
of constant direction acts on the disc in the plane 1-2. To understand the
influence of this force on the rotating shaft it should be noted that a harmonic
unidirectional force can be understood as being the resultant of two counter-
rotating forces. Such forces are given in the arrangement shown in Fig. 2.7.
Mathematically, this fact is evident from the equation,

F cos wt = 0.5 F (e iwt + e-iwt ) . ( 4.40)

If one assumes that the harmonic force acts in direction 1, then the equations
of motion of the models of Fig. 4.1 are

mz +di + k22 Z - k23'I/J 0.5 F(eiwt + e- iwt )


(4.41 )
Id~ + (d-jlp{})~-k23z+k33'I/J O.
4.4 Harmonic Excitation in one Direction 57

The action of the excitation is immediately evident from these equations, if one
notes the results of vibration due to unbalance.
The term 0.5Fe jwt gives a rotating force in the direction of rotation, the
term 0.5Fe- jwt gives an equal force in the opposite direction. It is to be expected
that, due to the first term, and when the excitation frequency W coincides with
one of natural forward frequencies, then a resonance occurs at W = W2 or W4.
Due to the second term, backward resonance is to be expected for W = -WI
or -W3. The following example elucidates this and gives more detailed results.
Consider model I with the already assumed data. The angular velocity of the
shaft {} = 0.8 Wo and it is excited with force F(t), as in equation (4.40), in
direction 1. The resulting forced vibrations are to be calculated.
Using the assumed angular velocities, the forward natural frequencies

W2 = 1.033 Wo W4 = 2.716 Wo

and the backward natural frequencies

WI = -0.698 Wo W3 = -1.451 Wo

The calculation of forced vibrations is carried out using equations (4.41) and, as
shown in Fig. 4.25, these have amplitudes dependent on the exciting frequency.
They are non-dimensionalised by the static value z. = hllF. There are reso-
nances for all four natural frequencies and the peaks decrease with excitation
frequency. The maximum value 7.4 has an approximate value of 1/2Do = 10.

8
7.1.
&
X.
6
.&Xs
1..9

I.
3.6
x,
Xs
2
.&
Xs 1.5

0
0 -w, w1 -W3 w4
w-
Fig. 4.25. Amplitudes due to excitation in direction 1, model I. a = 0.35 I, ip = 0.4 I,
Id = 0.5 I p , Do = 0.05, {J = 0.8 Woo
58 4 Shaft with Non Central Disc

The path of the centre of mass S in the plane 1-2 is determined by the
displacements ZI(t,W), Z2(t,W). Without calculation we observe the following.
For very low excitation frequency W :::::J 0 the point W follows the force;
it moves in axis 1 with amplitude :::::J z . At resonance the influence of the
corresponding excitation dominates, whereby the point S describes roughlr a
circle, in a backward motion for W = -WI and -W3 and in a forward motion for
the other two resonances. The orbit is not exactly a circle at resonance because
the other direction of excitation also has an influence. Between the resonances
the orbit is elliptical and becomes a line as the direction of rotation changes, as
shown in Fig. 3.14.
We can generalize these results. When the excitation consists of elements
which are co-rotating and also of elements which are counter-rotating, then
resonance is possible having all four natural frequencies. This is the case, for
example, when the force vector describes an ellipse (see equation (17.4.
Excitation predominantly in one direction results, for example, from pis-
ton machines through the piston force. For machines such as that shown in
Fig. 4.26, the generator rotor and flywheel are excited with vibration in this
fashion through the shaft and bearings. There are many excitation frequencies
present,
wp=pil ( 4.42)

Motor Generator Flywheel

Fig. 4.26. Motor generator set

with order numbers p, which depend on the type of machine. The possible
critical speeds are obtained from the so-called Campbell diagram, Fig. 4.27. Each
point of intersection of a straight line p il with a curve Wk gives a critical angular
velocity ilpk of order p and of degree k, that is the critical speed npk = ilpk /27r.
The severity of the resonances occurring at these critical speeds depends mainly
on the strength of the exciting harmonic.
4.5 Shaft with Flexible Bearings 59

6Q 5Q 4Q 3Q 2Q

Q_ Fig. 4.27. Campbell diagram

4.5 Shaft with Flexible Bearings

If the bearings of the models in Fig. 4.1 are flexible and isotropic, then the force
influence coefficients and correspondingly the values of the natural frequencies
are altered. The dynamic behaviour, however, stays fundamentally the same
as for rigid bearings. Therefore consider the general case of anisotropic flexible
bearings and of different stiffnesses of the two bearings. With this assumption,
model I may be redrawn as shown in Fig. 4.28. Its bearings have stiffnesses

Fig. 4.28. Model I with anisotropic flexible bearings.


60 4 Shaft with Non Central Disc

kA1' . , kB2, all having values other than zero. It is evident that, for this model,
the same equations of motion are valid as for rigid bearings, except that the
force influence coefficients are different.
We calculate the new force influence coefficients as in Sect. 4.1 by inver-
sion of the displacement influence coefficients. With the displacement influence
coefficients,
for rigid bearings and flexible shaft
for a rigid shaft and flexible bearings
for flexible bearings and flexible shaft,

(4.43)
For a shaft with rigid bearings, three influence coefficients suffice. However, here,
because of the anisotropy of the bearings, six influence coefficients are needed,

hik for the shaft are obtained for a constant cross-section from the relationship
in table 4.1 and formulae (4.6a) and (4.6b).
The influence coefficients, as a result of bearing flexibility are, with a = a/I
and f3 = b/l,

h' 1 1
44 = kA112 + k BI [2
(4.44)
h' _ _1_ _ _1_
33 - k A2 [2 + kB212 .

For model II with the same flexible bearings as applied for model I in Fig. 4.28,
the following influence coefficients result, with 'Y = c/l.

h' 1 1
44 = kA112 + k BI [2
h' 1 1
33 = kA212 + k B2 [2
(4.45)
Hence all the required influence coefficients h:~ can be calculated for models I
and II with flexible bearings. The force influence coefficients follQw from

h" -h" h"


kn -! , k14 = -----1:! , k44 = -1!.
..:11 ..:11 ..:11
..:11 h~l h~4 - h~/
( 4.46)
h"33 -h" h"
k22 , k23 ~ , k33 --B.
..:12 ..:12 ..:12
..:12 h~2 h~3 - h~32 .
4.5 Shaft with Flexible Bearings 61

The equations of motion are obtained from equations (4.35) by insertion of the
relevant force influence coefficients and are as follows,

m:ih + di: 1 + kn:l:1 + k 14'P4 F1(t)


m:C2 + di: 2 + k22:1:2 + k 23'P3 = F2(t)
(4.47)
I d r(;3 + dv>3 + Ip n CP4 + k23:1:2 + k 33'P3 = M3(t)
I d r(;4 + dCP4 - Ip n cp3 + k14:1:1 + k 44'P4 = M4(t)

where F1(t), F2(2) are exciting forces and M3(t), M4(t) are exciting moments.
This system can be reduced from four equations to two equations with com-
plex coefficients, see Kellenberger [31], by introducing sum and difference ma-
trices. Without this reduction, one obtains, for harmonic forcing, a system of
linear algebraic equations for the amplitudes of displacement and inclination
(see Sect. 24.2).
One can make the following comments about the natural frequencies without
carrying out calculations. Because of the flexibility of the bearings, the number
of degrees of freedom has not altered. Hence four natural frequencies are present,
as for rigid bearings. For n = 0, the models have two natural frequencies W20
and W40 for rigid bearings. Because of bearing flexibility their values are smaller
and because of anisotropy, two pairs of natural frequencies result, as shown
in Fig. 4.29. The behaviour for unbalance excitation is shown in the following
example.

Example 4/1 Consider the rotor of an induced-draught blower as shown in


Fig. 4.30. This example is taken from reference [32], page 183. The amplitude
of vibration of the disc is required, due to unbalance arising from a mass eccen-
tricity e.
Bearings rigid Bearings flexible. anisotropic
w4

w4

w40

W40
-W3
w30
-w 3
W2

W2
w20
W20
-WI w10 -WI

Q- Q-

Fig. 4.29. Natural frequencies for rigid and flexible anisotropic bearings
62 4 Shaft with Non Central Disc

I
I
~-r-L--------------4-+-U. +--~ M

I
i
~------4m------~

Fig. 4.30. Rotor of an induced-draught blower.

The mass of the shaft is small in comparison with the mass of the disc, and
so the rotor can be approximated with sufficient accuracy by modell II. The
relevant data are,
m 8000 kg
Ip 8520 kg m 2 = 1.03m
Id = 4260 kg m 2 0.73m = 0.707ip
kAl 333 N / p,m = 83.3N/p,m
kA2 667 N/p,m 167 N/p,m
D 0.02
operating speed 740 rpm.
Fig. 4.31 shows the calculated natural frequencies !k = Wk/27r. They are
given in cycles/min so that they can be compared with the operating speed.

/
fk /
cpm /
/ fk = 2n
4000 /
/
/
3000 /
/
/
2000
I
1790--------- -f 3
1125 - - - - - - - - f2
1000

-f
5000 rpm
n-
Fig. 4.31. Natural frequencies of the rotor of Fig. 4.30.
4.5 Shaft with Flexible Bearings 63

For rigid or isotropic flexible bearings the graphs of f2 and - h have a common
asymptote h = wl./27r, where Wl. is obtained from equation (4.28). For the
anisotropic flexible bearings considered here, there are two natural frequencies
when disc inclination is prevented. fn = 1125 cpm is the asymptotic value for
h and f31. = 1790 cpm for -h. Critical speeds are given as abscissa. Both
the forward critical speed n2 of 787 rpm and the backward critical speed n3
of 1900 rpm playa role. This can be recognized from Fig. 4.32, which shows
the amplitude of vibration of point W in directions 1 and 2. Resonances occur
at nl, n2 and n3' In this example the first resonance is insignificant, but the
next two (forward and backward) are equally strong. This can be seen better in
Fig. 4.33, which shows the orbits of W at the critical speeds. Important lessons
to be learnt from this example are:

10
.&e !
""
':
&
e
.&e
!:ve:
:'
"
"'
Z

,," ,,
,,
I I
5 I,
,
,
\
\

\' ....
.... _-------
o O-~~_+_.----.-~rr--~----,_---.------

o n, nz
1000 3000 rpm
n-
Fig. 4.32. Amplitude of vibration of disc centre for unbalance excitation.

10e
Xz

x,

Fig. 4.33. Orbits of the disc centre at the critical speeds.


64 4 Shaft with Non Central Disc

For anisotropic flexible bearings the backward critical speeds are excited by
unbalance.

Hence models I and II have three resonance regions.

The strengths of the individual resonances depend on the model parameters.


5 The Short and the Long Rotor

In Chapter 4, the rotor was assumed to be a disc and, in the examples, this disc
was assumed to be thin and circular, with Id = 0.5 Ip. The equations of motion
quoted in Chap. 4, equations (4.8) and (4.47) are valid, however, for arbitrary
values of Id and Ip, that is for shafts having a thick disc or a long stiff rotor
body, as shown for example in Fig. 5.1. The vibrational behaviour of such a
model will be investigated in the present chapter.
The stiff part of the model is shown as a rotor. A rotor has, from Fig. 5.2,
the moments of inertia,

Id = h = III and Ip = IIII ,


where h, III and IIlI are moments of inertia about axes through the rotor cen-
tre. A rotor is called symmetric when III = 1/ and
short for Id < Ip
spherical for Id = Ip
long for Id > Ip .

Fig. 5.1. Models with stiff rotor body

Fig. 5.2. Coordinate system for moments of


m inertia of a rotor
66 5 The Short and the Long Rotor

5.1 Unbalance Moment

When the rotor axis is oblique to the rotational axis, that is the axis III of the
rotor makes an angle with the axis of rotation, then a moment is introduced due
to unbalance forces, and this alters the inclination. We shall call this moment
the centrifugal moment and it is determined as shown in Fig. 5.3.
Due to rotation with angular velocity {} the rotor has the following torque
components at the instant shown,

Dd = Id{}II
Dp = Ip{}III

For incremental rotation d'lj; = {} dt, the changes in Dd and Dp are, from Fig. 5.3,

d Dd = -DdCOS"{ {}dt
d Dp = Dpsin, (}dt
Hence the time derivative of the rotor torque is

dD
-
dt
d Dd
= --
dt
d Dp
+- -=
dt
(-Id + I p ) {} 2 COS"{Slll,

that is, for, : 1


D = (-Id+1ph{}2 . (5.1)
The centrifugal moment Mu is equal to -D, and so for, : 1,

(5.2)

where
(5.3)
The expression UM in equation (5.3) will be called the unbalance moment to
differentiate it from the term unbalance U = me.

1\\
2 2 2

Dd - --------

Fig. 5.3. Torque components and their time derivatives


5.1 Unbalance Moment 67

Unbalance moment has the units kg m 2 , while centrifugal moment has Nm.
In Fig. 5.3 the centrifugal moment vector is coincident with axis 1 and rotates
in plane 1-2.
To obtain a better understanding, we derive again the equation for the
centrifugal moment but in a simpler way, by means of an example. The model
of Fig. 5.4a is considered. The rotor consists of a symmetric star of six stiff
massless beams each of length r at whose end a point mass is situated. The
rotor is supported on a massless shaft. It approximates to a spherical rotor for
which,
Id = Ip = I = 4 mr2 .
For Id - Ip = 0 the centrifugal moment is always zero (equation (5.2)), for any
inclined position.
Now increase the four masses in the plane 1-2 from m to m + Lim and
incline the rotor by a small angle 'Y (Fig. 5.4b). Referred to the axis of the star
we now have,

and the moment of the centrifugal forces is

The centrifugal moment is negative and it reduces the inclination 'Y.

Fig. 5.4. Spherical rotor with added masses.


68 5 The Short and the Long Rotor

In the second case, increase the two masses which lie on the axis by Llm
and likewise give the rotor the small inclination 'Y (Fig. 5.4c). The moments of
inertia are now,

and the centrifugal moment is

The centrifugal moment is positive and so it increases the inclination 'Y.

5.2 Natural Frequencies

N&tural frequencies of models with a thin disc, that is with Id = 0.5 Ip, were
described in detail in Chap. 4. Here it will be ascertained what influence the
rotor length, that is the relation Id/lp, has on these frequencies. To simplify
proceedings we confine ourselves to answer this question for models, whose
displacement and rotation are not coupled. It is thus assumed that the cross
coefficients k23 and k14 are zero. The bearings can be stiff or anisotropically
flexible. Damping is neglected. Thus, from equations (4.47), the equations for
natural vibration are,

m:ii 1+ kn :1)1 = 0
m:ii 2+ k22 :1)2 = 0 (5.8)
Id CP3 + Ip n t{;4 + k33 CP3 = 0
Id CP4 - Ip n t{;3 + k44 CP4 = 0
The first two equations give the natural frequencies for displacements :1)1, :1)2

namely,

W1=~ W2={:i.
The last two coupled equations for rotations give the equations

k33 - Id w2
[ (5.9)
-j Ip nw
where CP3 = ~3 ejwt , CP4 = ~ 4&"'t, in which ~3 and ~ 4 are complex.
With natural frequencies (for n = 0),

fk;
W30 = VI; (5.10)
5.2 Natural Frequencies 69

one obtains the characteristic equation

W
4- [2 2 (Iv) 2rl2] W 2+ W 302 W 402 = 0
W 30 + W 40 + Id U (5.11)

After some manipulation we obtain,

1 1 1 1
-[
4
1- -2 W30 W40 =f "4 [ 1+ '2 W30 W40 (5.12)

where [ 1is the expression in brackets in equation (5.11).


For stiff or isotropic flexible bearings k44 = k33 and one obtains the following
equation, which corresponds to equation (4.25)

(5.13)

In order to ascertain the effect of rotor length on the natural frequencies, the
simplified model of Fig. 5.5 is chosen. The rotor is symmetric about the shaft
centre and is supported by rigid or isotropic flexible bearings of stiffness k =
hI = k 2
The moments of inertia are

, Iv = 2"m R 2
(5.14)

The stiffnesses for rigid bearings are

6EI
P (5.15)
kll(~+lr

EI m X2 f
%/ ~_
~
k,
X, 'P3
k2

Fig. 5.5. Symmetric model.


70 5 The Short and the Long Rotor

For isotropic flexible bearings of stiffness k, the stiffnesses (5.15) should be


multiplied by the factor
2k
K,=--- (5.16)
kll + 2k
The natural frequencies WI and W2 are equal and independent of n. For n= 0,
W3 and W4 are also equal and the following relationships are valid

W30 W40
-=-= (5.17)

Putting v = n /W30, the natural frequencies W3 and W4 non-dimensionalised by


W30 are given from equation (5.13) by,

W3

W30
2iv)2 +1.
(1 1 (5.18)

The ratio of moments of inertia is seen to be the controlling parameter. Fig. 5.6
shows these non-dimensional natural frequencies versus non-dimensional angu-
lar velocity for

- a thin disc with Id = 0.5 Ip, that is L = 0

- a spherical rotor with Id = Ip, that is L = v'3R

- a rotor with Id = SIp, that is L = ..j27R.

o~------~------~--------~
o 2

Fig. 5.6. Natural frequencies of the model of Fig. 5.5.


5.3 Unbalance Vibrations 71

It can be seen that the gyroscopic effect decreases with increasing slenderness of
the rotor. This follows from equations (5.8), according to which, for harmonic
motion, the gyroscopic moment behaves with respect to the accelerating mo-
ment in the proportion Ip/ld. According to equation (5.14) Id/lp characterises
the slenderness of the rotor.

5.3 Unbalance Vibrations

In Chapter 4 the unbalance vibration due to mass-eccentricity of models with a


thin disc was investigated. It transpired that, for isotropic bearings, one critical
angular velocity exists, and for anisotropic bearings, three exist. For long rotors
having Id > Ip a further critical angular velocity emerges, because the asymptote
to the graph of W4 has a lower slope than that of the straight line w = il. As a
result, there exists a value il = W4 = il 4 We investigate this in a simple way
on models whose displacement is again uncoupled from the tilt deflections, but
which may be assumed to be supported on isotropic bearings.
From equations (5.8) the critical angular velocities for displacements :1)1 and
:1)2 are,

ill = W1 = ! il2 = W2 = ~ . (5.19)

For the tilt deflections '()3 and '()4, the critical angular velocities are obtained
from equation (5.11) with w = il and Id =I- Ip, as follows,

IJ 1(2 2) + W302 W402 (Ip)


IJ - T; '2 W30 + W40 =f 4:1 (w30
2
-
2 )2
w 40 Id
2

(5.20)
Resonance for these angular velocities occurs only for Id > Ip as will later be
shown in more detail.
For isotropic bearings, that is for W40 = W30,

(5.21)

Example 5/1 It is required that natural frequencies, critical angular veloc-


ities and unbalance vibrations be determined for the model of Fig. 5.5, with
anisotropic bearings. Relevant data are

1= 2 R , L = 10 R , k1 = 0.2 ko , k2 = 0.5 ko
with ko as the stiffness kn, or k22 for stiff bearings. The rotor of the generator
of a large turbo machine in a power station can have such a geometry. Hence
its vibration behaviour can be explained with the results of this model.
72 5 The Short and the Long Rotor

The calculated natural frequencies and critical angular velocities are pre-
sented in Fig. 5.7. For rigid bearings, we have the natural frequencies:
w; the first natural frequency for general n
wio the second natural frequency for n = 0
and the three critical angular velocities

n; = w; , ni = 0.96 wio, n: = 1.05 wio


For flexible bearings the natural frequencies are,

0.53 w;, W2 0.71 w; for general n


1.28 w;, W40 1.69 w; forn=O.

The critical angular velocities are

WI
0.998 W30

It can be seen that the gyroscopic effect is very weak for this model. This is not
surprising as the polar moment of inertia governing gyroscopic action amounts
to only 5.8 % of the diametrical moment of inertia through the centre of gravity
(see observations in Sect. 5.2).
The amplitudes of displacements :1:1, :1:2 as a result of an unbalance U = m e
are shown in Fig. 5.8. With this type of excitation the model behaves like the
Jeffcott rotor of Sect. 3.3. Resonance occurs at n l and n2 and whirling is in
the direction of shaft rotation. For D = 0.05, with D = d/2m wi the maximum
amplitudes are Zlmax = 5.4 e and Z2max = 7.1 e.

----- wZ
--============:- T------w;
W40+-------"7f------

Q-

Fig. 5.7. Natural frequencies and critical angular velocities of the model of Fig. 5.5.
For data see text.
5.3 Unbalance Vibrations 73

8
~
.&e

+'"'+
6
~
e
4

2 II, liz

0 i i
II, Qz
Q-

Fig.5.8. Amplitudes and orbits at resonance for the example having excitation
U=me.

For excitation by an unbalance moment UM = (Id - Iph, resonance occurs


at ,03, ,04. Fig. 5.9 shows the amplitudes of the displacements at the end of the
rotor in directions 1 and 2

non-dimensionalised by
L
:Vo = /2" '
the displacement at the end of the rotor in terms of /. The damping ratio was
assumed to be D= 0.05, where D= d
12Id wio. Hence the maximum amplitudes

8
~i

~ ~

+'"'+
Xo
~i
6
~
Xo

2 113 Q4

0 i i
Q3 Q4
Q-

Fig. 5.9. Amplitudes and orbits at resonance for the example having excitation
UM = (Id - Iph
74 5 The Short and the Long Rotor

are found to be

:il~max = 4.6 :1:0 and :il~max = 7.4 :1:0

il3 represents backward whirl and il4 forward.


For excitation at the end of the rotor by a small unbalance mass Lim at
radius R, the unbalance U = Lim R and the unbalance moment UM = UL/2.
Thus resonances can be excited at all four critical angular velocities.
With the results of this example and using Fig. 5.6, we can conclude that for
long stiff rotors for which roughly Id > 10 Ip the gyroscopic effect has practically
no influence on the unbalance vibrations.

The effect of an unbalance moment will now be generally ascertained for the
case of a symmetrical model with isotropic bearings. For this case the equations
of motion (4.36) are valid, and are uncoupled because k23 = O. Putting f3 = 0,
the complex tilt "p is given by

(5.22)

with the particular solution

(5.23)

U sing the terms

W30
-Ii
- Id' W8 =
~ Ip - Id ' WI =
~ Id - Ip

/.I
il il il
/.11=-
(5.24)
/.18=-'
W30 W8 WI

D -d-
-2Idw30
we have for the short rotor with Id < Ip

(5.25)
/.I~ 2D/.I
tanO! = - -
Va - V(1 + /.1;)2 + (2 D /.1)2 ' 8 1 + /.I~
5.3 Unbalance Vibrations 75

and for the long rotor with Id > Ip

(5.26)
2Dv
tanal = --2 .
1 - vI

The positions of the arrows for the angles 7 and tJi are shown in Fig. 5.10. The
magnitude of tJi depends for a short rotor on V. and for a long rotor on Vi.
Fig. 5.11 shows that V. increases with speed monotonically from zero to one.

ImljJ

Fig. 5.10. Positions of the arrows for the angles 7 and tP for the short and the long
rotor.

Q
W30-

Fig. 5.11. Amplification factor V. D = O.


76 S The Short and the Long Rotor

Therefore an unbalance moment excites no resonance for a short rotor. For long
rotors a resonance exists, as Fig. S.12 shows for a rotor with Id = 1.1 Ip, 2Ip
and SIp.
From equations (S.26) the maximum of V, is close to VI = 1, that is at
n = WI, where from equations (5.10) and (5.21), n = n4 The corresponding
amplification factor becomes

(5.27)

Fig. 5.12 shows that with constant damping ratio D= 0.05, the resonance curve
becomes flatter with decreasing Id/Ipo The dotted line is the resonance curve for
a rotor with Id = 1.1 Ip and zero damping. This shows that rotors with Id ~ Ip,
where Id > Ip and with small damping have a rather flat resonance region.
The resonance described is excited by the rotating unbalance moment, that
is by the centrifugal moment of constant value. This resonance is at natural
frequency W4 and is in the direction of shaft rotation. A harmonically varying
moment, but of constant direction can be separated into constant co-rotating
and counter-rotating terms, each of half the amplitude of the harmonic moment.
Hence the resonance consists of backward whirl at Wa and of forward whirl at
W4. Similar relationships exist to those described in Sect. 4.4 for a harmonic
force. A moment excitation, with constant direction occurs, for example, when
a bearing is excited in one direction in a harmonic fashion, with 1800 phase
difference from the excitation at another bearing.

f 10 I
I
I
I
VI I
I
\
\
\
\ 1.1
5 \

04-~~r=~~-----.-----r----
o 2 3 4

= 0.05, . - - D = o.
~

Fig. 5.12. Amplification factor Vi. - D


6 Oil-Film Bearings

Most rotors are supported in oil-film bearings or in rolling-element bearings. For


maintenance free operation and to give high speeds, gas and magnetic bearings
are also increasingly being used. The bearings influence the rotor vibrations to a
greater or less degree by their dynamic properties in comparison to those of the
rotor and the remaining components of the assembly. This influence originates
essentially from the ratio of the rotor stiffness to the bearing stiffness. For a
relatively elastic rotor the influence of the bearings is small, while the dynamics
of a stiff rotor are determined in the main by the bearings.
This chapter will be concerned with the static and dynamic properties of
oil-film bearings. This subject is very complex and in the last four decades
a large number of significant contributions have appeared in the literature. To
give these here would be too space consuming. We confine ourselves therefore to
the essential theoretical fundamentals and consider especially the stiffness and
damping coefficients, through which the oil-film bearings will be considered
using conventional rotordynamic theory.

6.1 Hydrodynamic Bearing Theory

The components used in the theory of hydrodynamic bearings are the bearing-
journal, the lubricant and the bearing shell. The journal rotates with angular
velocity a and is statically or dynamically loaded in the radial direction. The
bearing shell is rigidly supported. The relation between the force and the dis-
placement of the journal is required. The properties of the lubricant and the
geometry of the clearance will playa considerable part in this relationship. We
confine our consideration of the clearance geometry to the type of bearing shown
in Fig. 6.1: The bore should be constant along the axis in any radial direction,
that is consist of circular arcs. The cross-section should be approximately or
exactly circular. For a non-circular bore we establish a reference circle of radius
R. Its centre is at the midpoint of the bearing length L and is denoted by B a
a
The point B is the origin of the coordinate system X, Y, Z. The bearing jour-
a
nal has radius r and centre J in the XY -plane. It is assumed for simplicity
that the journal is parallel to the bearing bore and as well as rotation under-
goes only translational motion parallel to the XY -plane. In other words, any
moments arising from journal tilt are neglected.
78 6 Oil-Film Bearings

y y

Fig. 6.1. Fluid-fUm bearing. Diagrammatic layout of assumed clearance geometry.

The journal is unloaded when OJ and OB are coincident. The radial distance
between the journal surface and the bearing bore for a central journal is termed
the oil-film thickness hOe In general ho differs around the circumference, that is
ho = ho(rp). For a circular bearing bore ho = R - r, and is constant. This value
is here denoted by bearing clearance

b=R-r (6.1)

for an arbitrary bearing bore.


The position of the journal is defined by the eccentricity e and the angle '1
(Fig. 6.1). For an eccentric journal the oil-film thickness for e : r is given by

h(rp) = ho(rp) - ecos(rp - '1) (6.2)

and for a moving journal with e(t) and 'Y(t)

h(rp,t) = ho(rp) - e(t) cos [II' - 'Y(t)] (6.3)

The expression (6.3) is called the clearance function. Further, we assume,

1. The lubricant is massless, incompressible and adheres to the bearing sur-


faces.

2. The lubricant is Newtonian and its viscosity is constant in the whole of


the oil-film.

3. The flow is laminar.

4. The pressure of the lubricant is constant in the radial direction.

5. Flow velocity in the radial direction is neglected.


6.1 Hydrodynamic Bearing Theory 79

6. Velocity gradients in the radial direction are large in relation to those in


the tangential and axial directions.

7. The oil-film thickness is small compared to the journal radius.

8. The curvature of the oil-film is negligible.

9. The bearing surfaces are smooth and stiff.

Using assumption eight the oil-film clearance is developed out into one plane
(Fig. 6.2). It is therefore defined by the coordinates :1:, y, z and the function
h(:I:, t). The bearing shell is at rest and the journal has surface velocity U = ilr,
for angular velocity il. The lubricant has different velocities u, v, w in :1:- y-
z-directions and these will depend upon position in the oil-film.
For the fluid element d:l:, dy, dz the condition for force equilibrium is
aTzy op
--=- (6.4)
oy az
with p the lubricant pressure and T",y, Tzy the shear stresses.
For the shear stresses from assumption two the following equations are valid,
au ow
T",y = '1/ ay T zy = '1/ oy (6.5)

with '1/ the dynamic viscosity. Using equations (6.5) and (6.4) we have

(6.6)

and after integration,


1 op 2
U = 2'1/ 0:1: Y + Cly + C 2 (6.7)

v
y
u Journal
Oil film

Bushing X,u

X,u

z
w Fig. 6.2. Coordinates of oil-film.
80 6 Oil-Film Bearings

With the boundary conditions

u = 0, w = 0, for y = 0
u = U, w = 0 for y = k ,

then
U = -18 -
p (2
y - ky ) + -U
y 1 8p( 2
w = 2." 8z y - ky
)
(6.8)
2." 8~ k
Besides the condition of force equilibrium, the continuity equation must be
fulfilled. For this consider the elemental volume shown in Fig. 6.3. With
h

Qx = JUdY
o
as the fluid flow rate in the ~-direction and

J
h

Qz = w dy
o
as the fluid flow rate in the z-direction, the continuity condition is

8Qx 8Qz 8k
8~ d~ dz dt + 8z dz d~ dt + at dt dz ~ = 0 ,

that is
8Qx 8Qz 8k _ 0
8~ + 8z + 8t - (6.9)

Using equations (6.8),

. k 3 8p 1
Qx=-- -+- Uk (6.10)
12." 8~ 2

z Fig. 6.3. Fluid element for continuity


6.1 Hydrodynamic Bearing Theory 81

whose differentials with respect to z and Z in equation (6.9) lead to

~ (h 3 op) +'!. (h 3 op) = 6 U oh + 12 oh , (6.11)


oz 1J oz oz 1J oz oz ot
the so-called Reynolds Equation.
For further calculations we assume again the original geometry shown in
Fig. 6.1. With z = cpR and h(cp, t) from equation (6.3),

oh
oz
oh
ot
For r ~ R and constant viscosity 1J the Reynolds Equation takes the following
form

~ ~ (h 3 op) + h3 02p = (6.12)


R2 ocp ocp 0 Z2

6 11 [{} ~~ + e ({} - 2'1) sin (cp - ')') - 2e cos (cp - ')')]

With this equation and the appropriate boundary conditions and initial condi-
tions, the pressure p in the oil-film is determined. Integration gives the pressure
p(cp, z, t), which depends upon the position cp, z in the oil-film. For a trans-
versely moving journal this pressure is also a function of time.
The right-hand side of equation (6.12) shows the required conditions for
the formation of this oil-film pressure. In any case the existence of viscosity
is essential. Furthermore, at least one term in the square brackets is required.
Hence for a circular bearing whose rotating journal is centrally positioned, the
oil-film force is zero. For a non-circular bore the term oh%cp is responsible
for pressure build up and hence an oil-film force even for a cen.trally positioned
journal. For translatory motion of the journal, a dynamic oil-film pressure is in
general produced. Only for e = 0 and '1 = {} /2 can the pressure become zero
for a circular bearing.
The oil-film pressure field gives the incremental oil-film force dF when
integrated over the length L and multiplied by R d cp. Thus

dF = [7/2
-L/2
p(cp, z, t)dZ] R d cp .

This has components

dF1 = dF coscp , dF2 = dF sincp ,


82 6 Oil-Film Bearings

as shown in Fig. 6.4. Integration over the circumference gives the resulting force
components

(6.13)

The magnitude and angle {} of the resultant oil-film force Fa are given by

(6.14)

The oil-film force Fa is applied on both the bearing shell and the journal. To
ensure equilibrium a corresponding force FJ must be applied externally on the
journal, which is equal in magnitude and direction to Fa.

Fig. 6.4. Resultant oil-film force

For rotordynamic analysis the additional force is needed which arises due to
translatory motion of the journal. In general the relation between the force and
the motion is non-linear. For small displacements and velocities (Fig. 6.5) one
may however linearise. With X o , Yo as the displacements of the journal centre
as a result of the static load Fo = (FlO, F20 ), the journal force FJ = (F1' F2)
thus has components
OF1 OF1 oFl . oFl .
F1 = FlO + -0
:1;1
:1;1 + -0
:1;2
:1;2 + -0
:1;1
:1;1 + -0.
:1;2
:1;2 = FlO + LlFl
(6.15)
oF2 OF2 oF2 . oF2 .
F2 ~= F20 + --
0:1;1
:1;1 + --
0:1;2
:1;2 + -.-
0:1;1
:1;1 + -.-:1;2
0:1;2
= F20 + LlF2
The partial differentials with respect to displacement and velocity are the stiff-
ness and damping coefficients, respectively.
6.2 Short Bearing Theory 83

Yo -------.iJ,.:.--.L.--__
e :CJ x, x,
I
I
I Fig. 6.5. Coordinates for stiffness and
X F, damping coefficients.

6.2 Short Bearing Theory

In order to solve the Reynolds Equation and to calculate the static and dynamic
displacements a great deal of numerical work has appeared in the literature. We
shall only mention here references [33, 34] due to Someya and the reviews [35]
and [9] by Lund and Mitsui in Sect. 4.1 of reference [10].
As a rule calculation has to be carried out numerically. A closed algebraic
solution by Dubois and Ocvirk [36], however, allows the following qualitative
conclusion: For a relatively short bearing the change in pressure in the circumfer-
ential direction is small in relation to the pressure change in the axial direction.
Hence, in the Reynolds Equation (6.12), one may, to a reasonable approximation
remove the first term. Comparison of calculated results and measurements of
the displacement of the journal under static load for bearings having circular
bore show pleasing agreement up to a ratio of length to diameter of unity.
The determination of stiffness and damping coefficients from this theory has
been carried out by Holmes [37] and Smith [38]. In the following, calculations
will be carried out with somewhat different notation and procedures from those
followed in references [37, 38].
For a circular bearing the oil-film thickness is the same as the bearing
clearance. Hence ho = 6 and thus {}ho/{}cp = 0, that is the clearance function
(6.3) is
h(cp,t)=6-e(t)cos[cp-,(t)] . (6.16)
With {}p/{}cp = 0, the Reynolds Equation (6.12) takes the form

(6.17)

By integration one obtains the pressure function


84 6 Oil-Film Bearings

and with the conditions


L
dp = 0 ,or z = 0,
dz
p = 0 for z = -2
it follows that

p(cp, z, t)=!~[e(n-21)sin(cp-1)-2e COS(cp-1)] (Z2_~2). (6.18)

Integration over z gives the oil-film force per unit length in the circumferential
direction.
+L/2
J p(cp, z, t)dz 'TIL n_
3

q(cp, t) = = 2(52 q(cp, t)


-L/2
with (6.19)

21 ).
e ( - - 1 sm(cp-1)+-COS(cp-1)
2e
q(cp, t) = n n 3
[l-e COs(cp-1)]

and
e
e=- (6.20)
5
-as the eccentricity ratio.
In the static case, with i, 1 = 0 and n i= 0, q is constant with time. With
u = II' - 1 then

_( ) -e sinu
qu = 3 , (6.21)
(1- e cosu)

Fig. 6.6 shows the behaviour of q( u) for several values of e. In the region 0 < u < 71"
q(u) is negative. This corresponds to a tension force in the lubricant, which is
impossible to sustain. One therefore assumes in this region that q(u) = O. Hence
one obtains the pressure distribution shown in Fig. 6.7. Positive pressure occurs
in the lower half as a result of the eccentricity vector e in the plane X' Y'. The
resultant of forces dF = q Rdu loads the bearing shell and the journal. The
corresponding static equilibrium force on the journal is shown as Fo in Fig. 6.7.
In the dynamic case, with i, 1 i= 0 the components of the oil-film force,
that is of the journal force in the X', Y'-system are
2,..
F{ ,..J q(u, t) cos u R du = F,., b(e, e, 1)
2,..
(6.22)
F'2 J q(u, t)sinu R du = F,., !2(e, i, 1)
6.2 Short Bearing Theory 85

Fig. 6.6. Function q(O'} from equation (6.21).

~
o

Fig. 6.7. Short circular bearing. Forces due to static loading.

with
F. _ 1/ L3 {} R
'1- 262 (6.23)

q(O', t) is obtained from equation (6.19) and 0' = cp - ,. Integration gives

.. ( e 1+ 2e
n21') (1_e2e
2 2
h(e, e, ,)= 1- 2)2+'/!' {} (1_e 2)5/2
(6.24)
21' ) e 4e
1 - n (1 _ e2)3/2 - {} (1 _ e2)2
'/!' ( e
h(e, e, 7) = -"'2
The force components relative to the X, Y -system are

F{ cos, - F; sin,
(6.25)
F{ sin, + F; cos,
86 6 Oil-Film Bearings

With equations (6.22) to (6.25) the explicit relations for the magnitude and
direction of the journal force are obtained as functions of position g, , and of
velocities i, l' of the journal centre CJ. In particular, equations are obtained for
static loading and for stiffness and damping coefficients for any journal position.
For a horizontally supported shaft with static loading the following are the
component forces

o F{ cos,- F; sin,
(6.26)
-Fo F{ sin, + F; cos,
with
2 g2
, 11(g) = (1 _ g2)2
7r g
(6.27)
, h(g) = -"2 (1 _ g2)3/2
The angle between force and eccentricity is obtained from Fig. 6.8 and equations
(6.27) as

(6.28)

and the angle, in Fig. 6.7 is obtained from the first of equations (6.26) as

F' 4 g
tan, = ~ = -- ----,-:-::- (6.29)
F~ 7r (1 _ g2)1/2 .

Finally, it follows from the second equation of (6.26) and from equations (6.27)
and (6.29), putting

tan, 1
VI
sin, = -r=====
+ tan 2 ,

IX
+" Fig. 6.8. Special case of a static load
6.2 Short Bearing Theory 87

that the relative static force Fol F" is given by

Fo
FfJ = S * ="27r c
(1 _ c 2)2
J 1 - c2 +
(4;: c)
2
(6.30)

The reference force F" can be identified as a friction force on the journal, in the
following way. The friction torque for a central journal is given by

assuming r ~ R, and the friction force FF is given by

(6.31)

The reference force F" in equation (6.23) is thus identical to the friction force
FF for a bearing with reduced clearance

6' = 7r 1/; 6 (6.32)


fP ,
where
the clearance ratio
R (6.33)
L L
- - D ' the length to diameter ratio.
2R
The force ratio Fol F" on the left-hand side of equation (6.30) is a modified Som-
merfeld Number. In German literature, the Sommerfeld Number So is usually
given by
~ Fo 1/;2 p'lj;2
/:)0= - - - - = - (6.34)
2RL "If} "If}
and so
; : = S* = ~~ . (6.35)

The behaviour of S* with eccentricity c is shown in Fig. 6.1l.


To calculate the stiffness and damping coefficients one needs the differentials
of FI and F2 with respect to Xl and X2 and with respect to Xl and X2 (see
equations (6.15)). From equations (6.25) FI and F2 are functions of c, , and
e, i. The required differentials are thus,
of; of; oc of; a,
-=--+--
OXk 8c OXk a, OXk
of; of; oe of; oi (6.36)
-=--+--
OXk oe OXk oi OXk
where i and k are equal to 1 and 2.
88 6 Oil-Film Bearings

The displacements dzt, dZ 2 in the X, Y -system correspond to the displace-


ments de, ed'Y in the X', Y'-system. From Fig. 6.5 it follows that

de dZ 1 cOS'Y + dZ 2 sin '1


(6.37)
ed'Y -dz 1 sin '1 + dZ 2 cos '1

from which

1 .
-- Sln'Y
e
(6.38)
1
- COS '1
e

With these relationships one finally obtains after transformation the following
relationships for the stiffness and damping coefficients of the short circular (or
cylindrical) bearing

(6.39)
where

'111 = [211"2 + (16 - 11"2) e2]A(e)


11" 11"2 - 211" 2e2 - (16 - 11"2) e4
'112 4 e(1_e2)1/2 A(e)
(6.40)
11" 11"2 + (32 + 11"2) e2 + (32 - 211"2) e4
'121 -4" e (1 _ e2)1/2 A(e)
11"2 + (32 + 11"2) e2 + (32 - 211"2) e4
'122 1- e2 A(e)

{312 = {321 =- [211"2 + (411"2 - 32) e2] A( e) (6.41)

11" 11"2
-
+ (48 - 211"2)e 2 + 1I"2e4
A(e)
2 e (1 _ e2)1/2

(6.42)

The non-dimensional stiffness and damping coefficients are functions of eccen-


tricity e. From equation (6.30) they can also be represented as functions of
Sommedeld Number S*. Their graphs and a discussion of their practical signi-
ficance is given in Sect. 6.3.
6.3 Static and Dynamic Properties 89

6.3 Static and Dynamic Properties

Journal bearings can have different lubricants and different clearance geome-
tries. Liquids are, in general, used for the lubricant and the following features
relate only to such bearings. The clearance can have a completely specified
geometry or be formed from movable segments. Fig. 6.9 shows some bearing
types.
Bearings with prescribed clearance (Fig. 6.9 a-e) have, in principle, the
same static and dynamic behaviour as the short cylindrical bearing. If we load
the rotating journal with the static load Fo and the journal has an angular
velocity {} = 27m, then the journal centre CJ will take up a position defined by
eccentricity vector e from the bearing centre CB such that its direction relative
to the load direction is a: (Fig. 6.10). For the short cylindrical bearing this

-' - Rs.
r~
..
Y
b

-$~-
r
..

d
Fig. 6.9. Some bearing types
a) Segment bearing e) Unsymmetrical three-lobed bearing
b) Lemon bearing f) Unsymmetrical tilt-pad bearing
c) Pocket bearing g) Symmetrical tilt-pad bearing
d) Three-lobed bearing
90 6 Oil-Film Bearings

Fig. 6.10. Journal position under static load

so-called attitude angle a and the Sommerfeld Number S* are obtained from
equations (6.28) and (6.30) as functions of eccentricity ratio = e16. S* and
a are shown graphically in Fig. 6.11. Fig. 6.12 shows the journal static load-
deflection curve. It gives the position of the journal centre CJ for values of S*
between 0 and 00. For S* = Fol Frj and Frj from equation (6.23), Fig. 6.12 can
be interpreted as follows:
For constant static load Fo, the journal centre C J is coincident with the
point C z (Fig. 6.10) when {} = O. With increasing angular velocity the
point CJ moves along the static load curve and reaches the bearing centre
CB when {} = 00.
For constant angular velocity {}, the journal centre CJ is coincident with
the bearing centre CB when Fo = O. With increasing static load Fo the
journal centre moves along the static load curve and reaches the point C z
when Fo = 00.

s*

0.1/

0.01~--+---+---+---+-----1

I I I I I
o 0.2 0.4 0.6 O.S

E---
Fig. 6.11. Short cylindrical bearing. Attitude angle and modified Sonimerfeld Num-
ber.
6.3 Static and Dynamic Properties 91

C 0.1
o Bf 0.5

0.2
E

!
0.4
0.6
0.8
1
Cz Fig. 6.12. Short cylindrical bearing static-load curve.

Similar statements can be made about the influence of bearing parameters


L, R, 5 and viscosity 1/ on the value of F7J"
Oil-film bearings with other profiles behave in a similar way to the short
cylindrical bearing. The form of the static load curve and the eccentricity for a
certain value of Sommerfeld Number S* can, however, differ significantly. This
is shown in Figs. 6.13 and 6.14, which are taken from the measurements of
Glienicke [39]. The following bearing types were investigated
K cylindrical bearing with 1jJ 1.5 01 1
100

Z lemon bore bearing 2.1 0/00


T pocket bearing 1.33 0/00
SD symmetrical three lobed bearing 1.0 0/00
UD unsymmetrical three lobed bearing 1.44 0/00 .
For all bearings the bore diameter was 120 mm and the length to diameter ratio
0.8. Fig. 6.13 shows that the sideways deflection for the lemon-bore bearing is
greater and for the three lobed bearing smaller than for the cylindrical and
pocket bearings. Fig. 6.14 shows that a certain value of S* gives very different

Fig. 6.13. Static load curves of several bearing types,


from reference (39)

1 0/00 denotes parts per thousand


92 6 Oil-Film Bearings

10+---+---+---+-~r-~

0.01 +---+---+---+---+-~

o 0.2 0.4 0.6 0.8


E-

Fig. 6.14. Sommerfeld Number S* against e for several bearing types, from reference
[39]

eccentricity ratios for the different types of bearing. The eccentricity ratio for
all the bearing types investigated is always greater than that for the theoretical
short cylindrical bearing (dashed curve).
For dynamic loading of the journal, there results a corresponding movement
of its centre C J within the clearance. For small displacements and velocities the
components of the additional force are

LlFl = k11Xl + k12X2 + d11Xl + d12X2 (6.43)


LlF2 = k21Xl + k22X2 + d21Xl + d22x2

The coefficients of stiffness and damping can be obtained from equations (6.39):

(6.44)

For the short cylindrical bearing '"fik and {3ik are functions only of eccentricity
ratio e and are found from equations (6.40) to (6.42). S* also depends upon e
and so the coefficients '"fik and (3ik are also functions of S* , which is advantageous
for dynamic calculations. These functions are shown in Figs. 6.15 and 6.16.
For a better demonstration we consider the stiffness and damping forces of
the bearing for certain displacements and velocities. For displacement Xl the
spring force F( Xl) is given by

(6.45)
6.3 Static and Dynamic Properties 93

f 100~--~--~---+--~~
lik

0.1 +----+--+---+t--+--I--

I Ii ''''I . , .... hi i '''''''I i .iililll i I

0.01 0.1 10 100


5*-
I I i

0.05 I 0.4 I 0.8 0.9 Fig. 6.15. Dimensionless stiffness coefficients


0.2 0.6 _ for the short cylindrical bearing

t 100 +--->-<--+-------+---I--__+_
f3 ik

0.1 +----+--+----+--1--

j i """'I I i """I . Ii ""'1 """'I i i

0.01 0.1 1 10 100


5*-
I i I

0.05 I 0.4 I 0.8 0.9 Fig. 6.16. Dimensionless damping coefficients


0.2 0.6 _ for the short cylindrical bearing

having magnitude

(6.46)

and angle relative to the xl-direction


/'21
tan {}l = - - - (Fig. 6.17). (6.4 7)
/'11
94 6 Oil-Film Bearings

x,

F(x,) Fig. 6.17. Spring force for displacement :1:1

Correspondingly the spring force arising from displacement X2 is

and from velocities Xl and X2 the damping forces are

with magnitudes and angles from relationships corresponding to equations


(6.46) and (6.47). The specific spring forces are thus proportional to the co-
efficients

II = V,fl + I~l (6.48)


and the damping forces to the coefficients

(6.49)

These are shown to scale and at the corresponding angles in Fig. 6.18 for the
short cylindrical bearing, for displacement and velocity at different eccentric-
ity ratios. It can be seen that in general the force vectors are inclined to the
displacement or velocity vectors; they are not collinear.
For displacement Xl the angles range from 60 to 70 for the eccentricity
ratios considered. For displacement X2 the angles change markedly with eccen-
tricity ratio. For c = 0.66 the force has the same direction as ;];2. Also the
damping force vectors are not collinear with the velocities, as the figures for f3l
and f32 show. These properties can be the cause of rotor instability (see Sect.
7.2).
The bounds of linear behaviour can be recognized from the following load
conditions. A journal having a static load Fo is additionally loaded with a force
F., whose vector slowly rotates. For an angle of rotation cp = wt the components
of the additional force are

LlFl = Fr cos cp LlF2 = Fr sin cp (6.50)


6.3 Static and Dynamic Properties 95

-------.- X2- 70

E=0.8
0.6"
0:002
G
,
5

E= 0.2
X1
0.4 <-, E=0.2 X2
0.6 0.8 '0
0.8 0.6
0.4
s

X1
E=0.8 0.6 0.4
5 -------
Fig. 6.1S. Coefficients of forces F(zd, F(Z2) and F(:i:d, F(:i: 2).

For w ---t 0, the damping forces can be neglected and one obtains by inversion
of equations (6.43) the displacement components

Xl = h n LlFl + h12 LlF2


(6.51 )
X2 = h2l LlFl + h22 LlF2

having displacement coefficients

-k12 h _ kn
h12 = - - 22 - -:;1
Ll (6.52)

Using LlFl and LlF2 from equations (6.50), then equations (6.51) represent an
ellipse. With the assumption of linear behaviour, a slowly rotating additional
force thus causes the journal centre to describe an ellipse around the static
equilibrium point. In reality the path differs from this shape as Fig. 6.19 shows
for the short cylindrical bearing with the additional force Fr = 0.5 Fo. It can
be seen that the linear orbits differ progressively from the true orbits with
increasing eccentricity ratio. The arrows show the directions of the rotating force
for appropriate points on the orbit. We can recognize here again the difference
in direction between force and displacement.
The subject matter of this section is of fundamental importance in rotor-
dynamic analysis. Further influences, which were not considered here, are such
features as turbulence in the oil-film, variable viscosity throughout the oil-film
96 6 Oil-Film Bearings

Cz
Fig. 6.19. Short cylindrical bearing. Journal centre orbits for rotating force
Fr = 0.5 Fo.

and deformation of the bearing shell, that is variation of the clearance space
due to load and temperature differences. These and further influences have been
investigated and published as a result of much numerical and experimental re-
search. Literature on this topic, as well as many tables and diagrams of stiffness
and damping coefficients can be found in references [10] and [40].
7 Rotors with Oil-film Bearings

In Chap. 6 relationships were given, on the basis of certain assumptions, be-


tween radial movement of the journal and the accompanying force. Accordingly,
an oil-film bearing can broadly be described as a spring-damper element, with
the special property that the displacement and the force in general do not occur
in the same direction. In comparison with rotors with rigid bearings, a rotor on
oil-film bearings thus has different natural frequencies, critical speeds and other
resonance behaviour. As a result of the peculiar force relationships, the possi-
bility of instability arises, and a large number of theoretical and experimental
works appear in the literature. In the following, the essential theoretical results
will be given for simple rotor models with two bearings. They correspond in
the main to results obtained experimentally. In this chapter the rotor will be
assumed in general to be horizontal. For a vertical rotor, special considerations
hold, which will be discussed in Chap. 8. The influence of oil-film bearings on
the vibration of more realistic models will be considered in parts II and III.

7.1 Equations of Motion

We consider a symmetrical rotor with two equal bearings. The rotor can be
rigid or flexible and the bearings are rigidly supported. The model of the rigid
rotor is shown in Fig. 7.1. The shaft turns with constant angular velocity [}
and undergoes translational motion only, that is no tilt moti?n occurs. Thus
the movement of the rotor centre W is identical to that of the journal centres
C J Using XI, X2 as the displacements from the equilibrium position eJ, 0:
(Fig. 7.1)1, the equations of motion are

mXI + 2FJ1 (Xl, X2, XI, X2) = FI(t)


(7.1)
mX2 + 2FJ2 (Xl, X2, Xl, X2) = F2(t)

FJ1 , FJ2 are the components of journal force and FI, F2 the exciting force.
Equations (7.1) are non-linear, as discussed in Chap. 6 and are linearised in
the usual way. For unbalance excitation these equations become, using stiffness

1 We denote here the journal eccentricity bye] to differentiate it from the mass eccentricity.
98 7 Rotors with Oil-film Bearings

s
w

Fig. 7.1. Rigid rotor with oil-film bearings

and damping coefficients kik and dik ,


m:i: l + 2 (dllz l + dl2 Z2 + kllz l + kl2Z2) = meil 2 cos ilt (7.2)
m:i: 2 + 2 (d2l Zl + d22 Z2 + k2l Zl + k22 Z 2 ) = meil 2 sin ilt
The Jeffcott rotor will be taken as a model for the elastic rotor using the prop-
erties assumed in Chap. 3 (Fig. 7.2). The equations of the rotor centre W and
of the journal centres CJ are now different. They will be denoted by Y1, Y2 and
Z1, Z2, respectively, and represent displacements from an assumed equilibrium
position. In Fig. 7.2, a static load in the negative Y2-direction is assumed and
so the two coordinate systems have their origins displaced by a distance Ys for
static deflection of the rotor2. Thus the equations of motion are
mih + dYl + k (Yl - zt) = me0 2 cos Ot
(7.3)
mjh + dY2 + k (Y2 - Z2) = meil 2sin ilt
and, for a non-linear journal force,
k (Yl - zd = 2FJl (Z1, Z2, Z17 :i: 2)
(7.4)
k (Y2 - Z2) = 2FJ2 (:1:17 z2, :i: l , :i: 2)

CJ CJ Ii

y[!~~:
W Y1
Fig. 7.2. Jeffcott rotor with oil-film bearings

2In Chap. 3, displacements Yb Y2 relate to the position of the rotor centre., when the rotor
is unloaded.
7.2 Stability 99

or for a linear journal force


k(:l:l - yd + 2 (kn:l: l + k12:1:2) + 2 (dn:l:l + d12:l:2) = 0
(7.5)
k (:1:2 - Y2) + 2 (k2l:l:l + k 22 Z 2 ) + 2 ( d2l :l:l + d22 :l: 2 ) = 0

7.2 Stability

Solution of the homogeneous linear equations requires the eigenvalues of the


system to be found. These are conjugate complex or real and characterise the
natural vibration. The imaginery part corresponds to the natural frequency in
question and the real part gives the stability of the natural vibration. For a
negative real part the vibration decays with time, and for positive real part it
grows. The stability boundary of the system is reached, when the real part of
an eigenvalue is zero. Shafts with oil-film bearings can become unstable due to
the presence of coefficients k12 , k21' called cross coupling stiffnesses. In practice
instability must be avoided, and one must know as much as possible about
the conditions and about behaviour during instability. The subject is rather
complicated. We therefore first investigate the rigid rotor and then the Jeffcott
rotor. The understanding obtained in this way will be useful in the investigation
of large models in Parts II and III of this book.

7.2.1 Rigid Rotor

A shaft rotating at angular velocity n is loaded in its central plane with a


static force 2Fo and e = O. As a rule 2Fo is the dead weight G = mg. It can,
however, represent other forces as for example steam force, gear tooth force or
magnetic force. 2Fo may also be the resultant of a variety of these forces. Hence
the equations

mZl + 2 (dn:l: l + d12 :i: 2 + kll:l:l + k12:1:2) = 0 (7.6)


m[C2 + 2 (d2l :l:l + d22 :l: 2 + k21:l:l + k 22 Z 2 ) = 0
are valid and their solutions :l:l(t) and :l:2(t) are obtained using the substitutions

(7.7)

In order to obtain non-dimensional parameters, one uses the reference frequency

w. = V~
f2Fo or w. = V"8
{9.
wIth 2Fo = G = mg (7.8)
100 7 Rotors with Oil-film Bearings

and non-dimensional time T = w.t. Hence one obtains from equations (7.6) and
after transformation, the equation

[
A2 + ~11 ~ + 111 (7.9)
/3 21 A+ 121

with 1ik, /3ik obtained from equations (6.39) and

(7.10)

The condition for the existence of 'PI! 'P2 is that the determinant of the matrix
of (7.9) is zero. This leads to the characteristic equation

(7.11)

with the coefficients


-2 - -2 - -2
a4 = D ,aa = AaD , a2 = A2 + A4D ,a1 = AID, ao = AoD
Ao = 111122 -112121
Al = 111/322 - 112/321 + 122/311 - 121/312
(7.12)
A2 = /311/322 - /312/321
Aa = /311 + /322
A4 = 111 + 122

The roots of equation (7.11) are non-dimensional eigenvalues "Xr They are in
general complex conjugate.

r = 1,2 (7.13)

where a r = ar/w., the non-dimensional coefficient governing growth of vibra-


tion and wr = wr/w., the non-dimensional natural frequency.
The eigenvalues Ar depend upon D and 1ik, /3ik as can be seen from equa-
tions (7.12). In turn the coefficients 1ik, /3ik depend on e, where e is a function
of S* = Fo/ F". Using equation (6.23),

Fo 2Fo82 C.
F" 1J PDR - D

where
(7.14)

thus
m 82 .5
C. = v'9 1J R P with 2Fo = G = mg .
7.2 Stability 101

Hence the eigenvalues "XT ultimately depend upon the non-dimensional angular
velocity n and the parameter C., which is constant for a given system. C. is
approximately proportional to the power 1.5 of the physical size of the system.
Fig. 7.3 shows against angular velocity the natural frequencies W T and values
of aT for a rigid rotor with C. = 1 and 10.
For small values of n, one natural frequency exists and after a certain
value of n there are two. Correspondingly there are the "growth coefficients"
aI, a~ and a~ in the first n-region and a1 and a2 in the second n-region. For
n > nth, a1 is positive and the system is unstable. The angular velocity at the
borderline nth has a value 2.7 w. and 2.9 W. for C. = 1 and 10, respectively
while the unstable natural vibration has frequencies of 0.5 nth and 0.44 nth,
respectively.
At the stability borderline, R" "X = o. If one puts "X = j w in equation (7.11),
then the following equation is obtained
a 4W' - j a3rif3 - aGP +j a1w + ao = 0

W,
W2 2ws
IX,
IX2 ~ _ _ IX,
0
Q-

- 2ws

- 4ws ~--------------- IX2

Cs=10

W,
W2 2ws
IX,
IX2
0

- 2ws

- 4ws

Fig. 7.3. Rigid rotor with short cylindrical bearings. Natural frequencies and 'growth
coefficients' for C. = 1 and 10.
102 7 Rotors with Oil-film Bearings

for the stability borderline. Since its real and imaginery parts must each equal
zero, the following equations are obtained

(7.15)

and
(7.16)
Solutions of equation (7.16) are w = 0 and w2 = at/a3' The first solution is of
no interest as it occurs at n = 0, from Fig. 7.3. The second solution substituted
into equation (7.15) gives the condition

or using equations (7.12)

Hence the non-dimensional natural frequency at the stability borderline is


- nth AIA2A3
nth = - = (7.17)
w. A~ - AJA3A4 + AoA~ .
From equations (7.12), Ao, ... A4 contain the coefficients "'{ik, {3ik, which are
functions of journal eccentricity ratio e. Hence nth is also a function of e. In
Fig. 7.4 is shown this function for a rotor with short cylindrical bearings. An-
gular velocity is shown as abscissa and eccentricity ratio as ordinate directed
downwards. This is to illustrate the behaviour of the journal position as it would
appear for a shaft under gravity load and as it rises from startup.

Cs

Q5 10
E 0~0.1)
1 Startup curves

0.76

2ws 6ws
Q,Qth-

Fig. 7.4. Rigid rotor with short cylindrical bearings. Borderline of stability and start-
up curves.
7.2 Stability 103

For a system with given parameter G. the journal centre has an eccentricity
ratio of unity for {} = 0 and proceeds towards e = 0 for {} - - t 00. The stability
borderline (}th = f(e) is obtained as the intersection of the startup curve with
the borderline curve. For the parameters G. of Fig. 7.4, {}th lies between 2.7 w.
and 3 W
Using Fig. 7.4 the non-dimensional stability borderline {}th/W. can be shown
as a function of the parameter G. Such a figure is called a stability chart.
Fig. 7.5 shows such a stability chart, which was determined with data from
many different types of bearing.
The required coefficients 'Yik, {3ik and the static load locus function were
calculated for the short cylindrical bearing from [35] and by series solution in
[34]. The values from [39] were obtained from measurements. It should be noted
that in general, for the same bearing shape, alternative output data exist, so
that alternative results are possible for the stability borderlines. Thus, from
stability charts one can only obtain an idea of trends and relationships using
such figures as Fig. 7.5 and generally one cannot expect to get exact values.
With this proviso the following can be concluded from Fig. 7.5:
- With regard to stability the following rank order applies:
Circular bearing, pocket bearing, lemon bore bearing, three-lobe bearing.
From [39], for G. < 0.7, the stability borderline of the lemon bearing is
roughly the same as that of the circular bearing.
- The curves from [39] each have a distinct minimum. Hence increase in G.
can produce a change from stability to instability or vice-versa.
- The mass of the model and the bearing clearance have different exponents
in G. and w . Hence a discussion of the influence of these parameters on

Three-lobed bearing
Q th 100
Ws 50 Lemon bearing
Pocket bearing

10
5
Cylindrical bearing

1351 1341

i ' ,,( i I '''ij Ii "'I Ii., '''1


0.1 10
c
s -
Fig. 7.5. Rigid rotor with oil-film bearings. Stability chart for different types of
bearing, LID = 0.5 [34,35,39] (see bibliography).
104 7 Rotors with Oil-film Bearings

stability is rather complicated. One can thus better judge a given case,
witli reference to an example.

Example 7/1 A model similar to that in Fig. 7.1 is given with mass m =
10,000 kg and bearing data D = 2R = 400 mm, L = 200 mm,,,p = 1.5 %0' ", =
0.030 Ns/m2.
We wish to find the stability onset speed and its change for a 10 % increase
in bearing clearance and for a 10 % increase in oil viscosity. It is first assumed
to be a cylindrical bearing, then to be a lemon-bore bearing.
With the given data and 2Fo = mg,

C. = 1.017 and w. = 180.8 rad/ s.


and from Fig. 7.5, with the curve for the cylindrical bearing [39J

that is nth = 569.5 rad/ s,

or the borderline speed nth = 5438 rpm.


For a 10 % larger bearing clearance

C. = 1.291 and w. = 172.4 rad/ s

and hence nth = 4807 rpm, that is a 12 % lower value.


The viscosity has an influence only on C s For a 10 % greater viscosity we
obtain the borderline speed nth = 5542 rpm, that is an increase of around 1.9 %
above the original value.
For lemon-bore bearings in the original condition nth = 7079 rpm.
For 10 % greater bearing clearance nth = 7688 rpm,an increase of 8.6 %, and
for 10 % greater viscosity nth = 6820 rpm, a decrease of 3.7 % from 7079 rpm.
This example shows how different the behaviour can be. Observed borderline
speed changes occur for the two types of bearing, not only to different degrees
but even with different tendencies (1.15 : -12 %, and +8.6 % and 1.1", : +1.9 %
and -3.7 %).

7.2.2 The Jeffcott Rotor


To calculate the eigenvalues and stability borderline for the Jeffcott rotor of
Fig. 7.2, one assumes as for the stiff rotor, that the shaft rotating with angular
velocity n, is loaded statically with a force 2Fo in its mid plane and that the
centre of mass S of the disc is coincident with W, that is e = O. In addition, it
will be assumed for simplicity that the disc experiences no external damping,
that is d = O. Hence from equations (7.3) the following equations are obtained
mih + k (Yl - zt) = 0
(7.18)
mih + k (Y2 - Z2) = 0 .
7.2 Stability 105

These, together with equations (7.5) determine the displacements:l: ll :1:2 of the
journal centre eJ and Yll Y2 of the disc centre W, describing displacements
from the static equilibrium position due to 2Fo.
In the usual way substitution of expressions (7.7) for :el, :1:2 and correspond-
ing solutions
Y2 = f{)4 e
At At
YI = CP3 e (7.19)

lead to the characteristic equation for the eigenvalues. To obtain non-dimen-


sional parameters, the natural frequency of the rotor with rigid bearings is
introduced as a reference frequency namely,

(7.20)

together with non-dimensional time T = wnt. Further we put

= A = n
A=- n=-
Wn Wn
5 2 Fo
a=- Y.=-- (7.21)
Y. k
5
or a =w 2 _ with 2 Fo= G=mg
n9
Hence we obtain after some manipulation the characteristic equation
=6 =S _
b6 A + bs A + ... + bl A + bo = 0 (7.22)
with the coefficients

b6 = A2 , bs = (Al + a A3) n
b4 = 2A2 + (a 2 + Ao + a A4)!f
(7.23)
b3 = (2AI + a A 3) n ,
b2 = A2 + (2Ao + a ~) n
=2 = =2
, bl = Al n , bo = Ao n

and Ao, ... A4 from equations (7.12).


The characteristic equation is of sixth order and has, in general, the roots

r = 1,2,3. (7.24)

In an analogous way to equations (7.14) we introduce

Fo 2Fo52 en
=
Frj 1J pnR n (7.25)
2Fo52
where en
."PwnR
106 7 Rotors with Oil-film Bearings

Hence the eigenvalues 'Xr are obtained as functions of n and the parameters en
and a.
As an example, the natural frequencies (the imaginary parts of the eigenval-
ues) are presented in Fig. 7.6 together with the "growth coefficients" (the real
parts of the eigenvalues), when plotted against angular velocity of the rotor.
For this purpose it was assumed that the rotor was supported on cylindrical
bearings, with parameters en = 0.3125 and Y. = 0.3 6, that is a= 3.3.
There exist three natural vibrations, of which the second and the third
are stable in the whole speed range considered. In contrast, the first natural
vibration is unstable for nth = 1.2 W n At the stability borderline it has a
frequency Wt = 0.5 nth.
The parameters at the stabi~y borderline can be ascertained in the same
way as for the rigid rotor. With 'X = j w,
one obtains from equation (7.22) the
following equations
- bs WS + b4 ~ - b2 w2 + bo 0 (7.26)
(b 5 ~ - b w + bt ) w
3
2 0 (7.27)
From equation (7.27) the following solutions are obtained

w= 0 and w = 2~5 (b
2
3 V
b5- 4bt b5)
The second of these equations together with equations (7.23), gives

w=2 1 and w-
2
At
At
+ a A3 .
(7.28)

t
Wnl-~====~~~~------------W2
~---W1

Fig. 7.6. Natural frequencies and growth coefficients for a Jeffcott rotor with short
cylindrical bearings. en = 0.3125, y. = 0.3 5.
7.2 Stability 107

The solutions Z5' = 0 and't:f = 1 occur for n = 0 and are therefore of no practical
importance. The solution (7.28) substituted into equation (7.26) gives, after
some manipulation, the following simple expression for the stability borderline

nh - nth _ nth .1 A3 (7.29)


t - Wn - W. YAl + a A3
where nth/ws is obtained from equation (7.17).lt can be seen that the flexibility
of the rotor enters only through the factor a under the root sign in equation
(7.29).
Equation (7.29) gives a stability chart for the Jeffcott rotor, Fig. 7.7, which
corresponds to Fig. 7.4 for the rigid rotor. Both figures relate to short cylindrical
bearing supports. The startup curves are here specified by the parameter en
and the intersection with a borderline curve gives the stability limit. By this
means a stability chart such as Fig. 7.8 is obtained. This shows the stability
borderline for a Jeffcott rotor with different values of Ys/ C and oil-film bearing
data from measurements by Glienicke [39].

0.2 2 20
O~--~--~------T-T--------

0.51
1 Startup curves

0.5 5

~- _ _ _ _:= Borderlines

3wn
QJl th -

Fig. 7.7. leffcott rotor with short cylindrical bearings. Borderlines of stability and
startup curves.

Discussion of the influence of the model parameters on the stability bor-


derline is more complicated than for the rigid rotor because of the additional
parameter y./c, and will not be attempted here. In the same way we will not
pursue further the influence of bearing type. As a rule the same procedure is
valid as for the effect of bearing type on a rigid rotor. One can investigate a
particular case on Figs. 7.7 and 7.8 or by calculation with the given relation-
ships.
Finally it should be mentioned that in practice differences between theo-
retical and practical values of nth will usually be noted, if realistic models are
108 7 Rotors with Oil-film Bearings

5
~=
20

0.5

0.2
li'l ii' i" i I I Iii i

0.1 0.5 5
Cn -
Fig.7.8. Stability chart for a leffcott rotor with cylindrical bearings. Values for
from [39], L / D = 0.5.
lik, {3ik

analysed using these equations (see Parts II and III). For example, the mini-
mum clearance can decrease due to thermal distortion and hence the stability
borderline can be raised. A raised stability borderline can also depend on un-
known damping within the system. Also thrust bearings can be stabilising by
virtue of axial loads. Destabilising effects can arise due to magnetic attraction
in the rotors of electrical machines or due to clearance and sealing effects in
turbomachines (Chapter 10).

7.3 Unbalance Vibrations

The equations of motion for unbalance vibration of rigid and Hexible rotors
are given in Sect. 7.1. Solutions have been achieved by Glienicke [41], Merker
[42] amongst others. In the following the most important results will be given
for the horizontally-supported rotor. The behaviour of a vertical rotor will be
described in Chap. 8.
The unbalance forces cause the rotor axis to deHect from its equilibrium
position and to move in a closed orbit. The extent, shape and position of the
orbit depend upon the rotor speed. The orbit will be executed once every rev-
olution and resonance occurs when the running speed corresponds to a natural
frequency.
For a rigid rotor as in Fig. 7.1 an example will be considered from [42] in
Fig. 7.9. The model is prescribed by the parameter

SM=SO- ,
n (7.30)
WM
7.3 Unbalance Vibrations 109

o~-------------------------

l
E

a 0.5

5 10

~ & c9~ Non-Linear

+ & Linear

t 0.4
~
Linear
,,~
/.-_.-
m
6" Non-Linear
C 0.2

OO-~=-~----~----~-----.~
o
--
5 10 15 20
n
WM
Fig. 7.9. Rigid rotor with short cylindrical bearings. SM = 0.32, e = 0.2 5. From J.
Merker [42].

a) Eccentricity of the journal centre


b) Rotor orbits
c) Mean orbit radius

where So is the Sommerfeld Number from equation (6.34) and the reference
frequency is given by

(7.31)

SM will be assumed to be 0.32 and the mass-eccentricity e to be 0.2 o.


110 7 Rotors with Oil-film Bearings

Fig. 7.9a shows the "lift" ofthejournal centre with increasing angular veloc-
ity. In Fig. 7.9b are shown non-linear and linear journal centre orbits. Fig. 7.9c
shows a plot of mean orbit radius versus angular velocity. Here Tm = ..;;ib,
where a and b are the values of the half axes of the elliptical path, that is the
extreme excursions of the orbit, calculated on the basis of non-linear theory.
For n ~ 13 WM a certain resonance is indicated.
From Fig. 7.3 it can be seen that for certain C. values WI and W2 are
never equal to n, and so no resonance is possible. The linear results differ
only slightly from the non-linear. This is because, with e = 0.2 0 a relatively
small mass eccentricity is assumed. For e > 0.43 0 in this example, the jour-
nal amplitude would be greater than the bearing clearance, which of course is
impossible. Using non-linear calculations, more realistic journal-centre orbits
are produced, within the bearing clearance dimensions. An extreme example is
shown in Fig. 7.10.

x,

Fig. 7.10. Linear and non-linear journal-centre orbits. Rigid rotor with short cylin-
drical bearings. From [42]. SM = 4.826, e = 0.8 5, n = 24.14 WM.

The unbalance vibrations of a leffcott Tatar with oil-film bearings will now
be discussed for the example whose eigenvalues are shown in Fig. 7.6. The
model has short cylindrical bearings and the parameter Cn = 0.3125, while
Ys = 0.3 o. In Fig. 7.11 are shown the linear amplitudes of horizontal and
vertical displacement $1, $2 of the journal centre CJ and iiI, fh of the rotor
centre W, as a proportion of the mass eccentricity e. From Fig. 7.6, resonance
is to be expected for n = W3 = 0.94 Wn and for n = W2 = 0.96 W n
Fig. 7.11 shows a strong second peak for the disc amplitude only, with the
journal amplitude having a rather weak second peak. This is typical, as is shown
in figures from reference [10].
As in the examples using a rigid rotor, linear calculations will also be suf-
ficiently accurate here for mass-eccentricities e < 0.2 o.
For e = 0.2 0 the
7.3 Unbalance Vibrations 111

3 , "
Y2
"
~
,,, I'/e
\
e , I
I
\
,

2L
e
2 , I

I
, I
I

I
I
I

"

o~----~------~------~
o 2 3
Q
Wn -
Fig. 7.11. leffcott rotor with short cylindrical bearings. en = 0.3125, y. = 0.3 6.
Displacement amplitudes from [32].

maximum amplitudes of the journal centre are

Xl max = 0.20 0 and X2max = 0.22 0


and of the disc centre

ihmax = 0.42 0 and ihmax = 0.59 0 .


The oil-film bearings have a rather strong damping effect here as can be recog-
nized from the breadth of the resonance peaks.
The orbits of the disc centre are ellipses for small unbalance and the vibra-
tions in a random radial direction are harmonic with frequency il. For larger
unbalances the orbits are similar to those of Figs. 7.9 and 7.10. The vibrations
now contain components of frequency 2il, 3il etc. as well as il.
In Sect. 7.2 the stability of the statically-loaded rotor without unbalance
was considered. The influence of unbalance on the stability borderline was in-
vestigated by Lund [43] and by Merker [42]. In reference [43] a rigid rotor with
short cylindrical bearings was assumed and the stability was obtained for the
orbit due to unbalance. Average values of the stiffness and damping during a
cycle were used. It was found that the borderline was raised as unbalance was
increased and that this rise could be considerable. This feature could be use-
ful in the interpretation of measurements. It is, however, hardly suitable as a
means of stabilisation and one should endeavour to reduced unbalance as much
as possible. For a flexible rotor this influence is weaker. The frequency of the
unstable vibration is roughly the same as that without unbalance and is around
il/2.
112 7 Rotors with Oil-film Bearings

In reference [43] it was further ascertained that parametric excited vibra-


tions take place as a result of oscillating bearing forces. They lead to instability
for np < nth with a frequency of exactly n/2. The vibrations are, however,
weaker than for the other instability. The investigations in reference [42] lead
to similar conclusions, where the stability of unbalance vibrations was found
numerically as they built up after a sudden application of unbalance.
Two decades earlier both Hon [44] and Someya [33] came to roughly the
same conclusions. They were also deduced from experiments by Hon and oth-
ers. The essentials are brought together in Fig. 7.12. It shows versus angular
velocity n the amplitudes (of a point on the rotor in a certain direction) :Z:u
as a result of unbalance and :z:w as a result of bearing instability in a and b,
respectively. In c the corresponding frequencies are shown. For n = np , insta-
bility commences as a result of parametric excitation and at n = nth powerful
bearing instability starts as a result of the cross-coupling effect. The unbalance
vibration has frequency Wu = n (and for greater unbalance also 2n, 3n, ... )
and the vibration as a result of instability has frequency Ww = n /2. After nth
it has frequency Wr with r as the order number associated with the natural
frequency in question. This behaviour refers to leffcott rotors with cylindrical

t
Xu
a

Q--

t
b Xw

Q--

t Wu=Q

c w

Q--

Fig. 7.12. Amplitudes and frequencies for unbalance vibration and bearing
instability.
7.4 Summary 113

bearings. For other rotors and bearing types a somewhat different behaviour is
to be expected.

7.4 Summary

In Sects. 7.1 to 7.3 the theory of vibrations of a symmetric rigid or flexible


rotor with oil-film bearings was considered and the performance of the oil-film
bearings was determined by hydrodynamic bearing theory. This applies for cer-
tain assumptions, as described in Sect. 6.1, applied to the Reynolds Equation
(6.12), whose integration gives the pressure distribution in the lubricant film.
This is then integrated to give the resultant oil-film force acting on the jour-
nal. This depends, among other conditions, on the position of the journal in
the clearance circle and on the magnitude and direction of the journal centre
velocity. With the known journal force and shaft parameters, the equations of
motion can be obtained for the journal centre and the rotor centre. From their
solution the motions, that is the vibration of these points, can be ascertained.
Although some system features are not considered in this determination of the
journal force (see end of Chap. 6), the results agree quite well with experimental
measurements. To obtain a better overview some cases will now be discussed, in
which the results of Sects. 7.2 and 7.3, as well as of further publications, based
on computed and experimental investigations will be summarized. The cases
relate to the models of Figs. 7.1 and 7.2. They can also relate to the behaviour
of similar models if one exercises a certain degree of care.

1. The oil-film of a journal bearing acts like a spring and damper in parallel
combination.

2. The spring and damping properties depend upon the radial direction: the
bearing is non-isotropic.

3. The spring and damping force vectors each have in general a different di-
rection from the displacement vector and the velocity vector of the journal.

4. The spring and damping properties depend upon the bearing type and
upon the assembly, for example the load angle. For a certain bearing
type the Sommerfeld Number, equations (6.34, 6.35), and the length to
diameter ratio are significant. The properties thus depend among the other
things upon angular velocity {J of the shaft.

5. Computed and experimentally-obtained spring and damping coefficients


for many bearing types are summarized in ref. [10j. In practice differences
from these values can arise due to many different influences.
114 7 Rotors with Oil-film Bearings

6. A rigid rotor as in Fig. 7.1 has one or two natural frequencies (Fig. 7.3).
A flexible rotor as in Fig. 7.2 has, as a rule, three natural frequencies
(Fig. 7.6).

7. As a result of parametric excitation, a weak instability occurs at an angu-


lar velocity np , for cylindrical bearings. Stronger instabilities result at the
borderline angular velocity nth> np. The value of nth can be ascertained
from stability charts such as the ones shown in Figs. 7.5 and 7.8.

8. Flexible rotors are less influenced by oil-film bearing properties than are
rigid rotors. The ratio of static displacement y. to bearing clearance 5 is
significant in this respect.

9. For cylindrical bearings any unstable vibration has a frequency n /2 in the


region np < n < nth. After nth it has a frequency equal to the natural
frequency in question.

10. Practical measures taken to enhance stability above nth are, amongst
others, changes in bearing clearance, in oil viscosity (different temperature
or oil type), in length to diameter ratio, in orientation of the bearing axes,
in oil supply or in bearing type. The kind of change and its efficacy can
be concluded from the stability chart (as shown in example 7.1).

11. The stability borderline can be raised by unbalance.

12. For unbalance arising from mass eccentricity less than one fifth of the
bearing clearance (e < 0.2 5) linear assumptions give good results. El-
liptical orbits are produced with harmonic variations in radial directions
of frequency n. For large unbalances the ellipses are deformed (see, for
example, Fig. 7.10) and radial vibrations contain higher harmonics with
frequencies 2n, 3n etc.
8 Vertical Rotors

In Chap. 7 the vibration behaviour of a horizontally supported rotor with oil-


film bearings was described. This horizontal support gives rise to journal eccen-
tricity due to rotor weight except for the case of further forces producing a zero
resultant.
For a vertical rotor the weight acts in the axial direction, and it gives rise to
no journal eccentricity if no other forces act in a predominantly radial direction.
Hence in contrast to horizontal rotors, special dynamic features arise, which will
be considered in this chapter.
Cylindrical bearings and similar types develop little or no dominating force
as the journal takes up a central position, and so the journal describes more
or less large orbits about the bearing centre. Tilt-pad bearings are thus better
suited to vertical rotors and their properties will be described in the following
section.

8.1 Tilt-pad Bearings

A great number of publications on computed and experimental investigations


have been produced, relating to the static and dynamic properties of tilt-pad
bearings. To give a comprehensive coverage here would not be appropriate.
Instead the essential behaviour of tilt-pad bearings for vertical machines will
be described, using the results of Merker (45).
Consider a bearing as shown in Fig. 8.1. It has equal circular arc segments,
which are arranged around the circumference at equal intervals. These segments
are capable of tilting on their pivots. The geometry of a segment is shown in
Fig. 8.2.
The journal radius is r, the radius of the contact circle is R and the segment
radius is R . These give the following parameters

8 =R-r bearing clearance


8. = R.-r segment clearance (8.1)
8.
v clearance ratio.
8
The clearance ratio v is especially important here. As a rule one chooses R. >R
and so v> 1.
116 8 Vertical Rotors

Fig. 8.1. Symmetrical tilt-pad bearing

Fig. 8.2. Geometry of a segment

In ref. [45] the journal force was found as a function of the position (ej, ,)
and of the velocity (eJ, i) of the journal by solving the Reynolds Equation
(6.12). The special feature, that the segments are movable, was accomodated
in the following way by Ott [46].
The journal force is exactly opposed by the resultant of the segment forces
acting on the journal, as a result of the oil-film pressure. We consider first one
segment only. It was assumed in ref. [45] that the moment of inertia, the friction
force due to tilting and the shear force on the segment due to the lubricant are
all negligible. From this it follows that the vector Fs of the segment force must
pass through both the pivot point CT and the centre of curvature C s of the
segment; this is true only for a certain angle", (Fig. 8.2). This angle depends
upon the eccentricity es and is determined by iterative calculation.
8.1 Tilt-pad Bearings 117

If the journal centre OJ is moved in a circular arc B, whose centre is OT


(Fig. 8.3) and simultaneously the segment is turned in such a way that the
angle It remains the same, then the magnitude of Fs stays the same, while the
direction of F s is assumed to move through the angle U'.
Because Ii < r the same can be said to a good approximation for a dis-
placement along the tangent T. Any possible displacement is limited by the
bearing clearance and so is small in relation to the journal radius. It finally
follows that, for the possible positions of the journal centre OJ on the straight
line T of Fig. 8.3, the value of Fs depends only upon es and es. Hence we can
say that
IPs I = f(es, es) . (8.2)
To determine the journal force FJ we consider Fig. 8.4. The journal centre has
the coordinates (eJ, '1) and velocities (eJ, i). The segment defined by angle fJ
produces a force Fs at angle fJ'. Because Ii < r we can assume fJ to represent
the direction of the force. The required parameters to determine IPs I are given
in Fig. 8.4
es = eJ cos('Y-fJ)
(8.3)
es = eJ cos ('1 - fJ) - i eJ sin ('1 - fJ)
With these relationships we can find the force Fs for each segment ofthe bearing
and the journal force can be calculated by vector addition.
The advantage of this method is that the relation (8.2) needs only be de-
termined for one segment. With unequal segments, however, one must write
equation (8.2) for the different segments. The advantage, however, remains that
one does not need to calculate for the whole bearing. In the following are given
some results obtained using this method. The same bearing will always be used,
having the data given in Table 8.1.

Fig. 8.3. Explanation of the behaviour of the segment force Fs


118 8 Vertical Rotors

Fig. 8.4. Determination of journal force

Table 8.1. Tilt-pad bearing from Fig. 8.1

Data for Figs. 8.5 to 8.11 and 8.19:

Twelve segments
Segment angle 25
Pivot in centre of segment
Bearing length L = 0.2 D

As usual the results will be given in terms of non-dimensional parameters.


For the segment these are

es . es Fs 1/;2
es= 6 ' es = 5n ' Ss = D L 1/ n (8.4)

and for the journal,

(8.5)

Comparison with equation (6.34) shows that Ss and S] are Sommerfeld Num-
bers.
Fig. 8.5 shows the segment force as a function of the eccentricity eS for
different clearance ratios. The abscissa extends from -1 to +1. For es = +1
the journal touches the segment considered and for es = -1 it touches the
8.1 Tilt-pad Bearings 119

10 1

Ss 10

10-1

10-2

10-3

10- 4
I i i i i i i
I i i
I i i
I
-1 -0.5 0 0.5
ES-
Fig. 8.5. Segment force as a function of the eccentricity ratio, from reference [45].
Bearing data are from Table 8.1.

opposite segment. It can be seen that also for negative eccentricity ratios a
segment force exists if v > 1. For v = 1 the segment force approaches zero
as the eccentricity ratio approaches zero. This is understandable, as for this
condition Rs = R and hence a constant film thickness results, equal to the
bearing clearance 8 and no hydrodynamic pressure is produced. Elsewhere, Ss
increases with v. For larger values of es and v, however, Ss changes its trend.
The figure clearly shows the non-linear character of the segment force.
The influence of rate of change of journal displacement is in the radial
direction on the segment force is shown in Fig. 8.6. For positive is the jour-
nal moves towards the considered segment and the eccentricity ratio increases.
For negative is this is reversed. The static segment force (is = 0) is increased
for positive is and reduced for negative is. While for increasing is the force
increases, it becomes zero at a certain negative value of is. The dynamic com-
ponent of Ss depends upon is in a non-linear fashion.
Fig. 8.7 shows the segment forces for a bearing with clearance ratio v = 2
both for a central journal and for an eccentricity e] = 0.3 5 in the direction
of the pivot of segment 12. The length of each arrow represents the magnitude
of the appropriate force. For a central journal position the segment forces are
greater than zero because v > 1. They all have equal value, their resultant is
zero, and so the journal force is zero. For an eccentric position, the segment
forces remain unchanged, for those segments which lie at right angles to the
eccentricity (here numbers 3 and 9). The remaining segment forces are either
greater (here 12, 11, 1, 10, 2) or smaller (8, 4, 7, 5, 6).
120 8 Vertical Rotors

I i I I ' , I i , I
-1 -0.5 o 0.5
ES-
Fig. 8.6. Segment force for radial movement of the journal, from reference [45]. Bea-
ring data from Table 8.1, clearance ratio v = 2.

5 3

Fig. 8.7. Segment forces and journal force for e J = 0 and e J = 0.3. Bearing data
from Table 8.1 and v = 2.

For the eccentricity assumed in the direction of a pivot, the journal force has
the same direction as the eccentricity. This is also true when the eccentricity is
directed midway between two segments. In all other directions the force has a
somewhat different direction from that of the eccentricity. For twelve segments,
however, this is only noticeable at high eccentricity ratios.
Fig. 8.8 shows for our example the journal force as a function of eccentricity
ratio in the direction of the pivot of one of the segments. A range of clearance
ratios is covered.
8.1 Tilt-pad Bearings 121

10 2
f
SJ 10 1

10

10-1

10-2

10- 3
0 0.5
E-
J

Fig. 8.8. Journal force for different clearance ratios, from reference [45). Bearing data
from Table 8.1.

For low values of clearance ratio, the journal force becomes greater with
increasing clearance ratio. For very large clearance ratio the trend changes with
eccentricity ratio.
The change in journal force for small displacement from a certain initial
position and for small velocity can be determined using stiffness and damping
coefficients (see Sects. 6.1 and 6.2). As already mentioned, for displacements in
directions I or I' of Fig. 8.9, the journal force has the same direction as the
displacement. This can also be assumed for displacement in directions I I an~ I I'
perpendicular to I and I' respectively. Hence for each of these directions there is
only one stiffness and one damping coefficient. We denote these coefficients for
radial and tangential directions by kr, k t and d., dt respectively. The coefficients

2
I'

n'

Fig. 8.9. Coordinate system for stiffness and


damping coefficients.
122 8 Vertical Rotors

for other directions differ only slightly from these, if the bearing has many
segments and if the eccentricity ratio is not too large.
For system 1,2, which has the angle {J relative to the direction of the ec-
centricity, the stiffness coefficients defined in equations (6.43) can be written as
follows
kr cos 2 {J + kt sin2 {J (kr - kt ) sin {J cos {J"
(8.6)
k12 kr sin 2 {J + kt cos 2 {J

The relationships (8.6) also hold for the coefficients dl1 , d12 and d22 as functions
of dr and dt. It can be seen that, by coordinate transformation, cross-coupling
terms emerge. They cannot, however, be the cause of instability as they are equal
in value. In reference [45] the coefficients were presented in a dimensionless form
using the force
F* = D L 11 0 (8.7)
1/;2 '
the bearing clearance 5 and the angular velocity O. The coefficients may be
written in terms of these parameters as follows

kt5
, 'Yt = F*
(8.8)
a_ dt50
, ,..,t - F*

These are presented in Figs. 8.10 and 8.11 for our bearing with different clear-
ance ratios. They are valid for directions I, II. For a central journal position

10 2
t
lr 10'
It
10

10-'

10-2

10"3
r-r-..---.---r'""\j--"---'---''-.-,j Fig. 8.10. Stiffness coefficients for
0 0.5 1 tilt-pad bearing with data from
EJ - Table 8.1 (from reference [45]).
8.2 Vibrations 123

10 2

[3r 10 1
[3t
10

10- 1

102
150
10- 3
I
0 0.5 1
e:J -
Fig. 8.11. Damping coefficients for tilt-pad bearing with data from Table 8.1 (from
reference [45]).

the coefficients have minimum values; also IT = It and (3T = (3t. For 1/Jv = 1
and e] = 0 then IT) It = 0 and thus no stiffness exists. All coefficients become
greater with increase in eccentricity ratio. It should be particularly noted that
the damping coefficients (3T) (3t decrease with increase in v.

8.2 Vibrations

In order to describe the vibrations of a vertical rotor we shall use the model
shown in Fig. 8.12. The rotor can be rigid or flexible and the influence of the
axial bearing will be neglected.

Fig. 8.12. Model of a vertical rotor


124 8 Vertical Rotors

First consider the rigid rotor with short cylindrical bearings. This model
was investigated by Merker [42] and by Lund and Nielsen [43]. For a central
position (e = 0) and no radial force the rotor runs in the centre of the bearing
and for all values of n, e] = O. In this position the system has both a decaying
and an increasing amplitude natural vibration, each having the same frequency.
The natural frequency W1,2 and the growth coefficients aI, a2 depend upon
n. They can be made non-dimensional by using the reference frequency from
equation (7.31)

WM = 1J R (~) 3
4m 5
They can also be presented in terms of WM as in Fig. 8.13. This figure shows
that the coefficient a1 is positive for n > 0 and that the condition is always
unstable.
For mass eccentricity e > 0, that is for an unbalance condition, the rotor
centre moves in a circular orbit. For small changes in this orbit, one can assume
a linear system and so calculate eigenvalues and stability boundaries. Fig. 8.14
shows regions for stable and unstable circular orbits as functions of the angular
velocity n. For n < 29 WM a circular orbit is stable if of sufficiently large radius.
For n > 29 WM the region of stable operation is reduced by a second unstable
region.
The size of the orbit radius r e for unbalance excitation depends on the mass-
eccentricity e. The orbit will be described with velocity Ten, from which the
journal force can be found for a given radius, in both magnitude and direction.
The journal force must be in equilibrium with the unbalance force and from
this the corresponding radius will follow. The calculation shows that for certain
parameters, three circular orbits of different radii exist instead of one. At least

Wl,2
(Xl 20W
M
(X2 W1,2
(X1

0
50w M
Q-
-20w M (X2

Fig. 8.13. Rigid rotor with short cylindrical bearings. Natural frequencies and growth
coefficients for central position of the journal (from reference [42]).
8.2 Vibrations 125

0.515

o~~~~~~~~~~~~
a
Q-
Fig.8.14. Stability chart for circular-centred orbits (radius Tc). Rigid rotor with
short cylindrical bearings (from reference [42)).

one of these orbits will be unstable. Figs. 8.15 to 8.18 show examples for different
values of e/5.
In Fig. 8.15 a mass-eccentricity of 40 % ofthe bearing clearance is assumed.
From {} = 0 to {}I the circular orbit has radius TI. Fig. 8.15 indicates that from
{}I to {}2 instability occurs. For higher speeds a stable circular orbit is again
possible, with radius T3.
For e = 0.44 5 (Fig. 8.16) a stable orbit exists for every angular velocity. It
has initially a radius TI and later T3. From {}/ to {}Il a second orbit of radius

6.-----------------------
f

0.515

O~~~L+~~~~~-L~~

a
Q-
Fig.8.1S. Orbit radii for unbalance excitation, e = 0.4 6. Vertical rotor with short
cylindrical bearings [42].
126 8 Vertical Rotors

6.-------------------------
Stable

0.56

O~~~~~~~~~~~~
a 100wM
Q-

Fig. 8.16. Orbit radii for unbalance excitation, e = 0.44 5. Vertical rotor with short
cylindrical bearings [42].

1'2is possible. Similar results hold for e = 0.6 6 (Fig. 8.17). Here the orbit is
always stable at radius 1'1 and always unstable at radius 1'3' In a confined area
around 150 WM there are two stable orbits.
For e = 1.2 6 (Fig. 8.18) there is only one circular orbit. It is stable for all
angular velocities and 1'1 - 6 for {} _ 00.
If one employs tilt-pad bearings instead of cylindrical bearings then the
orbit is always stable. For unbalance excitation of an unloaded rotor the system

6.-----------------------

0.56

Unstable

O~~~~LL~~~LL~~-L

a
Q-
Fig. 8.17. Orbit radii for unbalance excitation, e = 0.6 5. Vertical rotor with short
cylindrical bearings [42].
8.2 Vibrations 127

6~-----------------------

t --=:::::;:;:;;::nTIT ~

0.56
Unstable

o~~~~~~~~~~~

o
Q-
Fig. 8.18. Orbit radii for unbalance excitation, e = 1.2 5. Vertical rotor with short
cylindrical bearings [42].

behaves like that with cylindrical bearings. As a rule a circular orbit occurs
for a certain group of parameters. There are, however, theoretical regions with
three orbits. Calculations [42] indicated that with 1'1 < 1'2 < 1'3 an orbit with
either 1'1 or 1'3 could occur. The orbit with 1'2 was unstable.
Fig. 8.19 shows the non-linear orbit radii against angular velocity {} for a
rigid rotor with tilt-pad bearings from Table 8.1 and with clearance ratio v = 5,

e
6.------------------------- G
f 4

0.56

0.1
O~T-~~~._~~~~~~
o WO
Q-
Fig. 8.19. Orbit radii for unbalance excitation. Vertical rotor with tilt-pad bearings
from Table 8.1 with v = 5.
128 8 Vertical Rotors

in terms of
(8.9)

For a mass-eccentricity e = 0.1 5 a broad resonance can be seen. For e = 0.2 5


and n > 1.19 Wo, there exist three theoretical radii, the central one being
unstable.
Up to now the rotor has been assumed rigid. For a flexible rotor one sees
the same fundamental phenomena. However, these bearing-induced phenomena
subside with increasing flexibility of the rotor, while the rotor resonance takes
on more significance. Further details can be found on this subject in reference
[42].
9 Rolling-element Bearings

Rolling-element bearings consist of rolling bodies which move in races. They


can be loaded radially or axially or in both of these directions. The basic forms
are shown in Fig. 9.1. The rolling elements can be balls, rollers, tapered rollers,
spherical rollers or needles (Fig. 9.2). These rolling bodies are positioned within
a cage, which holds them at a set spacing. Futher forms of construction for
ball-bearings are shown in Fig. 9.3.
Loading of the bearing leads to movement between the races. Their mathe-
matical relationship must be known in order to be able to determine the influ-

a b c
Fig. 9.1. Basic forms of rolling-element bearing

a) Radial ball-bearing
b) Roller-bearing
c) Axial ball-bearing

e B- B B ~ Fig. 9.2. Rolling elements

a b c d
Fig. 9.3. Further forms of ball-bearing

a) Standard ball-bearing
b,c) Angular-contact ball-bearing
d) Self-aligning ball-bearing
130 9 Rolling-element Bearings

ence of rolling-element bearings on the dynamics of a rotor. The force-deflection


behaviour of such bearings is evidently complicated. Essential parameters are
the geometry and the material of the bearing, as well as the lubricant film in
certain circumstances. In the following sections some fundamentals will be con-
sidered, by which the influence of rolling-element bearings on the vibration of
machine rotors can at least be estimated.

9.1 Deformation of the Rolling-element

In order to derive the force-deflection relationship of a rolling-element bearing


we consider first a single rolling body. In the case of the radial bearing we
ascertain the deflection of the inner race due to a radial force F (Fig. 9.4) .

Fig. 9.4. Radial ball-bearing with single ball.

This deflection arises from deformation at the contact points of the balls and
the race. We neglect the latter contribution, so that we are left only with the
problem of ball deformation. This was solved by Hertz [47] with the following
assumptions:
Linear elastic material
- Neglect of the shear stress at the contact surface
- Small deflection in relation to the dimensions of the element .
For the simple case of ball contact with a convex or concave surface, as a
boundary of a semi-infinite continuum (Fig. 9.5), the Hertzian theory gives for
loading by a normal force F, the deflection

(9.1)

where E is the elastic modulus


v is the Poisson's ratio
and R is given in Fig. 9.5.
9.1 Deformation of the Rolling-element 131

Fig. 9.5. Models describing parameters


in equation (9.1)

If on the right-hand side of equation (9.1) all the parameters except F are
c
collected in the function K , then one can write

(9.2)

For a radial ball-bearing the contact areas are somewhat more complicated.
Contact exists between the surfaces of the ball and the running sudaces of the
inner and outer races. The contact sudaces have different curvatures, the major
curvatures being .l..
r2
and .l..
T4
or .l..
T3
and .l..
TS
(Fig. 9.4). With rl as the ball radius
the equation for the contact sudace corresponding to equation (9.2) is

5 = cR{E, v, rl . .. r5) F~ . (9.3)

c
Details for the determination of R can be found in reference (48].
The model in Fig. 9.4 has two contact areas. From equation (9.3) the radial
deflection of the inner race is

(9.4)
where c R,; and cR,a correspond to the parameters of the inner and outer races,
respectively. By the use of equation (9.4) on standard ball bearings we can
obtain a simpler relationship as a result of such curvature ratios. With E =
20.6.10 4 N/mm 2 and v = 0.3, then for all bearing sizes
F 2/ 3
5 = 4.37 . 10- 4 dl / 3 mm (9.5)

where d is the ball diameter.


In a similar way, for a roller-bearing we have
FO.9
5 = 0.77 . 10- 4 10.8 mm (9.6)

where I is the effective length of the rollers.


132 9 Rolling-element Bearings

In equations (9.5) and (9.6), F is in Newtons and d and 1 are in mm. This
requirement can however be avoided by the use of non-dimensional parameters.
To avoid giving yet more relationships, such parameters will not be used here.
It can be observed that with roller bearings the force F has an almost linear
relationship with the deflection 8. Also the significant length parameter is the
effective roller length; the roller diameter does not appear in the equation.

9.2 Stiffness of Rolling-element Bearings

From a knowledge of the force-deflection relationship for a rolling body one can
now obtain the same relationship for the complete bearing with z rolling bodies.
From equations (9.5) and (9.6), the contact force Fi for the ith rolling body to
give a radial displacement 8i is

Fi= [8
~']n (9.7)

with
3
n=-
2 } (0' ball- be",;ng,
C = 4.37 . 10- 4 d- I / 3 mm N- 2 / 3
(9.8)
1
n= -
0.9 } (0' wlle,-bearing'.
C = 0.77 . 10- 4 1-0 .8 mm N-o.9
To obtain the displacement 8i it will be assumed that the outer race is held
rigidly and that radial play I exists between the races. We also assume that the
rigid inner race is displaced by the amounts x and y (see Fig. 9.6). Then

8i = X cos <Pi +Y sin <Pi - I (9.9)

where <Pi = i <PI and <PI = 2 7r / z.

4.,<:-
I

x Fig. 9.6 . Coordinates and angles for a rolling- element bearing.


9.2 Stiffness of Rolling-element Bearings 133

A contact force only arises when 6i is positive. For a bearing with no play
b > 0, y = 0,
= 0) and with a displacement in the :z: direction only, that is :z:
then we substitute into equation (9.7)

7r 7r

6i = {
:z: cos C{)i with "2 < C{)i <"2 (9.10)
o with
7r
"2 < C{)i < 3"2
7r

The components of Fi in the :z: and y directions are

Fxi = Fi cos C{)i Fyi = Fi sin C{)i (9.11)

For the assumed displacements :z: > 0 and y = 0 the components Fyi of the
rolling elements lying opposite are raised, that is those having the angles +C{)i
and -C{)i, and the sum of the components Fxi is the bearing force

(9.12)

where 0 < C{)i < 7r /2.


The calculation of the expression in brackets for practical rolling element
bearings shows that, to a good approximation, it is proportional to the number
of rolling elements, z. Thus

where S = S(n).
Hence the bearing force is

(9.13)

and the displacement is

(9.14)

The stiffness is found by differentiating equation (9.13) with respect to displace-


ment. Thus
1 n-l
dF n n 1 n ---
k = - = - Sz:z: - = - (Sz)n F n (9.15)
d:z: On 0
The stiffness thus depends upon the loading and is zero when the bearing is
unloaded. From equations (9.14) and (9.15),
134 9 Rolling-element Bearings

t
x

F --- Fig. 9.1. Calculation of stiffness

In particular, in Fig. 9.7


(9.16)

The constant S takes the value 0.23 for ball-bearings and 0.24 for roller-
bearings. Hence with n and C given by equations (9.8), the relationships (9.14)
and (9.15) are obtained for ball-bearings

F)2/3
:v = 1.2 d- 1 / 3 ( --; I'm
(9.17)
k = 1.3 Z2/3 d1 / 3 Fl/3 N / I'm

x 100
11 m

50

10 20 kN
F-
Fig. 9.8. Displacement z as a result of applying a load F for a rolling-element bearing
with no play in which z = 10 and d = I = 10 mm.
9.3 Lateral Stiffness, Critical Speeds 135

while for roller-bearings

x = 0.28 1- 0 .8 ( ~) 0.9 pm
(9.18)
k = 4.0 ZO.9 10 .8 F O. l N / pm

in which F is measured in Newtons and d and 1 in mm.


Figs. 9.8 and 9.9 show respectively the displacement x and the stiffness k
against load for both the ball-bearing and the roller-bearing, in which z = 10
and d and 1 are each 10 mm. As already mentioned k = 0 for F = O. Fig. 9.9
shows that for small loads the stiffness depends heavily on load.
Equations (9.17) and (9.18) relate to a bearing with no play. For a bearing
with play the displacement will be greater and the stiffness smaller.

k N
11m
400

200

O~---,---.----.----.----.--
o 10 20 kN
F-
Fig. 9.9. Stiffness k as a function of load F for a rolling-element bearing with no
play in which z = 10 and d = I = 10 mm.

9.3 Lateral Stiffness, Critical Speeds

In equation (9.15) k is defined as dF/dx. We now put Fx for F and write

k _ dFx
xx - dx (9.19)

We define kxx as the in-line stiffness. Correspondingly for a load Fy in the y


direction,
k _ dFy
YY - dy (9.20)

where kyy is defined as the lateral stiffness.


136 9 Rolling-element Bearings

The lateral stiffness, like the in-line stiffness depends upon whether or not
the load vector passes through the centre of a ball. More detailed discussion on
this follows in Sect. 9.4. We assume a configuration as shown in Fig. 9.6 and
determine the resultant force L1Fy for displacement z > 0 and L1y = eZ, where
e <: 1. The lateral stiffness is thus, with sufficient accuracy, given by

k _ L1Fy
l1li - L1y (9.21)

It is found that for e <: 1, L1Fy is of the form,

(9.22)

where T can be expressed as T(z) and hence from equation (9.21) the lateral.
stiffness is
k =L1Fy = ~ T zn-l (9.23)
yy eZ Cn
With the in-line stiffness from equation (9.15)

kxx n S
= Cn Z Z
n-l

the relationship between the stiffnesses follows,

kyy = T(z)
(9.24)
kxx Sz
It may be observed that the displacement z and hence the load Fx have no
influence on this relationship. A few values are given in Table 9.1. It can be
seen that the lateral stiffness is smaller than the in-line stiffness and that the
ratio increases with the number of rolling elements.
Consider now a rotor supported on rolling-element bearings. With regard
to rotor dynamics it is an anisotropically-supported rotor which, for a simple
case, was considered in Sect. 3.3. Before embarking on an extensive investigation
one should assess whether it is to be expected that rolling-element bearings are
likely to have a significant influence. The important quantity here is the critical

Table 9.1. Ratios oflatera.l to in-line stiffness.

z ball-bearing roller-bearing
8 0.46 0.49
12 0.64 0.66
16 0.73 0.74
9.3 Lateral Stiffness, Critical Speeds 137

speed, which is lower for rolling-element bearings than for rigid bearings. With
the critical speed
no with rigid bearings
and nc with rolling-element bearings
and the stiffness
ko for a rigidly-supported rotor
and k for rolling-element bearing
then from equation (3.24) the ratio of critical speeds is

(9.25)

If a reduction of critical speed of less than 3 % is regarded as of little conse-


quence, then the effect of rolling-element bearings in equation (9.25) can be
neglected, if their stiffness k is greater than about eight times the rotor stiffness
ko
For a rigid rotor, the stiffness (or flexibility) of the rolling-element bearings
plays its full part in determining the critical speed, if the bearing housings are
assumed to be rigid. The critical speeds are

n
y
=..!...
271"
j2kyy
m
(9.26)

and their ratio is


(9.27)

This ratio depends on the number of rolling elements, as is shown in Table 9.2.
For comparison the ratios of critical speeds of a flexible rotor with kxx = 4 ko
are also given. For this latter case the effect of the number of rolling elements
is thus likely to be less.

Table 9.2. Ratios of critical speeds of a rotor with ball-bearings

Rigid rotor Flexible rotor with kxx = 4 ko


z ny/nx nx/no ny/no ny/nx
8 0.68 0.94 0.89 0.94
12 0.80 0.94 0.91 0.97
16 0.85 0.94 0.92 0.98
138 9 Rolling-element Bearings

9.4 Consequences of Element Rotation

Until now it has been assumed that the rolling elements do not move. In fact
they do rotate and so their configuration changes the load vector and the inner
race moves a little. Meldau [49] was the first to investigate this phenomenon.
The consequences on the dynamics of a rotor with rolling-element bearings were
examined by Tamura [50]. In the following the essential results will be given.
Consider first movement of the rolling-elements. The inner race has angular
velocity n, whilst the outer race is at rest. For no slipping of the rolling-element
the contact point at radius T2 (Fig. 9.4) has velocity nT2 while at the contact
point at radius T3 the velocity is zero. The centre of the rolling-element thus
has velocity nT2/2 and it turns about the axis of the rolling-element bearing
with angular velocity
(9.28)

The cage acts in such a way that all rolling elements have the same motion.
The inner race thus rotates with angular velocity n and the set of rolling-
elements with angular velocity nR. Now consider a rolling-element bearing
which is loaded by a constant force F in the x-direction. As a result of ro-
tation of the set of rolling-elements the load vector passes through an element
centre or between two elements. In the first case and also if the load vector
passes exactly half-way between two elements, the displacement of the inner
race will be in exactly the same direction as the load, due to symmetry. For
all other positions, the result is a degree of sideways movement besides the dis-
placement in the load direction. This movement can be found as follows. Assume
a displacement in the x-direction and obtain with the equations of Sect. 9.2 a
resultant force, which in general has a component F y , and thus is inclined to
the displacement x vector by a certain angle. It thus follows that an effective
force F in the x-direction gives rise to displacements
Fx I Fy
x
I
=-;1; y = -- x (9.29)
F F
where Fx is defined as the x-component of F. The sideways movement is pe-
riodic in the positive and negative y-directions. In the same way there is a
periodic movement in the x-direction because the stiffness varies somewhat in
this direction. The inner race thus executes a closed path, which is traced out
once, when the inner ring has turned through the angle CPl = 211"/ z, that is in
the time

211"
Tl = -nCPl = -zn (9.30)
R R
Fig. 9.10 shows as an example the path for a ball-bearing with eight balls and
some bearing play. In this figure
9.4 Consequences of Element Rotation 139

6 = x - I ,the average displacement in the x-direction and

L1x, L1y are additional displacements in x- and y-directions relative to the point
(6,0).

-0,02 /
~
-0.05 0.05 b.y

';10' \V
0.02 \ \ '

with z = 8 and radial play ,=


Fig. 9.10. Orbit of inner race as a result of turning of the set of balls. Ball bearing
5.

The additional displacements L1x and L1y amount to only a few percent of the
average displacement 6. In this example the additional displacement in the y-
direction is 4.2 times greater than that in the x-direction. In general the orbit
becomes greater with increase in bearing play and smaller with increase in the
number of rolling elements.
From the described behaviour it follows that the rolling-element bearing
under constant load possesses periodically varying stiffness in the load direction
and perpendicular to the load direction. This leads to parametric excitation for
rotor-bearing systems, with resonances occurring at

and z [J~ = Wy
and from equation (9.28), to further critical angular velocities

[J' = TZ + T3 Wx and [J" = TZ + T3 Wy


(9.31 )
TZ Z TZ Z

Further details are published in reference [50]. These authors found, for an
example of a rigid rotor, that vibrations in the regions of resonance [J' and [JII
are unstable and that sub- and superharmonic as well as chaotic vibrations arise.
Whilst one is usually concerned with resonances at nx and ny, the phenomena
discussed here can also give rise to acoustic problems.
140 9 Rolling-element Bearings

9.5 Damping of Rolling-element Bearings

Rolling-element bearings have some damping as well as stiffness. This damping


arises on the one hand from the lubricant, on the other from micro-movement
and hence friction at the points of contact. The theoretical and experimental
evaluation of this damping is very difficult. Any test facility to measure this
damping will usually have its own damping, which will swamp the damping to
be measured. Hence they will be very difficult to separate. Basically, although
it would be desirable to know its quantitative value, and its precise influence
on the vibration of a rotor-bearing system, one often has to be content with
an approximate value. In any case the extraneous damping in a system is often
greater than that from rolling--element bearings and this can usually only be
known approximately. With the following treatment the damping of a rolling-
element bearing can be roughly estimated.
We consider a linear spring of stiffness k and a linear damper of damping
d, one end of which is fastened to ground and the other moved harmonically
with frequency w. If the amplitude of movement is ii: then the amplitude of the
force in the spring is fA = kii: and the damping force Fd = dwii:. If Weq is the
frequency at which these two force amplitudes are equal, then

(9.32)

If k and Weq are known, then the damping coefficient can be calculated from

k
d=- . (9.33)
Weq

In rolling--element bearings the damping will follow quite a different law from
the stiffness and so a constant ratio kid and hence a constant Weq cannot be
expected. From the literature one finds that

Weq = (0.4 to 4) 105 (;)

and so
d = (0.25 to 2.5) 10- 5 k (:~) (9.34)

where k is in N / pm.
To obtain an estimate of the value of this damping, the idea of the damping
ratio of a simple vibrating system is employed, in which W is substituted for Wn
in equation (2.19). Hence
W
D=- (9.35)
2weq
9.5 Damping of Rolling-element Bearings 141

For rolling-element bearings with the above Weq values and f = 50 cycles/s,
that is W = 314 Tad/ s, the damping ratio lies in the region

D = 0.0004 to 0.004 .

It can be seen that rolling-element bearing damping is rather small and so for
most assumed frequencies, the rough estimate is a fair representation.
10 Shaft Seals

Turbomachines have seals situated in positions along their shafts, to prevent


the working fluid (liquid or gas) escaping to atmosphere or spilling over from
one stage to another. Fig. 10.1 shows an example of a boiler feed pump. Such
a pump is used to pump water at a pressure of between 250 and 450 bar, with
an intake pressure of only a few bar. In this example the seals are to be found

a - between the shaft and housing,

b - between the impeller shroud and the housing (neck ring seal),

c - between the hub and the housing (interstage seal),

d - at the balance piston.

The seals should be so constructed that the leakage loss is as small as possible.
This can be achieved in several ways, using either contactless or contacting
seals. We consider here the former, which can be classified by

Fig. 10.1. Boiler feed pump (KSB)


144 10 Shaft Seals

- smooth, cylindrical, conical or stepped seals (Fig. 10.2),

- labyrinth seals (Fig. 10.3) and

- floating ring seals (Fig. lOA).

Fig. 10.2. Smooth seals

tl~'lj
~~ Fig. 10.3. Labyrinth seals

Fig. 10.4. Floating ring seal

10.1 Restoring Force

To understand the influence of a seal on the dynamics of a shaft one needs to


know the force which acts on the shaft as a result of the action of the seal. It
is to be expected that there will be some influence on translational vibration
arising from a restoring force in the radial direction, that is from the resultant
force due to fluid pressure. For a perfectly circular seal and a central shaft
this restoring force is zero. A non-zero restoring force arises from a statically-
eccentric position of the shaft. For an angular tilt of the shaft, that is for a
change in its inclination, a restoring moment is also introduced, which can be
significant for a long seal. We do not consider this latter action here, but instead
assume the following situation (Fig. 10.5)

- Seal not moving.

Shaft turning with angular velocity fl and having translational movements


x(t) and y(t).
10.1 Restoring Force 145

Flow-

Fig. 10.5. Notations and coordinates of seal

Fluid (liquid or gas) with pressure PIon one side and P2 < PIon the other
side of the seal.

Because of this pressure difference a fluid flow exists in the clearance space,
which, as a result of the peripheral speed {} R and the translational velocities :i:( t)
and y(t) ofthe shaft, has not only an axial component but also a circumferential
component. In most cases there also exists at entry to the seal a certain swirl,
for example as a result of drag arising from friction.
Thus complicated flow relationships exist and corresponding pressure char-
acteristics, such that, in general,anon-linear restoring force is produced, with
components -FAt), -Fy(t).
For small displacements, however, these forces can be linearised. For dis-
placements around the central position, the following relation is found between
the displacements and restoring forces.

[:' ~,l {:} 1{:} + [ -: :

+[ -~ ~ 1{ : } ~ -{ ;,;:; } (10.1)

The restoring forces are thus proportional to a combination of displacement,


velocity and acceleration.
The velocity and displacement matrices indicate that the forces in question
are parallel neither to the velocity nor the displacement, but are inclined to
them by certain angles. In this respect seals thus behave like oil-film bearings.
The dynamic behaviour of a seal is determined from equation (10.1), using
the above assumptions, by the five coefficients kd, kc (stiffness), Cd, Cc (damping)
and md (inertia). Before proceeding to find these, we shall consider in the next
section a simple model showing the influence of seals on the dynamics of a rotor.
146 10 Shaft Seals

10.2 Jeffcott Rotor with Seals

Rotors can be considered as Jeffcott rotors for the purpose of obtaining an


approximate vibration behaviour. This is especially valid for the rotor of a
turbo pump such as that shown in Fig. 10.6. We replace it by the model shown
in Fig. 10.7, having parameters mo, Co, ko for the shaft and md, Cd, Co, kd, kc
for the seals. The equations of motion for free vibration x(t), y(t) are thus

(10.2)

Fig. 10.6. Cross-section through a turbo pump

Fig. 10.7. leffcott rotor with seals


10.2 Jeffcott Rotor with Seals 147

or, with z = x + jy,


(10.3)

With the substitution z = Z e.\t on obtains the characteristic equation


(lOA)

having eigenvalues

(10.5)

A general discussion of these eigenvalues would be too inconclusive and so a


special case will be taken as an example.

Example 10/1 Given: A shaft with mass

mo = 1000 kg, natural frequency Wo = 200 Tad/8 and damping


ratio D = 0.03, has two equal seals with the following parameters
kd = 0.25 ko , kc = 0.05 ko
Cd = 0.25 Co Cc = 0.05 Co
md = 0.1 mo

It is required to show the influence of these parameters on the eigenvalues.


Solution: The stiffness of the shaft follows from

Wo r;;, giving ko = 4.107


= y-;;;; N/m.

The damping coefficient follows from

1) = co/2v'komo, giving Co = 1.2 104 N 8/m.


Hence all the required parameters are known.
The eigenvalues are presented in Fig. 10.8. The real part (growth constants
all a2) are shown on the abscissa and the imaginery part (natural frequencies
WI, W2) on the ordinate. To achieve a sound understanding the model parameters
are introduced one after the other, so as to appreciate their influence.

Case a: Jeffcott rotor without damping and without seals.


al = 0, WI = 200 Tad/8
a2 = 0, W2 = -200 Tad/8

Case b: Consideration of rotor damping. For D = 0.03,


a = -0.03 200 = -6.0Tad/8 and so
al = -6.0 Tad/8, WI = 200 rad/8
a2 = -6.0 Tad/s, W2 = -200 rad/s
148 10 Shaft Seals

W1
W2 rod/s

c "-.250 ~e'f
.-d
---9
b-.200 -0

"r---r---,,---r--~l~-'----,r---r-
-20 -10 Y 10 rod/s

Fig. 10.S. Eigenvalues for example 10/1

Case c: Taking account of the direct stiffness kd of the seals, the values of the
natural frequencies will be increased, while the constants aI, a2 remain
almost unaltered:
al = -6.0 rad/s, WI = 245 rad/s
a2 = -6.0 rad/ s, W2 = -245 rad/s
Case d: Allowing for the cross-coupled stiffness ke of the seals:
al = 2.2 rad/s, WI = 245 rad/s
a2 = -14.2 rad/s, W2 = -245 rad/s
Thus al > 0 and the system is unstable.

Case e: Allowing for the direct damping Cd, the system is found to be stable:
al = -0.81 rad/ s, Wj = 245 rad/ s
a2 = -17.2 rad/s, W2 = -245 rad/s
Case f: Allowing for the cross-coupled damping Ce , there is hardly any change:
al = -0.86 rad/ s, WI = 246 rad/ s
a2 = -17.1 rad/s, W2 = -244 rad/s
Case g: Finally, allowing for the mass md, the values of the natural frequencies
are reduced and the a values are only slightly altered:
al = -0.42 rad/ s, WI = 234 rad/ s
a2 = -15.9 rad/s, W2 = -233 rad/s
10.2 Jeffcott Rotor with Seals 149

The above results naturally depend upon the choice of the parameter values,
but the trends are typical. Essentially, the most important feature is the desta-
bilising influence of the cross-coupled stiffness k c Occasionally, there can also
be a negative direct stiffness kd, whence the natural frequency becomes smaller.
Stability depends upon the relationship between the damping and the ex-
citation. An expression for this relationship is the ratio

(10.6)

where ~ C and ~ kc are the sum of the damping coefficients and the sum of the
cross-coupled stiffnesses, respectively and Wi is the natural frequency in ques-
tion. Stability can be expected when S > 1. In example 10/1, case d (unstable):
Wi 245 rad/s
~c Co = 1.2.10 4 Ns/m
~kc 2kc = 20.05 ko = 0.4 .10 7 N/m
and so S = 0.74.
For case e (stable),
Wi 245 rad/s
~c Co + 2Cd = 1.8.104 Ns/m
~ kc 0.4 . 10 7 N/m
and so S = 1.10.

The solution z = Ze>.t means that the natural motion is a circular orbit
with time-varying radius (a circular spiral). For positive imaginary part in .A
that is positive Wl, forward whirl is indicated and for negative imaginery part W2
backward whirl. Instability in forward whirl can be expected from the following
reasoning.
As a result of displacements x, y of the shaft a restoring force is called into
play, from equation (10.1). It has the components

(10.7)

If r is the radius of the circular orbit the energy supplied to the shaft along the
arc cpr will be
(10.8)
This is shown in Fig. 10.9. If this energy is greater than that necessary to simply
overcome the damping then instability will result.
In example 10/1 the seal coefficients were assumed to be constant. In reality
they depend upon the working conditions and hence also do the eigenvalues
and the resonance behaviour. Let us consider as an example the influence of
the speed of a turbocompressor. The average pressure generated is roughly
proportional to the square of the speed. The seal coefficients are proportional
150 10 Shaft Seals

Fig. 10.9. Explanation of equation (10.8).

to the pressure difference and hence also roughly to the square of the speed. For
the direct stiffness of the seal one thus has

(10.9)

where no is the reference speed for k dO . Neglecting Co and the other seal coeffi-
cients compared with kd' the natural frequency of the model (Fig. 10.7)

k
1+2~ - (n)2 , fo = J:... !ko . (10.10)
ko no 211"V~
Fig. 10.10 shows /I/ fo versus n/no for various values of kdO/ko. The speed is
critical, n = nC) when n = /I. With ft! fo = n/no = ne/no, it follows from
equation (10.10) that,

( k)
ne = no 1 - 2 ::
-1/2
. (10.11)

As is shown in Fig. 10.10 the critical speed increases greatly with increase
in stiffness of the seal and for kdO ~ ko it approaches infinity. For kdO ~ ko
resonance prevails in a wide range of speed. Fig. 10.11 shows how this affects
the vibration due to unbalance. In this figure is shown the radius of unbalance
vibration for kdO/kO = 0, 0.3, 0.4. For this purpose the damping ratio Dc
was assumed to be 0.1. The radius l' as a ratio of mass-eccentricity e is, from
Sect. 3.1, given by

(10.12)
where
{} kdO D-~
11 = wo' a = 1- 2ko- ' c -
2mOwe
, We = 211" ne . (10.13)
10.2 Jeffcott Rotor with Seals 151

0.1.5

3 0.30

0.15
2

~-=~~~~------------------o

O~------~~--~---,--------,-~-----
o 2 3

Fig. 10.10. Natural frequency from equation (10.10).

L 6
e -------0.1.

I.

0.3

----------------0
O~~~----r---------'--------'---------'---
o 2 3 I.
..!l_
no
Fig. 10.11. Radius of unbalance vibration from equation (10.12). Dc = 0.1.

The figure shows that the resonant peak becomes broader with increasing
stiffness of the seal. For a stiffness ratio of 0.4 the amplitude of vibration is large
over a wide speed range, in fact T ~ e/2D c Indeed the amplitude of vibration
is even greater than for this simplified calculation, because the cross-coupled
stiffness increases with speed and the system becomes less damped.
152 10 Shaft Seals

By consideration of other coefficients and for more realistic models the be-
haviour becomes more complicated, but the trends shown here remain essen-
tially valid.

10.3 Smooth Seals

In this section the relationships required to obtain the coefficients of the restor-
ing force will be given for smooth seals of the type shown in Fig. 10.5. In the
simplest case the shaft will be assumed to be stationary. Due to the pressure dif-
ference at the front surface an axial flow ensues, which when the seal is eccentric
has an unsymmetric velocity and pressure distribution around its circumference.
The resulting force acting on the shaft has an equal and opposite direction to
the shaft displacement. This case was investigated by Lomakin [11, 51) and from
his results one can obtain the direct stiffness kd of equation (10.1).
By rotation of the shaft the flow in the circumferential direction becomes
diverted, by means of which the restoring force has a component perpendicu-
lar to the shaft displacement. This is accounted for in equation (10.1) by the
coefficient kc. If the displacement is not static but takes place with a certain
velocity, then the restoring force contains the further terms in equation (10.1).
This general case was considered by Black et al. in [52, .. 55) and by Childs in
a review of this area in [56). He also published an improved theory in [57, 58],
whose results are summarized in the following.
A fundamental relationship is expressed by the equation for pressure difference,

(10.14)

where ~ is the entrance loss coefficient


p is the fluid density
V is the average axial stream velocity

and
L
(T =).- (10.15)
C
where). is the resistance coefficient.

In [57) and other work a factor Llp R7r /). is assigned to describe the components
of restoring force. Overall results are presented using the factor Llp LR/ C, by
which means the factor 7r / (T can be varied independently. Hence, from equa-
tions (31) and (32) of reference [57) the following relationships are given for the
coefficients of the restoring force
10.3 Smooth Seals 153

Cc = a2 (nT) c* (10.16)

with the reference values of the corresponding quantities

k* = Llp L R , c* = k* T , m* = k* T2 (10.17)
C
and the transit time
L
T=- (10.18)
V
The dimensionless coefficients are

ao 2.5 A E

al = 2 A[! + ~ (E + ~) ] (10.19)

a2 : (E+~)
with
E_ l+e
, - 2 (1 + e+ B IT)
(10.20)
V
, b= R n .
e,
The dimensionless coefficients are thus functions of IT and b. These coefficients
are given in Fig. 10.12 for the frequently occurring value = 0.5. e
When n = 00, then from equation (10.20) b = o. It can thus be seen that
the coefficients depend only weakly upon n.
The direct stiffness kd has a component which depends upon n 2 (equation
(10.16)). However, for most practical purposes this has little importance. The
cross-coupled stiffness kc is proportional to n and so the system becomes less
stable as speed is increased.
To determine the values of ao, al and a2 from Fig. 10.12 the prevailing
values of IT = ALIC and b are required. From reference [57],
2] 0.375
A = 0.066 R;;l/4 [1 + (2\) (10.21)

for a smooth annulus and for turbulent flow, where

V C , the Reynolds Number for axial flow (10.22)


v
v p, , the kinematic viscosity. (10.23)
p
154 10 Shaft Seals

Co 0.5

o I i I I 1 I I
Fig. 10.12. Coefficients aQ, at
0.1 10
u- and a2 for e= 0.5.
The value of b is defined by equation (10.20). With

the Reynolds number for circumferential flow, (10.24)


v
Ra
b (10.25)
Rc
To obtain the value of A we need the average velocity V to be inserted into
equation (10.22). Finally, from equation (10.14)

V=
Vp(I+~+20")
2..::1p (10.26)

Since the desired value of A is also a function of 0", it is best obtained iteratively.
Fig. 10.13 shows the resistance parameter from equation (10.21). In fact A is
larger than Fig. 10.13 would indicate, due to any roughness of the seal surfaces.
The significant parameter affecting any increase in A is the relative roughness
e/C, where e is the actual roughness.
Expressions (10.16) are valid, on the assumption that the fluid on entering
has a tangential velocity (inlet swirl) U c = U /2, where U = R [}. In reality the
inlet swirl can assume other values. These can be accomodated in the general
10.3 Smooth Seals 155

0.010

0.001

Fig. 10.13. Resistance parameter


Ra- from equation (10.21).

relationships of reference [57]. An example is given in ref. [57] in which kd, kc


for U c = U /2 and kd' k~ for U c = 0 give

k~ = 1.08 kd and k~ = 0.48 kc .

It can be seen that inlet swirl influences appreciably the cross-coupled stiffness
and that therefore stability can be improved by reducing the inlet swirl.
In reference [57], the length of the seal was considered in only an approxi-
mate way (short-bearing solution). An improved theory (finite-length solution),
considered in ref. [58], gave certain differences. For L < 0.2 D the results remain
questionable.
In this section, we have only considered results from theoretical calculations.
In most publications the calculated values are compared with measurements
from test facilities. In this way the basic theory is essentially confirmed, although
certain differences remain. Further differences are evident in real machines. It
is recommended therefore that a tolerance of 30 % be used on the coefficients
in rotordynamic calculations.
For rough estimates, some numerical values are given in the following list.
These have been obtained from calculations on actual seals:

c ~ 0.003 R
T 0.5 ... 2 ms
kd (0.4 ... 0.6) k*
kc (0.2 ... 0.5) k*
Cd (0.7 ... 1.4) c
Cc (0.1 ... 0.6) c
md (0.2 ... 0.7) m .

Other values are however possible but the mass md is in most cases unimportant.
156 10 Shaft Seals

10.4 Labyrinth Seals

More shapes are in evidence for labyrinth seals than for smooth seals. Any
discussion about their force-displacement behaviour is correspondingly more
complex. In the following a short overview will be given of results from theory
and experimental testing. The work [59) of Alford should first be mentioned,
where, amongst others, the behaviour of a seal with a chamber is discussed.
Using improved assumptions Spurk and Keiper [60) report somewhat different
results. Wachter and Benckert [61, 62, 63) describe comprehensive experimental
investigations, a typical result from the latter being shown in Fig. 10.14. This
shows the direct stiffness kd and the cross-coupled stiffness ke for a labyrinth
seal, when non-dimensionalised by dividing by k* (equation (10.17))1 and plot-
ted against E u , where
- P 2 1
(10.27)
Eu = - U po
2 L1p+ -2 u a2

in which Po, P are the density at inlet and the average density,
U = OR the tangential velocity of the shaft,
ua the axial stream velocity at inlet,
L1p = P1 - P2 the pressure difference.

kd 0.08
k"
~
k" 0.06

0.04

0.02

I:)
0 0.2 0.4 0.6
E
u -

Fig. 10.14. Stiffnesses of a labyrinth seal. From reference [62].

IThese authors used other expressions. To obtain a better understanding we use our own
expressions throughout.
lOA Labyrinth Seals 157

In addition to the variable E u , the following will also be used

- po 2 1
E. = -2 Uc po (10.28)
L1p+ -2 u a2

where U c is the stream velocity in the circumferential direction. The expression


Pou~/2 is the stagnation pressure and also the specific kinetic energy of entrance
swirl. E. can thus be regarded as the swirl energy, non-dimensionalised by the
total pressure energy.
The variable Eu is proportional to {J2. Accordingly, the stiffnesses in
Fig. 10.14 have a roughly quadratic dependence on rotor speed. In some pub-
lications other dependences on rotor speed are given. Similarly, very different
dependences are given relating the other important variables. However, in gen-
eral, it can be said that the non-dimensional coefficients of labyrinth seals are
somewhat smaller than for smooth seals.
Iwatsubo [64] gives an analytical method, in which the stiffness and damping
coefficients are presented as summations. Childs and Chang-Ho Kim [65] give
extensive comparisons between theoretical and experimentally-determined co-
efficients for labyrinth seals with different geometries under realistic conditions.
The differences between theory and experiment lie within tolerable bounds.
Scharrer [6q] gives equations by which stepped seals with increasing or de-
creasing diameter can be considered. Parameter studies show that seals with
increasing diameter are more stable than are those with constant diameter.
Such stability depends slightly on whether the steps are on the rotor or stator.
Certain restrictions about the geometry of the seal and about the type of
flow were set when computer methods were used in the above research pro-
grammes. Nordmann and Dietzen [67, 68, 69] allowed more freedom in the
above respects. Their published method utilized finite difference solutions of the
Navier-Stokes equations. Weiser [70] gives a clear comprehensive overview and
further calculations for coefficients for seals of any geometry. Finite-difference
methods were used on a bulk-flow theory similar to the work of Childs and
Scharrer [65, 66] as well as approximate equations for the dynamic coefficients.
All procedures were appropriate to compressible, turbulent flow. Further publi-
cations are referred to in the bibliography. Although there is still much research
to be done, rotordynamic problems can be solved reasonably well using the
known methods. This is demonstrated, for example, in the work of Schmied [71]
and of Kanki et al. [72].

10.5 Floating Ring Seals

A simple floating ring seal is shown in Fig. lOA. The sealing ring is mounted
in the housing, so that it cannot rotate but is free to move axially. Because of
158 10 Shaft Seals

the difference between the pressures on the inner surface of the housing and the
outside, the ring is forced against the housing in an axial direction and held fast
by a sufficiently high friction force. The free movement during startup, when
the pressure difference is still small, allows a much smaller radial clearance than
for the rigidly held seal. In use C / R ~ 0.001, while for smooth seals about 0.003
is usual.
Another arrangement of a floating ring seal for a gas compressor is de-
scribed by Kirk et al. in [73, 74J. The seal consists of an inner ring and an outer
ring (Fig. 10.15). In the space between the two rings oil is supplied at a pres-
sure somewhat greater than the gas pressure within the housing. An additional
labyrinth seal operates such that neither gas can enter the oil space nor oil enter
the gas space. Some oil penetrates through to the outer ring space.
Floating ring seals have a smaller clearance than do the other kinds of seal.
Hence the axial stream velocity is lower and the flow remains laminar. In addi-
tion the compressibility of medium oils can be neglected. Therefore Reynolds'
lubrication theory may be used to find the restoring force, in much the same
way as is done for oil-film bearings. The rings are usually short (L/ D ~ 0.02
to 0.20) and so the short-bearing theory of Sect. 6.2 suffices.
A special problem in dealing with floating-ring bearings is the determination
of the position of the ring in the locked-up condition. In reference [73J numerical
investigations were carried out and as a result the ring was found to follow an
irregular path at startup and, after a few revolutions of the rotor, to become
fixed. In the example considered six revolutions were completed by the outer
ring and twelve by the inner ring. The authors found that, in the fixed position,
the eccentricity had a value similar to that for static loading due to the weight
of the ring. This eccentricity can be calculated from equation (6.30) or found
from Fig. 6.11, if Fo is made the weight of the ring.

Fig. 10.15. Floating ring seal for a gas compressor. From reference [73].
10.5 Floating Ring Seals 159

Schmied came to a somewhat different conclusion in reference [71]. He found


by numerical estimation that the ring initially performed a whirl with eccen-
tricity ratio about 0.4 to 0.5, and whirl frequency about 0.4 {}, until it finally
settled. In the locked-up position it thus achieved the afore-mentioned eccen-
tricity in a random position within the clearance circle. It is thus recommended,
in rotordynamic calculations, that the likely effects on stability and natural fre-
quencies by taking up different positions be appreciated.
11 Steam Whirl

In reference [13] Thomas reported instabilities in steam turbines, which could


not be explained at that time by any known cause. In particular, it was not
understood why the instability commenced at a certain level of power. Bearing
instability, that is classical bearing instability, could not be responsible, as this
commences at a certain rotor speed. In reference [13], it was suggested that
a plausible cause might be so-called steam whirl. This arises in an axial flow
machine when a radial displacement of the shaft takes place as a result of
different leakage losses around its circumference. A resultant force is produced
which is perpendicular to the displacement. Alford [75] described a similar effect.
In the following a simple means of calculating this excitation force will be derived
and it will be shown how this force influences the vibration behaviour of the
rotor.

11.1 The Steam Whirl Force

Consider one stage of a steam turbine as shown in Fig. 11.1, generating power P
and having angular velocity il. With R as the radius to the mid-blade position,
the tangential force Fu is given by

(11.1)

The non-dimensional tangential force per unit length lu is thus given by

Fu P
Ill. = 2 7r R = 2 7r il R2 (11.2)

When the rotor is positioned centrally the clearance between the blade and
the housing around the whole circumference is equal to c. For an eccentric
position a variable clearance exists around the circumference. Thus for a small
displacement x > 0, y = 0, the clearance at a position defined by the angle cp is

c", =c- x cos cp = c + Llc . (11.3)

The tangential force changes somewhat with blade clearance, becoming smaller
as the clearance increases. For simplicity a linear relationship is usually assumed,
162 11 Steam Whirl

y y

Fig. 11.1. One stage of a steam turbine.

(11.4)

where fuo is the reference tangential force for a blade with zero clearance and
the factor 'Y takes account of the clearance effect.
Putting
Llfu = B/
Be
u
Lle = - fuol'Y (-x cos I{) ~ fu I'Y x cos I{) (11.5)
the tangential force at a position defined by the angle I{) is

fu,/{, = fu + Llfu = fu (1 + x f cos I{)) , (11.6)

if the ~ sign is replaced by an equality.


The resultant of all these tangential forces around the circumference has
the components

2,..
J fu,/{' R sin I{) dl{) = 0
o (11. 7)
2,.. 'Y
[ fu,/{, R cos I{) dl{) = 7f1 R fu x = ksx

where
(11.8)
To derive ks , a known value of:l: > 0 with y = 0 would have to be assumed.
Thus for a small random displacement having components :1:, y,

(11.9)
11.1 The Steam Whirl Force 163

A force F. = ks r is thus produced, which is always perpendicular to the resul-


tant displacement r. F.. is termed the steam whirl force and ks the coefficient
of steam whirl.
The formation of Fs can be explained by reference to Fig. 11.2. For an
eccentric position of the rotor, the forces acting around the circumference on
the side having the reduced blade-clearance will be greater than those acting
on the side with the increased clearance. Hence the resultant circumferential
force on one side is F + t1F and on the other side F - t1F. Their resultant
Fres = 2t1F is the steam whirl force Fs.
Expression (11.8) describes the fundamental parameter of steam whirl. The
factor I can be concluded from references [76, ... 79J. From [80J 1=2 K2 with

K2 ~ 2 - 0.4(1/1 -1.5) 0.2 (11.9a)

for a 50 %-reaction stage. For a weak reaction stage

K2 ~ 3.2 - 0.27(1/1 -1.5) 0.2 (11.9b)

Fig. 11.2. Formation of the force Fs

In these equations
.1. = 2 t1h. (11.9c)
'f' 2'
U

is the pressure or head factor for the blading. I is typically between three and
five.
For blades with shrouds, the influence of the pressure between the shroud
and the housing adds to the steam whirl. This is the same effect as that produced
by seals as discussed in Chap. 10. The coefficient ks of steam whirl has the same
meaning as the cross-coupled stiffness kc in seal excitation. Both are added in
the equation of motion. Mostly, they are of about the same order of magnitude.
Baumgartner [81J gives an overview of the subject and a scientific method for
calculating the coefficients of seal excitation and of steam whirl for blading.
164 11 Steam Whirl

11.2 J effcott Rotor with Steam Whirl

Steam whirl has a destabilising influence. The most important effects on the
stability of a rotor can be appreciated by considering a model of a Jeffcott rotor
if it allows steam whirl at its disc. With parameters mass m, stiffness k, damping
d and coefficient k. of steam whirl, the equations of natural vibration are

mz + dz + kz + k.y = 0
(11.10)
my + diJ - k.z + ky = 0

or with z = z + jy
mz + di + (k - j k.) z =0 . (11.11)
The substitution z = Ze>.t leads to the characteristic equation
(11.12)

with the solutions or eigenvalues

(11.13)

in which
d 2 k 2 k.
5=- W =-
m W =- (11.14)
2m n
m
If one relates the eigenvalues to W n , that is the natural frequency of the shaft
without damping and steam whirl, then one obtains the expression

-
Ai
= -D VD2 - 1 + J. -k. = -
G:i
J -Wi , i = 1,2. (11.15)
Wn k Wn Wn

5
where D = - .
Wn
An eigenvalue has zero real part at a stability borderline and so A = jw. In
this case the characteristic equation takes the special form

- m w2 +j dW +k - j k. = 0 . (11.16)

This gives
- m w2 +k =0 and d W - k. =0 (11.17)

Hence at the stability borderline the natural frequency is


11.2 Jeffcott Rotor with Steam Whirl 165

and
ks = d Wn ( 11.18)

The condition for stability is thus

ks < d Wn (11.19)

Using the damping ratio D,

Using equation (11.8), the limiting power for stable running is given by

D
Pth = 4 - k 1 [} R (11.20)
I

The forces acting on the disc, at the boundary of stability are shown in Fig. 11.3.
This illustrates equal and opposite forces, on the one hand from excitation and
damping and on the other from centrifugal force and spring force.
The fundamental character of the natural frequencies and decay constants
and their dependence on steam whirl is evident from Fig. 11.4, in which D = 0.1.
The value of natural frequency increases with ks but varies only a little from
w n One real part increases and becomes positive at the stability boundary
(ks/k = 0.2 = 2 D). The other real part decreases at the same rate and the
associated backward whirl thus becomes ever more damped.
From equation (11.19) it can be seen that instability due to steam whirl can
be avoided by sufficiently high damping and natural frequency. This is valid not
only for the model considered, but is generally true for real rotors. However,
the stability for real rotors depends on more parameters, especially on those of
the journal bearings. Consider therefore a model of a leffcott rotor supported
on oil-film bearings. The equations of motion can be obtained from equations
(7.3) and (7.5), in which ks Y2 is added to the first term on the left-hand side
of equation (7.3) and -k s Yl to the second term.

y I--------:;;q:r

x
Fig. 11.3. Forces on a leffcott rotor having
steam whirl, at the stability boundary.
166 11 Steam Whirl

1.02
w,
wn 1.01
_ W2
wn 1.0
0.2 0.4
0.99 ~-
k

0.1 ~
Wn
~ 0
wn
0.2 0.4
(X2 - 0.1 ks_
wn k
-0.2

-0.3 Wn

Fig. 11.4. Natural frequency and decay constants for leffcott rotor with D == 0.1 and
subject to steam whirl.

References [82, ... 86] describe investigations using this model. A general
discussion is not possible because of the presence of the many controlling pa-
rameters. An example will therefore be considered.

Example 11/1 From reference [82]. Investigation of the HP stage of a


steam turbine having a power output of 100 MW.
Data for HP stage

Power P= 25 MW
Operating speed n= 3000 rpm
Rotor mass m= 5000 kg
Rotor stiffness, rigid bearings k= 21.9 . 107 N/m
Natural frequency, rigid bearings In = 2000 cpm
Steam whirl coefficient at 25 MW ks = 1.25 . 10 7 N/m
ks/k = 0.057
It is required to determine the influence of natural frequency, of damping
and of bearing type on the limiting power for stability.
To carry out the investigation the HP rotor is separated from the rotor
train and assumed to be a Jeffcott rotor in oil-film bearings. With cylindrical
bearings, one employs the four stiffness and four damping coefficients used in the
calculations of reference [82]. Fig. 11.5 shows the characteristic of limiting power
11.2 Jeffcott Rotor with Steam Whirl 167

100
~h d =5.10 4 Ns 1m '-...
MW
Unstable d=O Stable

50 52MW
Operating power
------_\._--- 30MW

0
0 1500 2000 cpm
I
1540 fn -
Fig. 11.5. Ezample 11/1. Power limit and its dependence on natural frequency

against natural frequency of the rotor. The lower curve refers to no damping
at the disc and the upper curve to damping with d = 5.104 Ns/m, that is a
damping ratio of D = 0.024 at In = 2000 cpm. In the case of no disc damping the
limiting power is zero for In = 1540 cpm. It follows that, at the operating speed
of 3000 rpm, bearing instability occurs, if the natural frequency of the rotor
remains at 1540 cpm. At the given natural frequency of 2000 cpm the limiting
power is 30 MW, lying some 20 % above the operating power. Increasing the
natural frequency by 10 % to 2200 cpm gives an increase in power limit of 73 %
to 52 MW. Consideration of disc damping with D = 0.024 gives a 57 % increase
in permissible power to 47 MW. It can be seen that the curve corresponding to
this damping for natural frequencies below about 1500 cpm is fairly flat. Here
one must increase the natural frequency by a great extent in order to increase
the power limit. The flat region of the curve can be regarded as representing
bearing-induced instability in the absence of disc damping.
The influence of bearing type on the power limit is shown in Fig. 11.6 for
the case of no damping at the disc. In comparison with cylindrical bearings,
the other bearing types allow a 2.3 times higher power limit of 68 MW, at a
natural frequency of 2000 cpm. This example shows that the influence of natu-
ral frequency and of damping depends strongly on whether or not the rotor is
operated near bearing instability conditions.
Further special cases can be considered such as additional damping at the bear-
ings, which can be more easily realized than additional damping on the rotor.
If steam is introduced through nozzles, then a resultant tangential force acts on
the rotor. This has an order of magnitude similar to the rotor weight. Hence
different stability features can exist from those for which the bearings are loaded
only by rotor weight.
168 11 Steam Whirl

Four-lobed bearing

~h MW

100

-6~-~'-- 68 MW

50

- d - - - - - 30 MW

o
o 1500 2000 cpm

Fig. 11.G. Example 11/1. Influence of bearing type on power limit. Disc damping
d= O.
12 Internal Damping

As well as bearing instability and steam whirl a further cause of self-excitation


is internal damping of the shaft material or due to micro-movement at shrink
fits or couplings. The first was reported by Newkirk and Kimball in 1925 [14, 15,
16]. The mechanism is very complex but can be understood by the assumption
of a viscous law. This will be considered in the next section.

12.1 Jeffcott Rotor with Internal Viscous Damping

The vibration of a Jeffcott rotor, represented by displacements :z: and y follows


from Newton's Second Law in the following equations.

mi = -F""e + FAt)
(12.1)
my = -Fy,e + Fy(t) - W
where F",(t) and Fy(t) are the excitation forces and W = mg, the disc weight.
All remaining forces acting on the disc are represented by F""e and Fy,e. It is
assumed in the following that they act on the rotor in the opposite sense to
F",(t) and Fy(t), that is through the negative sign.
To deduce F""e and Fy,e the rotor will first be assumed to be stationary.
Using the coordinates depicted in Fig. 12.1, the stiffness k and the coefficients
da and d j representing the external and internal viscous damping give,

F""e (il = 0) = k:z: + (da + di):i:


(12.2)
Fy,e (il = 0) = ky + (da + di)y .
If the shaft is now rotating and :z: and y i- 0, the individual fibres of the shaft
have additional velocities and therefore generate additional damping forces.
To appreciate these, we first put dj = 0 in equations (12.2) and transform to
rotating coordinates. We then introduce for the internal damping the terms dju
and djv.
It is appropriate to use complex quantities

z=:z:+jy (12.3)
From equations (12.2), with d; = 0 we have

(12.4)
170 12 Internal Damping

I
I
---t---
I
I

Fig. 12.1. Coordinates for the leffcott rotor

If one now puts


( = u + jv = re ia (12.5)
then
z =r ei(OHa) =( eint
z = (i n ( + C) eint
(12.6)

Correspondingly for the forces

(12.7)
and
Fz = Fe i (nt+f3) = F, eint (12.8)
Substituting these equations into equation (12.4) gives the desired relationship
in rotating coordinates

Finally, including the force arising from internal damping, we have

F, = k ( + da (j n ( + C) + di C . (12.9)

To employ equations (12.1) one must transform equation (12.9) back to sta-
tionary coordinates. This can be done by using equations (12.8) and (12.6).
Thus
Fz = (k - j d; n)z + (da + di)i (12.10)
or in terms of real equations

kz + din y + (da + di)z


(12.11)
-din z + ky + (da + di )1i

since equation (12.10) must be satisfied by individual consideration of its real


and imaginary parts.
12.1 Jeffcott Rotor with Internal Viscous Damping 171

From equations (12.11) and for a rotating shaft, the force (Fx.,,,, Fy ,.,) consists
of steady-state components

kx + di il Y
(12.12)
-di il x + ky

and dynamic components

(da + di)x
(12.13)
(d a + dJiJ
The dynamic components each consist of an external and an internal damp-
ing force. Each is proportional to vibratory velocity and has the same direc-
tion. From the steady-state component one has together with the spring force
(kx, ky) the term (d;ily, -d;ilx) as a result of internal damping. To clarify
this, consider the case Fx,. = 0, Fy,. = - W, that is a static load due to the
weight of the rotating shaft. Solving equations (12.12) for displacements x, y,

di il d; il
x = k 2 + (d; il)2 W:::::! kz W
(12.14)
-k 1
y = k2 + (d i il)2 W:::::! -k' W .

In the majority of cases (d; il)2 k 2, so that in equations (12.14) the right-
hand terms are sufficiently accurate.
As well as static displacement in the negative y-direction the rotor centre is
displaced in the x-direction. The resultant displacement has a direction inclined
at an angle a = arctan (d i il / k) with respect to the load direction. This is shown
in Fig. 12.2.

d1I
-'-w
k 2

Fig. 12.2. Displacement due to static load of rotating shaft having internal damping.
172 12 Internal Damping

The sideways displacement can be explained as follows. The stress in the


cross-section of a fibre with strain e and rate of strain i at a random point on
the shaft is
u=Ee+E'i (12.15)
in which E' represents the corresponding damping modulus (N 81m 2 ). With dA
as the cross-sectional area of the fibre the stiffness force acting at the same
point is
dF.=EedA (12.16)
and the damping force
dFd = E' i dA . (12.17)
Now consider the fibres A, B, C, D of Fig. 12.3. For the loading shown the
fibres A and B each have strain -e and hence experience a compressive force
-dF. The fibres C and D have strain +e and hence experience a tensile force
+dF. These forces produce a moment dM = 2dF.a. The resulting moment for
the whole cross-section is in equilibrium with the external moment due to load
w.

Stiffness forces Damping forces

Fig. 12.3. Fibre forces for a statically loaded, rotating shaft. Equilibrium forces not
shown.

The damping forces are shown in the right-hand diagram of Fig. 12.3. Be-
tween I and II the fibres will be extended and between II and I shortened. Hence
in Band C the rate of change of strain i is positive and in D and A negative.
The damping forces have corresponding signs. For equilibrium these forces must
oppose an equal stiffness force dFc Hence

or
E' i dA + E ec dA = 0
12.1 Jeff'cott Rotor with Internal Viscous Damping 173

whence strain
ec = -E
E' e (12.18)
The damping forces thus give a strain -ec in Band C and +ec in D and A.
This is true for all fibres in each half of the cross-section. Hence the shaft bends
and gives rise to displacement in the x-direction.
As for the loading with weight W, so for a loading due to random static
load in the x, y plane. Working in reverse from a displacement r = (x, y), forces
can be seen to act on the shaft, as shown in Fig. 12.4.

Fig. 12.4. Forces acting on the rotating shaft for static deflection, T.

It can be appreciated that a forward-whirl will lead to instability, if a suf-


ficiently large damping force does not act to counter the force dJJr. This is
indicated in the following equations of motion. Using equations (12.11) and
(12.1), we obtain

ma: + (do + di)x + kx + d; {}y= FAt)


(12.19)
my + (da + di)iJ d; {}x + ky = Fy(t) - W
The left-hand sides of these equations correspond to the left-hand side of equa-
tions (11.10) for steam whirl, where do + di replaces d and dj { } replaces k.
Hence from equation (11.19), for stability

(12.20)

and the angular velocity at the stability boundary is


n da+di
Uth = -d-j- Wn (12.21)

If the external damping is zero, then stability is maintained only for speeds
up to {} = wn This is because the internal damping cannot only excite but
174 12 Internal Damping

also exert a damping effect. Increasing external damping delays the stability
borderline to a correspondingly higher angular velocity (Fig. 12.5).
The unstable motion ensues as a forward whirl offrequency W n For example,
if do = 1.2 di then the stability borderline occurs at nth = 2.2 Wn , with the
motion having a natural frequency Wn = n th /2.2 = 0.45 nth.

stable

O-------~------_.--
o 0.5
Fig. 12.5. Stability chart for Jeffcott
rotor with internal damping.

12.2 Damping due to the Shaft Material

In Sect. 12.1 damping due to the shaft material was represented as internal
damping to distinguish it from external damping. It was given the coefficient d;
for use in considering the vibration of the J effcott rotor. In this section the prop-
erties of this material damping will be briefly discussed and the determination
of the coefficient di elucidated.
Consider a bar of length I and cross-sectional area A, which is loaded by a
periodic force F. It is thus subject to a stress U = F/A and experiences a strain
e = 111/1. During one period the function u(e) describes a so called hysterisis
loop as shown in Fig. 12.6. For a static load the loop shrinks to the dashed
curve. Hence, as can be seen in Fig. 12.6, Us is the static and Ud the dynamic
component of stress.
In the following we shall confine ourselves to a linear elastic loading, in
which case the dashed curve becomes a straight line, governed by Hooke's Law.
The path around the u, e loop is equivalent to the work done per unit volume
against damping during one period, namely

w = f u(e) de . (12.22)

The work done per unit volume in deformation during the first quarter-period
is, from Fig. 12.6
1 ~ ~
u = - u e (12.23)
2
12.2 Damping due to the Shaft Material 175

E-

Fig. 12.6. Hysterisis loop

Now define the ratio


w
"p=- (12.24)
u
as the relative damping (specified damping capacity) of the element and

"p
77= - (12.25)
27['

as the loss factor.


The force-displacement law of internal friction required for vibration predic-
tions follows from the dependence of Ud on the strain c; and its time derivative.
In general this dependence is non-linear. A somewhat simpler concept is the
specific damping"p. However, measurements have indicated a rather more com-
plicated dependence. It is beyond the scope of the present work to consider this,
but reference can be made to many publications ([16], [87, ... 91] and others).
Here suffice it to say that the specific damping of the shaft material depends
upon the material, the type and magnitude of its loading, the frequency and on
other influences. In any given case one must decide how far one must consider
these influences. As a guide linear visco-elastic behaviour is assumed, which
will now be developed in the context of shaft vibration.
From equation (12.15)
U = E c; + E' Ii

For a harmonically varying strain, put

c; = f sin wt, Ii = fw cos wt = WVf2 - C;2 (12.26)

Hence
U = E c; + E*vf2 - C;2 = Us + Ud (12.27)
176 12 Internal Damping

where
E* = E' w , (12.28)
the loss modulus. The static component (T. = E e, as mentioned previously,
conforms to Hooke's Law. The dynamic component (Td = E*J2 - e 2 is zero
for e = e and is equal to E* e when e = o. Between the two extremes the
characteristic is elliptical. This produces the hysterisis loop shown in Fig. 12.7.
The work done against damping, equation (12.22), is

W = 7r E* tJ

The work done in deformation is


1 ~
u=-Ee
2
and hence the relative damping is
E*
.,p=27r- (12.29)
E
and the loss factor
E*
'1/=- (12.30)
E
To obtain a relationship for the coefficient di of internal damping of the leffcott
rotorfor no rotation (0 = 0), one proceeds using the work done by the damping
force and by the stiffness force during harmonic displacement:c = :c sinwt. With
amplitude Fd = d; w z the damping work done in one period is

(12.31)

as=EE

Fig. 12.7. Hysterisis loop for viscoelastic material


12.3 Damping from Assembly Components 177

and with amplitude F. = k x the deformation work is


u= ~ k li 2 (12.32)
2
For a shaft of volume V, it follows that the relative damping is

(12.33)

For a given "p or given loss factor 1/,

(12.34)

By analogy with damping ratio equation (2.19), for internal damping the equiv-
alent damping ratio is given by

D _ d; Wn
(12.35)
- 2 k
This gives the relationship
W
1/ = 2 D i - (12.36)
Wn

To calculate d; from equation (12.34), "p or 1/ must be known and a vibration


frequency assumed. The vibration calculation then gives a natural frequency Wx
at the stability boderline. If this value agrees to a sufficient accuracy with the
assumed value w, then this value can be accepted. Otherwise one must iterate
until the difference is sufficiently small.
Approximate values of relative damping can be found in the literature. For
metals, rough values range from 0.001 to 0.1, and in general the higher the
stress, the higher the 1/J value. For W = W n , Di = 0.0005 to 0.05 and it follows
that in general the internal damping for metals is very small.

12.3 Damping from Assembly Components

In Sect. 12.2 linear material damping was assumed for the shaft and as a result
its effect on the stable operation of a rotating system could be ascertained.
Similar effects arise from shrinkfits and flexible couplings (Fig. 12.8), which are
defined generally as comp'onent damping. An estimate at the end of this section
shows that component damping is essentially greater than material damping
for a comparable length of shaft. Hence component damping is much more
important than material damping in determining stability.
In order to consider the component damping in the equations of motion one
needs the relation between the bending moment and the difference angle along
178 12 Internal Damping

Fig. U.S. Machine components


having damping

the component . For the toothed coupling the model shown in Fig. 12.9 will be
used. This model can also be used for shrink fits as will be explained at the end
of the following derivation.
The figure shows the two halves of the coupling in separate sketches. The
angles of rotation (tilt) can be different and both will be assumed small, and
they can be displayed as vectors. Using the notation of Fig. 12.9 the components
of the difference angles are <P = <pr - <PI and "p = "pr -"pl. The components of
the bending moment are shown as M<p and M.p . In a coordinate system fixed to
the shaft and having angular velocity {} the corresponding notations are Q, (3
and MOt, M{3.
The bending moment at the coupling arises from the tooth forces and their
distance from the axis of rotation. We assume a linear force-displacement re-
lationship such as applies to a spring and distribute the forces continuously
around the circumference of the circle of radius R. For the difference angle Q,
the relative displacement at the position R, /j is

Xc = Q R sin /j

lit

Fig. 12.9. Model for component damping


12.3 Damping from Assembly Components 179

and the displacement force on the arc of length R dli is

dF =~
271"
a R sin Ii dli

where k is the axial stiffness of all the teeth.


The moment of this force about the axis of rotation is

dMo. = dF R sinli .

By integration the resulting moment

Mo. =! -
o
271" k
271"
a (R sin 1i)2 dli = R2
k - a
2
(12.37)

The rotational stiffness of the model is thus


M R2
~
k= ~ = k - (12.38)
a 2
If one further assumes uniformly distributed damping of coefficient d for the
whole circumference then the moment produced for a rate of change of angle a
can be obtained in a similar way as

(12.39)

where coefficient
(12.40)

For later comparison we now determine the work done during one period by
these moments for harmonic variation aCt) = a sinwt.
Thus

! (Mo. + Me" a dt
271"/w

We ,0.
o

! (k
271"/W

d sin wt+ w cos wt) &2 cos wt dt


o
71" d W&2 (12.41 )

This consists, as we know, of only the damping work and is represented by the
surface of the ellipse in Fig. 12.10.
Consider again the moment itself. From equations (12.37) and (12.39), for
rotation a( t)
180 12 Internal Damping

aw& ----::=-------'
11-
Fig. 12.10. Damping moment for harmonic
variation of Q.

and correspondingly for rotation (3(t)

These are the components of the moment in coordinates fixed to the shaft and
also for a stationary shaft, for which n = o.
The components are now required for a system rotating with angular veloc-
ityn. These can be found in an analogous way to the method used in Sect. 12.1.
Thus,
k cp+ d; n.,p + (d a + d;)cp
(12.42)
- d; ncp+ k .,p + (d a + di)"j;
~ ~

in ~hich external damping of coefficient d a is included and d is distinguished


by d;.
With these equations we can consider a toothed coupling in the equations
of motion as is shown in Subsect. 16.2.3. In a similar way one can, at least ap-
proximately, consider the effect of a shrink fit, if one assumes the corresponding
shrink-fit coefficients k and d as in equations (12.38) and (12.40).
Internal damping within a toothed coupling or a shrink fit causes instability,
just as does the material damping of the shaft. If one assumes such an element
in the centre of the shaft, an analogous equation to (12.21) can be obtained for
the stability borderline, as the following will show.
For circular motion of the shaft centre, with frequency Wn the rotational
motions are
cp = rp cos Wn t f , .,p =;fi sin Wn t
and the work done by the moments in one period is

27r/wn
We = / (M<p cp + M", "j;) dt = 2 7r [(d a + d;) W n - d; n] rp2 (12.43)
o
12.3 Damping from Assembly Components 181

Stability is assured for positive work done We. The stability borderline is en-
countered for an angular velocity

n _ da + d; (12.44)
Uth - ~ Wn
d;
According to the above considerations linear viscoelastic behaviour has been
assumed, which corresponds roughly to reality. More realistic is Coulomb damp-
ing, which will be now investigated.
For Coulomb damping in a one-dimensional element, the force arising from
a displacement a:(t) is given by

F(t) = Fe sgn(:i:) (12.45)

For constant Fe and harmonic displacement a:(t) :v sinwt, F(t) has the
characteristic shown in Fig. 12.11 and in one period the work done is

We = 4 Fe:v . (12.46)

Lund [92] distinguishes between solid damping with Fe = ke :v and dry friction
with Fe = constant.
Considering the model of Fig. 12.9, the equations corresponding to equations
(12.45) and (12.46) are as follows.

M",(t) = Me sgn (a) (12.47)

and
We =4 Me a. (12.48)
Coulomb damping is non-linear. To carry out an equivalent linear calculation
an effective coefficient of viscous damping will be used, for which the work done
We will be taken ~o be that given by equation (12.41). As a result, the dry
friction coefficient de is given by
~ 4 Me
de = - --::::: (12.49)
1rWO!

f
F

t
Fc

x-
I Fig. 12.ll. Coulomb damping force for harmonic
x displacement z(t) = i sinwt.
182 12 Internal Damping

where
~ R2
Me =ke ii = ke - ii (12.50)
2
Thus for solid damping
~ 4 2 R2
d e =- - = - ke -
ke
(12.51 )
11" W 11" W

For a flexible coupling, for example a toothed coupling as shown in Fig. 12.8,
reference [92] derives a more accurate relationship, which will be reproduced
using our notation, as follows.
The coupling is loaded by a turning moment M T Thus for axial displace-
ment z(t) and dry friction conditions, the required force is

F(t) = p. i sgn (:i:) (12.52)

where R is the radius of the pitch circle and p. is the coefficient of friction.
From Fig. 12.9, for rotation of the coupling flanges with vibratory velocities
eX and {3, the velocity :i:8 at the point described by angle 5 is given by

(12.53)

where
(12.54)

Assuming the friction force at the teeth is uniformly distributed a.round the
circumference, then for an element of arc this force is,

and this lea.ds to the following components of moment

J J
2~ 2~

Me:< = p. ~: sgn(:i:8)sin5 d5 , M~ = -p. ~: sgn(:i:8) cos 5 d5 . (12.55)


o 0

From equation (12.53), :i:8 is negative for 0 < 5 < 'Y, positive for 'Y < 5 < 11" + 'Y
and again negative for 11" + 'Y < 5 < 211". Hence, performing the integration,
2
Me:< = - P. MT cOS'Y ,
11"

and using equations (12.54),


2 eX 2 {3
Me:< =- p. MT -;::====;;= , M~ =- P. MT -;===== (12.56)
11" "';0.2 + {32 1 1 " " ' ; 0 .2+ {32
12.3 Damping from Assembly Components 183

The moments M", and M", for fixed coordinates are obtained by substituting

{3=tb-ilcp. (12.57)

The moments are thus non-linear in their velocities. The equivalent coefficients
of viscous damping are found using equation (12.49) with
2
Me = - P, MT (12.58)
11"

Thus
~ MT 8
de = - p, ----::: (12.59)
wa 11"2

or, with displacement amplitude $ at radius R,

(12.60)

Expressions (12.59) and (12.60) are very questionable due to uncertainties in


w and ii, that is $/ R. However, they are sufficient to get a first estimate of
stability.
To compare the damping of a toothed coupling with the material damping
of a comparable length of shaft we consider the following example.
A toothed coupling is given, having radius R = 150 mm on a shaft of diameter
d = 150 mm. The turning moment MT gives a torsional stress in the shaft of
T = 50 N/mm2 The value of MT is thus T Wp, where Wp = 1I"d3 /16. This gives
MT = 3.31 . 10 7 Nmm. From equation (12.58) and with p, = 0.10, the Coulomb
damping moment is
Me = 2.10 . 106 Nmm .
For comparison assume a 500 mm length of shaft of 150 mm diameter. It may
be bent at its ends by a harmonic moment, so as to give a central displacement
of amplitude :v = 100 p,m. The amplitude of angular displacement is obtained
from
ii = 8 :v/l = 1.6.10-3
The bending stiffness is
~EI 11" d4
k=- ,with 1=-
I 64
and, by analogy with equation (12.34), it follows that the coefficient of damping
IS
~ k
d i ='1/ -
w
and the amplitude of the damping moment

Mi =d i w ii = '1/ E I ii (12.61)
I
184 12 Internal Damping

With the values E = 2.1 .10 5 N/mm 2 , d = 150 mm,,,, = 0.002

then Mi = 3.34 . 10" N mm and the ratio Me / Mi = 63.

Other values can be assumed, but this example shows that in general, flex-
ible couplings provide much greater damping than do comparable lengths of
shaft. Coulomb damping gives a closer approximation to material damping than
does viscous damping but in practice they are not very different. Microslip the-
ory may offer an improvement. In reference [92] Lund describes its application
to rotors, and in his summary he gives some further points which however will
not be considered here.

12.4 Conclusion and an Example

In Sects. 12.1 to 12.3 the fundamental features relating to the effect of internal
damping on rotors were described. Further details can be found in an extensive
literature dating back to 1924. Of these the following des~rve mention.
Tondl [93] gives a detailed description of the basic principles and Kel-
lenberger [94] considers cylindrical shafts with internal and external damping.
Gunter and Trumpler [95] investigate the influence of anisotropic bearings on a
Jeffcott rotor. They ascertain that stability can be improved by anisotropy. Shi-
raki and Umemura [96] report practical experience and describe cases of twin-
shaft machines with a toothed coupling, for which instability was observed. By
further calculations and measurements on a model rotor they made the impor-
tant discovery that the kind of natural mode shape is significant in determining
instability. Such instability arises for each natural frequency for which vibration
across the coupling is especially pronounced. Pedersen [97] gives an extension
to the theory by investigating a cylindrical shaft on oil-film bearings. Non-
linear properties of a cylindrical shaft were investigated by Muszynska [98] who
presents results using many clear diagrams.
By referring to [99, 100, 101] Nataraj et al. showed in ref. [102] the influ-
ence of a coupling with Coulomb damping on the dynamics of a rotor. Their
model consisted of a stiff rotor supported on stiff bearings, and having two
halves divided by an elastic coupling. A force relationship analogous to that for
moments M<p and M.p in equations (12.56) was assumed at the coupling. In addi-
tion their model was supported by an anisotropic spring and damper. A closed
form solution was given for isotropic bearings, in which case the model was
essentially like a Jeffcott rotor with internal viscous damping (see Sect. 12.1).
For anisotropic bearings, as well as elliptical-type natural motions there were
further vibrations of higher frequencies, which, however, were shown to have
much less significance.
12.4 Conclusion and an Example 185

Huhlmann and Luzi [103] report a very interesting case, whose main essen-
tials will be given, followed by the results of a complete calculation in example
12/1. This publication concerns a 21 MW gas turbine having an operating speed
of 7729 rev/min and driving a generator at 1800 rev/min. Each of the connect-
ing shafts has two toothed couplings (Figs. 12.12, 12.13). At low load, that is
between 2 and 6.4 MW, severe vibration was encountered, especially at the gear-
box. Analysis showed that mainly sub synchronous vibration was present, whose
frequency increased with load. This frequency ranged from 86 Hz at 2 MW to
120 Hz at 6 MW. Measurements and calculations showed further that the part
ofthe shaft between toothed coupling II and the gearing (Fig. 12.13) described a
circular conical motion. The dependence of frequency on load could be explained
by the fact that, with increasing load and thus torque, this part of the shaft was

Gasturbine Planetary Generator


gear

Fig. 12.12. Gasturbine and generator, from [103]

Gasturbine Intermediate shaft End section


I I I !
~ Tooth,d "0,[;09 1 Tooth,d "0,[;09 IT
I I I I I I
. . ~ ----~--------~
I I I I I I
I I I I
I I I I

t:-
Bearing A Bearing B Gear
990 1810
I"
1150 1
1
-1-
3330 ~

1
258 kg 768kg 200kg
Fig. 12.13. Mathematical model (gasturbine, toothed couplings, intermediate shaft,
end section). All dimensions are in mm.
186 12 Internal Damping

under increasing strain from the planet gear and thus was held progressively
more stiffly. Since the vibrations at 6.5 MW, that is for correspondingly higher
natural frequency, fall off to a permissible level, the remedy became obvious:
The significant natural frequency must be sufficiently high. Thus the shaft was
shortened and made lighter. As a result the natural frequency was increased to
146 Hz at only 2 MW. 146 Hz was some 13 % above the running speed. The
machine, as a result, ran in a stable state for all load conditions.
Although it was not mentioned expressly in ref. [103], one may assume that
a case of self-excitation was evident here, as a result of friction damping in the
coupling. To investigate this further, additional calculations need to be carried
out, whose results are given in example 12/1.

Example 12/1 1 For the case described in reference [103], we shall calculate
natural frequencies and unbalance vibration amplitudes and investigate the ef-
fect on machine stability of friction in the couplings.
This procedure was carried out using program MADYN (Chap. 25). It is
sufficient to consider only the high speed shaft between the gas turbine and the
gearbox, using Fig. 12.13. This shaft consists of the turbine rotor, the interme-
diate shaft and the end section, comprising the hub of coupling II and the sun
gear of the gearbox. The toothed couplings were first assumed to be flexible and
without friction. The gas turbine is supported in oil-film bearings A and Band
the end section by the planet gear. For the bearings the coefficients given in
Table 12.1 were assumed. As a simplification their dependence on rotor speed
was neglected.

Table 12.1. Stiffness and damping coefficients

10 7 N/m 104 Ns/m


kn k12 k21 k22 dn d12 d21 d22
Bearing A 6.07 5.11 -2.80 42.7 5.60 0.80 -0.70 25.6
Bearing B 3.20 2.13 -1.42 10.1 3.80 -0.60 -0.30 8.8

The constraint in the end section in the planet gear depends upon the torque
transmitted. This section was simulated by isotropic springs at 150 mm spacing
and kl and k2 were each assumed constant at 20.10 7 N/m. The model has 17
elements with 64 degrees of freedom and a total mass of 1226 kg.
The eigenvalues and eigenvectors were calculated using the so-called Hes-
senberg procedure. At the natural frequencies the individual elements of the
model move in elliptical paths, and the instantaneous mode shapes vary con-
IThe writer thanks the Sulzer Company, Winterthur, and BHS-Voith, Sonthofen, for the
essential data and for their permission to publish these findings.
12.4 Conclusion and an Example 187

sider ably from one to the other as a result of the oil-film bearing coefficients
(for details see Sect. 17.1). At the operating speed the first seven mode shapes
projected onto one plane are displayed in Fig. 12.14. On the left are the natural
frequencies and damping ratios. Six natural frequencies lie below the operating
speed of 7729 rpm/60 = 128.8 rev Is. The mode shapes and damping ratios show
that the first three and the fifth natural frequency are strongly influenced by
the oil-film bearings. On the other hand the fourth and sixth natural frequen-
cies are almost solely determined by the intermediate shaft. Bearing damping
here has a barely noticeable effect as the low values of 0.001 for D4 and D6
show. On the right in Fig. 12.14 the type of natural vibration is characterised.
Hor. or vert. signifies that the natural motion is mainly horizontal or vertical
in character. At the fourth and sixth natural frequencies the elements move in
approximately circular orbits, the fourth in a backward whirl, the sixth in a
forward whirl.
To calculate unbalance vibration three typical cases were assumed.

Case I Single unbalance at the left-hand end of the shaft (Fig. 12.13), with
U = 1.15 kg mm corresponding to a mass of 115 kg at a 10 /Lm eccentricity.

Case II Pair of unbalances with U = 3.68 kg mm at a distance of 140 mm to the


right of bearing A and U = -3.68 kg mm, 130 mm to the left of bearing
B. 3.68 kg mm represents a mass of 368 kg at an eccentricity of 10 /Lm.

Case III Single unbalance at toothed coupling II with U = 0.575 kg mm, cor-
responding to a mass of 57.5 kg at an eccentricity of 10 /Lm.

Nr D H$ili--~~I'!-t-----<m Kind of vibration


cps I A BI II Gear hor. vert. Bw. Fw.

------ -----
47.1 0.125
:----- - x

2 48.5 0.139 x
V

3 75.2 0.170 r---.. x

-----
4 96.9 0.001 x
/
0.134
r---- x
5 102.8

-------
6 108.2 0.001 x
/

7 168.1 0.054 r---


// r--------- x

Fig. 12.14. Example 12/1. Natural frequencies, damping ratios and mode shapes for
n = 7729 rpm.
188 12 Internal Damping

The vibration amplitudes of the rotor at bearings A and B and at the


toothed coupling II were calculated for the horizontal and vertical directions at
speeds up to 12000 rpm. The results are shown in Fig. 12.15. The behaviour
of the toothed coupling I I is especially noteworthy. It shows a main resonance
at the sixth natural frequency, with amplitudes which are some 10 to 30 times
greater than the resonance amplitudes at the bearings. In case III, with an
unbalance at the coupling, there is a resonance only at /6, corresponding to
stimulation in a forward whirl. For cases I and II resonances are evident at the
bearing positions, whose varying strengths can be explained by the mode shapes
of Fig. 12.14.
If some external damping had been assumed in the area of the intermediate
shaft, then the resonances at the toothed coupling II would, of course, have been
weaker. We can conclude, however, from this calculation that this coupling is
especially sensitive to unbalance excitation.
From the calculations carried out up to now the two toothed couplings
were assumed to be friction free links. To investigate stability, calculations were
carried out with the assumption of viscous da~ping at both couplings using
a variety of different values for the coefficients d i . As a result, a very variable
behaviour of the eigenvalues was in evidence, which will not be discussed h;:re.
Checks against the actual behaviour of the machine finally gave the value d;=
~03 Nms. To appreciate the scale of this damping, we compare the value of
1:. i n, which has the dimension of torsional stiffness with the torsional stiffness
k = EI/l of the intermediate shaft. With diameters 232 mm and 213 mm and
length 1320 mm, k= 6.52.10 6 Nm. With n = 809 rad/s as the nominal speed,
di n is by comparison 0.8~9. 106 Nm, that is 0.12 k.
Due to the terms d i n in equations (12.42) the eigenvalues and the
stability depend on the rotor speed. The eigenvalues were thus calculated in
the speed range in question. Results were obtained which suggested that the
sixth natural frequency would be unstable. Further it was found that no other
natural frequency would be unstable.
Fig. 12.16 shows the behaviour of natural frequency /6 and damping ratio
D6 in the region 5000 to 10,000 rpm. At 6420 rpm (107 rev / s) the damping
ratio becomes negative and one encounters the stability borderline. At this speed
/6 = 107 cycles/s, equal to the rotor speed. This equality means that no external
damping prevails, as can be seen from equation (12.44). The only assumed
external damping, namely that of the oil-film bearings does not manifest itself,
as has been already ascertained.
This example shows that with today's knowledge, added experience is
necessary to ascertain where the stability borderline lies. The following results
of this chapter should however be sufficient to ensure the avoidance of instability
due to internal damping by suitable design procedures:
Case Bearing A Bearing B Toothed coupling IT

40 40 ~ 600

I 20 20~ I /\ 300
1 .:f.."-
0 00 i....,..-l---~ i 0
E
::1. 1\ I III I 1\ I III I 1\ I III I
<II 1 3 5 7 1 3 5 7 1 3 5 7 1)
"'0
::J
.... 40 40 J 1\ 600
0.
E II 20 300
20
0 0 0 ......
t-.:)
1\ I III I II I III I 1\ I III I
~
1 3 5 7 1 3 5 7 1 3 5 7 1)
0
0
40 40 600 ='n
~
m 20 20 300
-'"o
='
~
0 0 0 ='r:;;..
0 5 nap 10 0 5 nap 10 0 5 nap 10
~
='
t?=j
Speed (1000 rpm) ><
1) Natural frequencies f,+f7 at nop =7729 rpm ~
S
"C
Fig. 12.15. Example 12/1. Unbalance vibrations. Natural frequencies It ... h for ('I)

nap = 7729 rpm. - horizontal, - - - vertical.


-
......
00
<I:l
190 12 Internal Damping

120

cps

110

100

L I
5000
I, I
10000 rpm
n-

t 0.04
D6
D6
0
5000 10000 rpm
-0.04 % n-
nth
-0.08

Fig. 12.16. Example 12/1. Natural frequency /6 and damping ratio D6

Internal damping can have both a damping effect and destabilising effect on
a rotor. The destabilising influence depends upon rotor speed, as a result
of which instability ensues at a certain speed.

Material damping is negligible in comparison with damping within the var-


ious machine elements.

The natural frequency at which the location of the internal damping vibrates
relatively strongly in relation to other rotor positions, and also in a forward
whirl is significant in deciding the stability borderline.

External damping delays instability. However, its value is mostly unknown


and so the significant natural frequency should be made greater than the
maximum operating speed.
13 Non-circular Shafts

The vibration of a shaft depends amongst other factors on its stiffness. This is
defined as the quotient of a force applied to the shaft divided by the resulting
deflection in the direction of the force. If this stiffness is independent of the
direction of the force in the plane in question, then the shaft is regarded as
circular, otherwise it is non-circular.
A shaft is accordingly defined as circular when the second moment of area
of its cross-section about any axis through the centre of area is invariable.
Examples of such cross-sections are shown in Fig. 13.1. It can be 'seen that
shaft sections defined as circular do not have to be geometrically circular.
A shaft is defined as non-circular when the second moment of area of its
cross-section over the whole shaft or over a part of the shaft is dependent upon
the position of the reference axis through the centre of area. This is the case
when the principal second moments of area h, III are different. The radii of
gyration then describe an ellipse as shown in Fig. 13.2, while Fig. 13.3 gives
examples of non-circular cross-sections found in practice.
For h = III, that is for circular cross-sections, the momental ellipse becomes
a circle, whence it is evident that, for this case, the second moment of area is not
dependent on the position of the reference axis. A very important practical case
of a non-circular shaft is the rotor of a two-pole turbogenerator. It has between
the poles longitudinal slots to accomodate the windings. The cross-section of
the socalled active part essentially has the form (e) shown in Fig. 13.3.
The most important vibrational features of a non-circular shaft can be
discussed in the context of a non-circular leffcott rotor and will be considered
in Sect. 13.1. A summary will then be given in Sect. 13.2.

a b c d e
Fig. 13.1. Cross-sections defined as circular.
192 13 Non-circular Shafts

Fig. 13.2. Momental ellipse


Second moments of area h = f :z:2dA, III f v 2 dA. Radius of
gyration i .. .J
= 1../A.

a b c d e
Fig. 13.3. Non-circular cross-sections.

13.1 Non-circular Jeffcott Rotor

Prandtl [104J reported on the behaviour of a leffcott rotor with non-circular


cross-section in 1918. A fundamental investigation was also carried out by Smith
in ref. [105J. There followed further publications of which ref. [106J by Taylor
and Chap. II in ref. [93J by Tondl will be referred to. The following explanation
is taken from ref. [32J.
The leffcott rotor of Fig. 13.4 has a non-circular cross-section with principal
second moments of area hand hI. The displacements in a stationary coordinate
system are :z: and y. In coordinates rotating with shaft angular velocity il the
corresponding displacements are u and v. It is assumed that the displacements
u and v are parallel to the principal axes. The centre of gravity S of the disc
is eccentric to the axis of rotation, its position being defined by eccentricity e
and angle {3.
The stiffnesses kx, ky depend upon the angle ilt while k'J! kv remain con-
stant. Hence it is simpler to set up equations of motion in u and v, rather than
in :z: and y. We first derive equations in u and v for a circular shaft and then
extend them to the case of a non-circular shaft.
13.1 Non-circular Jeffcott Rotor 193
y

Fig. 13.4. Coordinates in a non-circular shaft.

For a circular shaft with k", = ky = k the following equations apply

mz + di: + kz = m e {}2 cos ({}t + (3)


(13.1)
my + diJ + ky = m e {}2 sin ({}t + (3) - G ,

and with z = z + jy
mz + di + kz = m e (}2 ei(!1t+~) - j G (13.2)

With
(13.3)
we have
z r ei (!1t+a) = ( ei !1t

i (i{}(+ C) ej !1t (13.4)


z (_{}2( + j 2{}C + () e j !1t

From equation (13.2) the following equation results

In equation (13.5) the real part and the imaginery part must each be satisfied
and from these can be obtained equations of force equilibrium in u and v di-
rections. If expressions kuu and kvv are inserted for the components of shaft
restoring force in place of ku and kv then the equations of motion for a non-
circular shaft are obtained as follows:

mii + du - 2m{}iJ + (ku - m{}2) u - d{}v = me {}2 cos{3 - G sin {}t


mv + 2m{}u + diJ + d{}u + (k" - m{}2) v = me {}2 sin{3 - G cos {}t
(13.6)
194 13 Non-circular Shafts

with EIIl
and kv =48 P (13.7)

for a constant shaft cross-section.


As usual, consider first natural vibrations and then forced vibrations arising
from unbalance and rotor weight.
For natural vibrations we start by putting
'/J =V eAt (13.8)

for
d = 0 and WI = fFi WII =
r;;:
V;;, (13.9)

in the following algebraic equation for the eigenvalues

,2 -
Al,2 - - _12 ( w I2 + w II
2 + 2 n2)
U -41 (WI2 + wI2I + 2n2) 2 -
U (wI2 - n2) (W I2I
U _ n2)
U

(13.10)
The solutions Ak, Ai; = ak jvk (k = 1,2) depend upon rotor speed n. In the
region 0 < n < WI and WII < n < 00 we obtain the solutions

At, A~ = jVl , A2, A; = jv2 (13.11)

and in the region WI < n < WII the solutions

(13.12)

Thus in the first two regions harmonic natural vibrations offrequencies VI and V2
exist and in the n region between WI and WII a natural vibration with frequency
V2 exists as well as another motion, increasing and decreasing by exponent al t.
For no damping this region is thus unstable.
These natural motions take place in a system rotating at angular velocity
n. With transformation (13.4) that is with
:I: U cos nt - v sin nt
y u sin nt + v cos nt (13.13)
the natural motions in fixed coordinates are obtained. This results in harmonic
natural vibrations of frequencies n + Vk and n - Vk, denoted by WI ... W4, with
exponentially increasing and decreasing motions in the unstable region. Fig. 13.5
shows as an example the natural frequencies of a rotor with WII = 1.8 WI, that
is with III = 3.24 h, which represents a shaft with extreme non-circularity.
External damping reduces the unstable region. The influence of damping on
the natural frequencies is minimal and we shall not discuss thus feature further.
For A = jv one finds from reference [32) the following equations for the stability
borderlines

-= (13.14)
Wo
13.1 Non-circular Jeffcott Rotor 195

<f4
W,

Q-

Fig. 13.5. Natural frequencies of a non-circular Jeffcott rotor shaft with WII = 1.8 WI
in a fixed coordinate system.

with

Wo "21 (2 2)
WI + WII (13.15)

d
Do = (13.16)
2 mwo

Using equation (13.14), the stability chart shown in Fig. 13.6 is obtained. This
shows, that for each value of damping ratio Do there exists a certain value of
non-circularity at which instability becomes possible. Further it can be con-
cluded from Fig. 13.6 that, for non-circularity under about 10 %, instability
has little practical meaning. On the one hand an operating speed in the re-
gion of WI and WII would be avoided, and on the other hand one can always
run through the existing unstable region reasonably quickly, so as to avoid any
vibration problem.
Unbalance is introduced in the terms me[J2cosf3 and me[J2sinf3 in equa-
tions (13.6). Accordingly, an unbalance gives rise to a constant displacement
with time of magnitude r = ";u 2 + v 2 in a rotating coordinate system. In a
stationary coordinate system the displacement of the disc centre describes a
circular motion with angular velocity [J. The radius r of this circle depends
on the one hand on [J and on the other on angle f3 of the eccentricity. The
resonance curve r([J) shows for sufficiently small damping two general peaks in
the vicinities of WI and WII. For greater damping only one peak is displayed.
Fig. 13.7 shows resonance curves relating to a rotor with WII = 1.1 WI and
Do = 0.1 for f3 = 0,45 and 90. For these parameters only one peak is in evi-
196 13 Non-circular Shafts

t Unstable

1.05

1.0
o 0.95 1.0 1.05
.R_
Wo
Fig. 13.6. Stability chart for non-circular leffcott rotor shaft.

dence and one can in general say that the frequency at which this peak occurs
is about WI(WII), if the eccentricity lies near principal axis l(II). The height of
the peak (r/e) has a value in this example of 7.4 to 9.5 and is thus higher than
1/2Do = 5.0. In other examples it was also found that (r/e)max > 1/2Do and
so in general one can speak of the non-circularity effect acting to reduce the
influence of damping.
Rotor weight is represented in equations (13.6) by the terms -G sin nt and
-G cos nt. As a particular solution we hence find harmonic motions u( t) and
v( t) of frequency n, giving an elliptical orbit in a rotating coordinate system.
Transformation to a fixed coordinate system gives the displacements

:v(t) =:Vo + rg (n)cos[2nt - c,o(n)]


(13.17)
y(t) = Yo + rg (n)sin[2Ot - c,o(n)] ,

L 10
e
45
0
90
5

o
Q-

Fig.13.T. Resonance curves for anon-circular shaft withwII = 1.1 WI and Do = 0.1.
13.1 Non-circular Jefcott Rotor 197

that is a circular orbit of the disc centre W at frequency 2{} with its centre
C(:l:o, Yo) as shown in Fig. 13.8.
Along any radius line through C the disc centre moves harmonically with
frequency 2{} and amplitude T g ({}). We speak of 2{}-vibration or 2n-vibration,
where n = {} /27r is the speed.

2Q

Fig. 13.S. leffcott rotor with non-circular shaft. Orbit of disc centre resulting from
gravity excitation (2n-vibration).

The radius Tg of the circular orbit has the value

0.51YI - YIII , for {} = 0

with G G
YI= - - YII = - - (13.18)
kI klI
as the maximum static displacements resulting from rotor weight. The radius
reaches a maximum in the vicinity of
n _ WI WII Wo
Ucr,g - 2 + 2 Wo < -2 (13.19)
WI WII
and approaches zero as {} approaches infinity (Fig. 13.9).
The value of Tg,max can be concluded from Fig. 13.10. It has an order of
magnitude related to the average static displacement

(13.20)
The centre C(:co, yo) of the circular orbit describes an approximate circle, de-
pendent upon {}, as shown in Fig. 13.11. For {} : Wo, C lies near to the point
(0,11) and for {} - - ? 00 near to (0, Yoo).
For {}cr,g the lateral displacement is a maximum. The value of :CO,max can be
obtained from Fig. 13.12.
198 13 Non-circular Shafts

1'1-
Fig. 13.9. leffcott rotor with non-circular shaft. Radius of circular orbit due to gra-
vity excitation (2n-vibration).

Do
0.02 0.05 0.1
rg.max
IYI
0.1

1.01 1.02 1.03


WI
w;--
Fig. 13.10. Radius Tg,max of 2n-vibration.

Yo

Xo,mnx

Xo

Yn Yoo
Q=Q".g
if

YI

Fig. 13.11. Position of centre C and its dependence on angular velocity.


13.2 Supplementary Comments 199

t 0.02 0.05 0.1


Xo.max
Iyl 0.1

O~~-r-r-.-.--.-.-'-

--
1.0 1.1
Wn
WI

Fig. 13.12. Maximum lateral displacement of point C

13.2 Supplementary Comments

The results of Sect. 13.1 are essentially true for real rotors with non-circular
cross-sections although they describe their vibrations in an incomplete way.
Amongst other things, the behaviour at higher natural frequencies and for
anisotropic bearings is of interest.
The first can be considered using the results for a cylindrical, non-circular
shaft with a uniformly distributed mass. Such a shaft has a similar behaviour
to a Jeffcott rotor as has been demonstrated by Kellenberger [107]. This applies
to fundamental as well as higher harmonic vibrations. A special case arises
for vibrations excited by gravity. Only symmetric mode shapes are excited.
Resonances of the 2il-vibration variety thus occur in the regions of about Wl/2,
W3/2, W5/2, etc. in which WI, W3, W5, etc. are average values of the pair of natural
frequencies in question. A pair of natural frequencies with identical mode shape
arise due to non-circularity.
For the non-circular shaft in anisotropic bearings, even in a rotating-
coordinate system, the equations of motion feature periodic coefficients. So-
lutions can be found in, for example, references [105, 108, 109, 110].
Due to anisotropy, as is well-known, two natural frequencies occur in place
of each one for isotropic bearings. Investigations show that, in this case, three
unstable regions are possible, one each at the two afore-mentioned frequencies,
denoted by Wa and Wb and between them another at Wm (Fig. 13.13). Due to
damping, these regions are reduced in a similar way to that shown in Fig. 13.6.
Further unstable regions arise theoretically at wm/n (n = 1,2,3 ... ). How-
ever, they disappear for relatively small damping and thus are of little practical
significance.
Finally the following comments can be made in respect of two-pole rotors of
turbogenerators. In order to avoid exciting 2n-vibrations, the designer should
200 13 Non-circular Shafts

11-
Fig. 13.13. Stability chart for non-circular rotor with anisotropic flexible bearings

attempt to remove as much as possible any non-circularity due to the winding


slots. To do this, either several cross-slots or additional longitudinal slots should
be provided over the length of the poles. As it is not easy to estimate the stiffness
- it is determined by the stiffnesses of the rotor, the windings and the keys and
their keyways - there always remains a certain residual non-circularity. This is
however, almost always so small that the existing damping does not allow the
above mentioned instabilities to occur.
Part II

Rotors with Several Bearings


In Part I the model of the Jeffcott rotor was used, from which the most
important phenomena encountered in rotordynamics were presented. A know-
ledge of these phenomena is fundamental to an understanding of the behaviour
of complex models, which correspond better than does the Jeffcott rotor to the
real rotors of turbomachines. In contrast to the Jeffcott rotor, models will be
considered here, which have variable stiffness and mass along their lengths and
have more than two bearings. Furthermore, the bearing pedestals will likewise
be considered as vibrating systems.
14 Computer Model

Figures 14.1 to 14.3 show typical examples of rotor systems whose characteristics
will be investigated in this part. These systems consist of rotor, bearings and
bearing pedestals and models will be devised which will enable the bending
vibrations of real rotors to be obtained sufficiently accurately, using reasonable
computing time.

Fig. 14.1. Turbine-generator line out with four bearings.

Fig. 14.2. Turbine-generator line out with six bearings.

Fig. 14.3. Pump-turbine unit


206 14 Computermodel

We begin with the rotor, which will be divided into sections. Each section
will again be divided into a number of elements. The points on the element
boundaries lying on the rotor axis will be called nodes (Fig. 14.4). The dis-
placements :l:ll :1:2 at a given node and the angular displacements (rotations)
<P3, <P4 in the planes in question represent the motion of the rotor at these
positions. The positions and number of nodes should be so chosen that the
mass and stiffness properties of the rotor can be readily obtained. Using many
nodes will increase the accuracy, but will demand more computing time. It is
thus necessary to choose a sensible number of nodes, which give an acceptable
compromise.

Element

k-1 k k+1

Fig. 14.4. Rotor model. Nodes and displacements.

Each bearing pedestal will be replaced by a system such as that shown in


Fig. 14.5. This consists of a point mass and spring-damper elements, which
are arranged in mutually perpendicular positions (here horizontal and vertical).
The model is a simple planar vibrator, having displacements :1:3 and :1:4.

Fig. 14.5. Model of a bearing pedestal.


15 Influence Coefficients

The finite-element method leads to equations of motion, which describe the


equilibrium of the forces and moments which act on the system. Each elastic
force and moment appears as the product of an influence coefficient and a
displacement, either translational or angular (rotational). In this way we can
find all the terms on the left-hand side of the equations of motion. This means
that the determination of the influence coefficients is very important. This will
be explained in the following by a simple example and then a general case will
be taken.
Consider a beam as shown in Fig. 15.1. The given forces and moments act
at nodes a and b. These forces and moments will be called collectively "forces"
for convenience and will give rise to translations and rotations (collectively
called displacements). The system is linear elastic and so these displacements
are proportional to the forces.
Thus for the displacements, we have:

:1:1 hl1Fl + h12M2 + h13F3 + h14M4


'(J2 h21Fl + h22M2 + h23 F3 + h24M4
(15.1 )
:1:3 h3l F l + h32M2 + h33F3 + h34M4

'(J4 = h4lFl + h42M2 + h43F3 + h44 M 4


where hik (i, k = 1 ... 4) are the corresponding constants of proportionality, the
so-called displacement influence coefficients.

a
Fig. 15.1. Example explaining influence coefficients.
208 15 Influence Coefficients

Correspondingly for the forces

FI knz i + k I2 1(J2 + k13 Z 3 + k I4 1(J4


M2 = k2I Z I + k 22 1(J2 + k 23Z 3 + k 24 1(J4
(15.2)
F3 = k3l Z I + k 32 1{J2 + k33 Z 3 + k 34 1(J4
M4 = k4I Z I + k 421(J2 + k43 Z 3 + k 44 1{J4

where kik are the force influence coefficients or stiffnesses.


In matrix form these equations become

x= H r and r= K x (15.3)

where x = (Zb 1{J2, Z3, 1{J4)T , r= (Fb M2 , F3 , M4)T


H = (h ik ) K = (kik) = H- I .
The force influence coefficients or stiffnesses can be found from the displace-
ment influence coefficients by inversion.
From Maxwell's reciprocal theorem hki = hik and kki = k ik . The H-matrix
and the K-matrix are symmetric, if the forces and displacements are suitably
arranged, as in the foregoing example.
It will be observed that each influence coefficient has one of three different
dimensions. For the force influence coefficients the following units (in the SI-
system) are featured:

JV/Tin for kll' k l3 , k31' k33


JV Tin for k22' k24' k42' k44
JV for the remaining eight kik

The corresponding units for the displacement influence coefficients are Tin / JV,
1/JVTin and 1/JV. From equations (15.2) the following can be concluded:

A force influence coefficient kik is the force Fi which arises when the
displacement Zk has a value of unity and all remaining displacements
are zero.

In our example the deformations shown in Fig. 15.2 are accordingly prerequisites
for the calculation of k ik .
The definitions of the force influence coefficients used for this beam are also
valid for two- and three-dimensional linear elastic structures. They are also
valid for a random pair of nodes or for a single node, and for a random pair of
perpendicular coordinates at a node.
15 Influence Coefficients 209

a b
Fig. 15.2. Deformations leading to the calculation of kik.
16 Equation of Motion

For any model such as described in Chap. 14, it is required to calculate its
bending vibrations. The displacements will be regarded as small and all relations
will be assumed linear. With the aid of the transfer-matrix method of Myklestad
[6] and Prohl [7] the models considered here can be solved with reasonable
efficiency. This method has retained its usefulness up to the present time and
has been extended to rotor-foundation systems (Wang, Lund [111,112]). Details
of its application are described in numerous publications, of which only the book
on basics and cited in reference [113] will be mentioned. The transfer matrix
method has proved itself useful for single drive-line models. Possible numerical
difficulties have since become apparent and we know how to control these [114].
Such experience, however, has not been acquired for extensive models, such as,
for example, the rotor-foundation model discussed in Part III. As an alternative
the finite element method has proved to be very useful and will be applied in
the following.
In rotordynamics the finite elements are relatively large. We could thus
rather speak of the method of macro-elements. In the study of the statics of
beams, the method of displacements has been in use for a long time. This name
would also be very pertinent here, as in rotordynamics the finite element method
represents an extension of the former method to dynamics. We shall however
stay with the usual name and consider first the static case. Extension to the
dynamic case should then be easily understood.

16.1 Statics

Take as an example a stationary rotor (!1 = 0) as shown in Fig. 16.1. It has


a variable cross-section, three elastic bearings and is statically loaded at each
of three positions by a force and a moment. We require the displacements at
nodes a to m and the rotations of the cross-sections in question as a result of
the loading.
To effect a solution the required displacements and rotations (Fig. 16.2) are
placed in the displacement vector
212 16 Equation of Motion

f f

Q=O

1 I I I 11111111
abc d efgh kim
Fig. 16.1. Rotor model with static loading.

M6 M8 M18
,.......,....... ,.......

tFs tF7 tF,7


412 416 ,.......
418 4112 4118 4124
,....... ,....... ,....... ,....... ,.......

tX1
0-. ..
txs tX7 tX11
.. --0- .
t X17
.. -0
t X 23
~.
--0-
a c d m
Fig. 16.2. Coordinates of the model.

and the loadings in the load vector

f = (0, ... 0, Fs, M6 , Fr, Ms, 0, ... 0, F17 , MIs, 0, ... of


From equations (15.3), using the stiffness matrix K, which is still to be found,
we have
Kx=f. (16.1)
By solving for x, the required displacements and rotations are thus obtained.
To obtain the stiffness matrix one first considers a single beam element
as shown in Fig. 16.3 and having parameters 117 E, II, the displacements
Xl, <P2, X~, <p~ and the boundary forces Fh M 2, F~, M~.
16.1 Statics 213

2 4' 4" 6
~ ~ ~ ~

11 h h 15
+-+
1,.1,
+--+ 11.11

2 4 6
~ ~ ~

11 b 15
+-f--} Fig. 16.3. Connecting two beam elements.

Using equations (15.3) one obtains the relationship

kll k12 k~3 k~4

k21 k22 k~3 k~4 1f'2


(16.2)
k;l k;2 k;3 k;4
k~l k~2 k~3 k~4
where kll , k 12 , ... are the force influence coefficients or stiffnesses defined in
Chap. 15.
For a second element one correspondingly obtains

F"3 k"35 k"36 X"3

M"4 k~3 k~4 k~5 k~6


(16.3)
k~3 k~4 k55 k56
k~3 k~4 k65 k66
Bringing together elements 1 and 2 as in Fig. 16.3 we obtain

(16.4)
These are the conditions for compatibility. Also the sum of the internal forces
is equal to the external force and so

(16.5)
These are the conditions for equilibrium of forces.
214 16 Equation of Motion

Hence, for the combined element one obtains the relationship

FI kll k12 k~3 k~4 Xj

M2 k21 k22 k~3 k~4 'f'2

F3 k~1 k~2 k33 k34 kif35 kif36 X3


= (16.6)
M4 k~1 k~2 k43 k44
kif kif
45 46 'f'4

Fs kif kif k55 k56 X5


53 54

M6 kif kif k65 k66


63 64 'f'6

where

[ k" k", 1~ [ k;, +kg, k;.+ k~'l (16.7)


k43 k44 k~3 + k~3 k~4 + k~4
The combination of the two elements thus appears to be a simple procedure,
in that the corresponding coefficients at the common node are added. This is
in general valid, as will be shown in further examples. It is assumed that the
coordinates of the shared nodes are parallel to each other.
In a similar way the K-matrix of all eleven beam elements can be composed
for the rotor model of Fig. 16.1. The K-matrix of the whole rotor thus exhibits
the following band structure

4 21 24

K= (16.8)

21

24

The elements in the double- hatched portions consist of the sums of coefficients
described above.
The constituents of the K-matrix of each beam element may be obtained
from the theory of beam bending as follows,
16.1 Statics 215

al a2 -aj a2

a2 a3 -a2 a4
K= (16.9)
-at -a2 al -a2

a2 a4 -a2 a3

where
12 EI 6 EI
al = 1 +e P a2 = 1+e P
(16.10)
4+e EI 2-e EI 12EI
a3=-- - a4=-- -
1 +e 1 1 +e 1 e = GK,AP
and K, is the shear factor. This depends on the form of the cross-section. K, = 0.89
for a circle and 0.53 for a thin circular ring. The shear factor relates to the shear
deformation and reduces the stiffnesses.
Up to now we have not considered the bearings. These consist of springs
at nodes a, f and m having stiffnesses ka, k f and k m . They increase the local
stiffness of the shaft at the nodes in question, which is acknowledged by adding
ka, k f and k m to kl,1) k 11 ,11 and k 23 ,23, respectively. Hence the complete K-
matrix of the model shown in Fig. 16.1 is obtained.
For rigid bearings, that is for ka, kj, k m = 00, XI, X11 and X23 = O. There are
thus three fewer unknowns in equation (16.1) and the K-matrix of this model
results from the previous matrix by deleting the columns and rows designated
1, 11 and 23. In the same way a restriction in rotation can be allowed for by
deleting appropriate columns and rows. If, for example, the shaft at bearing a
is rigidly held, then, as well as the first column and row, the second column and
row would also be deleted.
Consideration of rigid bearings shows that one cannot simply leave out
node displacements in which one is not interested. This is only permissible if
the displacements in question are always zero. One can, on the other hand,
eliminate displacements using the following method, known as static reduction.
If one wishes in our example to determine only the displacements XI) X3, . X23
and to eliminate the rotations i{J2, i{J4, i{J24, then the system of equations
(16.1) is arranged so that first the displacements then the rotations appear.
One can then write

(16.11)

If there is no moment loading, then f M = 0 and so

K",x Xx + K",,,, x'" =0


216 16 Equation of Motion

from which
XI" = - K;~ Kcpx Xx

Hence from equations (16.11) the reduced equation becomes,

K red Xx = fF (16.12)
where
K red = Kxx - Kxcp K;~ Kcpx (16.13)
The method consists in arranging the system of equations in preferred coor-
dinates and adjacent coordinates and in eliminating the adjacent coordinates.
The method is exact and all coordinates at which no external force or moment
exists can be eliminated. The method can also be used to calculate the K-
matrix of a superelement made up of several beam elements. Super elements
are suitable for rotors having many changes of diameter.
Consider, for example, Fig. 16.4, which shows a superelement consisting of
three parts. With
Xa = (:I:b CP2, :I:7, cps)T fa = (FI, M 2 , F7 , Msf

Xb = (:I:3, CP4, :I:s, C(6)T fb = (0, ... of


we have

and from the above equations

K red Xa = fa (16.14)

where
(16.15)

2 6 8
,--... ,--... ,--...

t1 h t5 h
+-f--f--t-
2 8
,--... ,--...

t1 t7
+-={---E--t- Fig. 16.4. Superelement
16.2 Dynamics 217

16.2 Dynamics

In the section on statics it was assumed that the rotor was stationary and that
the loads acted so slowly that the velocities and accelerations had no influence on
the results. In this section on dynamics the rotor will be assumed to rotate about
its axis and to execute bending vibrations. This brings forces and moments into
play, which depend on the accelerations and velocities of vibration and on the
rotor speed. In order to use the finite element method the corresponding forces
and moments at the element nodes are required.

16.2.1 Rigid Masses

As rigid masses, we consider here lumped masses, rigid discs and rigid rotor
bodies.
We begin with a lumped mass m having moments of inertia Id = Ip = O.
To explain its inertia effect consider the model shown in Fig. 16.5. It consists
of two massless beam elements and has at node b a mass m3. To simplify the
treatment rotations as well as displacements are designated as :I: and moments
as well as forces as F.
The model is first loaded statically with force F3 and moment F4 Then for
node b,
F3 - k31 :1:1 - k32 :1:2 - - k36 :1:6 = 0
(16.16)
F4 - k41 :1:1 - k42 :1:2 - - k46 :1:6 = 0

For dynamic loads F3(t) and F4(t) and using the same summations of forces,

6
m3 Z3 F3(t) - Lk3r:l:r
r=l
6
(16.17)
o F4(t) - I)4r :l:r
r=l

X2 X4. F4 X6
~ ~ ,--....

tXI tx3,F3 txs

a c Fig. 16.5. Model to explain the mass matrix.


218 16 Equation of Motion

The equations for nodes a and c remain the same as for static loading.
With the acceleration vector
"
X =
(.... .. }T
3:ll 3:2, 3:6 ,

the excitation force

and the massmatrix

M = diag (0, 0, m3, 0, 0, O) (16.18)

the equation of motion for the model is

Mi+ Kx = f(t} (16.19)

It can be recognized in this simple example that the equation of motion arises
from the equation of static equilibrium by simple addition of Mi terms. The
inertia force of a lumped mass is thus catered for by placing its mass in a
position on the diagonal of the mass matrix corresponding to its displacement.
If the rotor vibrates not only in a plane but in space and if the displacement at a
node is defined by coordinates 3: and y, then a mass appears at the corresponding
position on the diagonal for both the 3: and y coordinates.
For a rigid circular disc of mass m and moments of inertia Id and I p , both
forces and moments are introduced as shown in Fig. 16.6 (see Sect. 4.1). For the
node in question an extension of equations (16.17) is appropriate as follows,

tIp Qx 3

tX4 tId x
4

f X2 f mX2

~m:'~$ -.... mx,


""""
I d X3 """"
I pQX4
Fig. 16.6. Rotating and vibrating disc. Forces and moments introduced as a result
of inertia effects.
16.2 Dynamics 219

Fl(t) - E k1r:l)r
F2(t) - E k 2r :l)r
(16.20)
Idz3 + I pilz4 F3(t) - E k3r :l)r
Idz4 - Ip ilz3 F4(t) - E k4r :l)r
For the whole model with several discs the following equation of motion holds

Mi + G:i: + Kx = f(t) (16.21)

with mass matrix

(16.22)

and the matrix of gyroscopic moments, a skew-symmetric matrix, as follows:

G= (16.23)

Now consider the element shown in Fig. 16.7, consisting of a ring B which is
flexibly connected to the hub A. The parts B and A are assumed to be rigid and
the flexible connection massless. Nodes a and b lie at the appropriate centres of
gravity.
The element has eight degrees of freedom and the matrices

M (A
dlag m , m A, IA
d, IA
d, m B, m B, IB
d, IB)
d (16.24)
G (9ik) with 934 = -943 = I: il, 978 = -987 = I: il (16.25)

Fig. 16.7. Ring-hub element.


220 16 Equation of Motion

K corresponds to the design of the flexible connection between A and B. If we


join this element to a shaft in such a way that node a on the element becomes
coincident with node c on the shaft, then of course these two nodes will have
the same displacement. As has already been shown in Sect. 16.1, such joining
simply results in the addition of the corresponding elements of the matrices of
the ring element and shaft. If the hub is not distended, that is I = 0 and if the
resulting disc is assumed rigid in the radial direction, then X5 = Xl and X6 = X2
and the matrices are correspondingly simplified.
If a rotor has regions which are much stiffer than the rest of the shaft, then
we can assume these to be infinitely stiff, to a first approximation. Such a case
is shown in Fig. 16.8. The motions at nodes b, S, c of the rigid portion are
inter-dependent and coupled kinematically. It is therefore sufficient to consider
only the motion at one of these nodes in the equations of motion. As a result the
matrices must be transformed to the coordinates at this node. This so-called
rigid body transformation is considered in a general way in Subsect. 19.1.3, and
we shall confine ourselves here to the application of the results. Any random
node on the axis of the rotor body can be chosen as a reference node. We choose
point b and must therefore transform the matrices of the rotor body and of the
beam c, d to the coordinates of b. The matrices of beam a, b remain unaltered.
Had we chosen the centre of gravity as a reference point, then the matrices of
the rotor body would have remained unaltered and the matrices of both the
beam elements would have been transformed.
The transformation is performed with matrix T using equation (19.27).
Here it has the reduced form
1 (
1 -(
T= (16.26)
1
1
The rotor body has mass m, polar moment of inertia Ip and diametrical moment
of inertia I d , about an axis through the centre of gravity. Thus its central mass
matrix is

and from equation (19.31), putting ( = -I., the transformed matrix for point b
IS

m -mi.

(16.27)

-mi.
1 Elementsnot shown are zeros.
2Elements not shown are zeros.
16.2 Dynamics 221

Fig.1S.8. Rigid rotor body on flexible shaft components.

The gyroscopic matrix remains unaltered. That is

G'=G (16.28)

-Ipil
In order to transform the matrices of beam c, d the T-matrix must be expanded
to the 8 x 8 matrix

(16.29)

and in equation (16.26) , must be written as -lc. With K as the stiffness matrix
of beam c, d the matrix of the expanded element for the rigid portion c, b is
obtained as
K' = T;xp KTexp (16.30)
The mass matrix of the beam is transformed in a similar way.

16.2.2 Beam Element

The beam element has already been introduced in Sect. 16.1 for static plane
bending. In most dynamic cases the bend is not all in one plane and so one
needs to introduce a second plane of bending. The corresponding beam element
is shown in Fig. 16.9. It has the notation already introduced in Subsect. 16.2.1
and applies to the two-dimensional case. The cross-section is constant and
circular along the length of the element and so from Fig. 13.2 h = III = I.
The stiffness matrix is formed simply by doubling the matrix (16.9) of the plane
222 16 Equation of Motion

--- 5 ""'""-
7

Fig. 16.9. Beam element bending in two dimensions.

element, observing the changed coordinates. Thus


al -a2 -al -a2
al a2 -al a2
a3 -a2 a4
a3 a2 a4
K= (16.31 )
al a2
symmetric al -a2
a3
a3

where al, ... a4 are obtained from equations (16.10). To determine the mass
matrix and gyroscopic matrix one could as a rough approximation distribute
the mass and moment of inertia of the element between its boundary nodes.
More exact matrices are obtained, if the bent form is represented by a Ritz-
substitution dependent upon the boundary displacements. The required matri-
ces are obtained by equating the work done by the boundary forces during a
virtual displacement, with the work done by similar forces along the element.
Usually the following expression is substituted for the bent form.

(16.32)

where ( = zll, the non-dimensional axial coordinate.


This applies to elements such as that shown in Fig. 16.9, for which the
horizontal plane is depicted as shown in Fig. 16.10.
A corresponding substitution is made for the other plane of bending, hav-
ing coordinates X2, CP3, X6, CP7. Substitution functions are selected as Hermite
polynomials
16.2 Dynamics 223

Hl = 1 - 3(2 + 2(3
H2 = ( - 2(2 + (3
(16.33)
H3 = 3(2 - 2(3
H4 = _(2 + (3 .

These are equal to the statically-bent form for the corresponding boundary
displacements (Fig. 16.10). The dynamically-bent form is naturally different
from this assumed form, and so the finite-element procedure is approximate. It
has, nevertheless, been ascertained that any errors are very small.
After some mathematical manipulation [115] the mass-matrix M of the
beam element is found to be

Cl -C3 C2 C4

Cl C3 C2 -C4

Cs C4 -C6

Cs -C4 -C6
M= (16.34)
Cl C3
symmetric Cl -C3

Cs

Cs

I" "I ~
0

-E----3-
H3 H1
, x1 , Xs H1.H3
'---' '---'
4>4 4>8 -0.2

~ t
0
z ~
0.2
w(z.t)
H2.H4
Fig. 16.10. Explanation of equation (16.32)
224 16 Equation of Motion

(16.35)

with e given as the last of equations (16.10). /Ld is the mass per unit length,
that is
m
/L=-=pA (16.36)
and I
/Ld = pl. (16.37)
The gyroscopic matrix of the beam element is

0 91 92 -91 92
0 92 91 92
0 94 92 93
0 92 93
G= (16.38)
0 91 -92
skew-symmetric 0 -92
0 94
0

where

(16.39)

and
/Lp = 2 /Ld = 2 P I (16.40)

16.2.3 Flexible Couplings

The separate shafts of a complete line-out can be connected together by ei-


ther rigid or flexible couplings. For rigid couplings the connection is treated,
16.2 Dynamics 225

as already explained in Sect. 16.1, as an example of two connected beam ele-


ments. Now consider the different technical possibilities arising from a flexible
coupling as being one of two types, namely a) a frictionless joint or b) a joint
with stiffness and damping.
For a frictionless, radially stiff joint the displacements at the two sides of
the joint are identical and the angles different. Fig. 16.11 shows, on the left,
the coordinates of two shaft elements which are to be joined. With :Z:9 = :Z:5 and
:Z:1O = :1:6 the new coordinates are shown on the right for the elements when

joined together.
The stiffness matrix of the combination has the following structure

1 2 3 4 5 6 7 8 11 12 13 14 15 16
1:z: :z: :z: :I:
2 :z::z: :z::z:
3 :z::z: :z::z:
4 :z: :z: :z: :z:
5 :z: :z: * :z: + + +
6 :z: :z: *:z: + + +
K= 7 :z: :z: :z::z:
8
11 + + + +
12 + + + +
13 + + + +
14 + + + +
15 + + + +
16 + + + +
(16.41 )
:z: relates to the left shaft element (aI, ... a4)
+ relates to the right shaft element (aI, ... a4)
* relates to al (left) + al (right).
with al . .. a4 from equations (16.10).
In the second possibility, namely the joint with stiffness and damping, we can
use the results obtained in Sect. 12.3. With the assumptions used there, the
relationship given in equations (12.42) holds for the moments and their rotations
and time-derivatives of rotations. Using the coordinates of Fig. 16.11 and J =
J7 - JIb 1/J = Js - J12 gives

M7 k (<(J7 - Jll) + di n(<(Js - J12) + (d a + d i ) P7 - c,Oll)

Ms = - di n (<(J7 - Jll) + k (<(Js - J12) + (d a + di ) (c,Os - c,012)

Mll k (<(Jll - J7) + din (<(J12 - Js) + (d a + d i) (c,Oll - c,07 )

MI2 = - di n(<(JU - J7)+ k (<(J12 - JS) + (d a + di ) (c,012 - c,OS)


226 16 Equation of Motion

~
13 ..........
15
~
5~
7 ..........
~ 11
1~
3

Fig. 16.11. Coordinates of two joined elements

Thus
f = Kx + Hie (16.42)

where

f (M7' M s , M n , M12)T
x (S07, SOs, SOn, S012f .
The stiffness matrix is

K = [ k
-k
-k
k
1 where k= [ ~
- d; {}
71~k {} 1 (16.43)

and the damping matrix

D =[ d -d 1 where d = [ 71 a + 71; ~ 0~ 1 (16.44)


-d d o da + d;
The K-matrix of the connected elements shown in Fig. 16.11 is finally obtained,
in which the matrix (16.43) is added, after corresponding extension, to matrix
(16.41).
16.2 Dynamics 227

16.2.4 Bearing Supports

The shaft line-out has several bearing supports. One bearing support consists
of the shaft journal, the bearing (:6.uid-film bearing or rolling-element bearing)
and the pedestal. The bearing shell is assumed to be fixed rigidly to the bearing
body. The two have a combined mass m. The shaft journal has the displacements
:1:1, :1:2 and the pedestal has the displacements :1:3, :1:4. One usually neglects
rotational (tilt) de:6.ections (Fig. 16.12).

Bearing shell
Bearing body

- X1

Journal Pedestal
Fig.16.12. Coordinates of a bearing support.

Following the procedure of Sect. 6.3 it can be said that, for a :6.uid-film
bearing,

(16.45)

where the stiffness and damping coefficients are given in equations (6.44). For
rolling-element bearings, from Sect. 9.2

(16.46)
and from Sect. 9.5

du = d22 = d, equation (9.34) and d 12 = d 21 = 0 (16.47)

For the forces at the bearing pedestal, we have

(16.48)
228 16 Equation of Motion

Hence, finally for the complete assembly,

-K D -D
+ (16.49)
K' -D D'

where

( 16.50)

The mass of the bearing pedestal is regarded as a point mass. It appears at its
corresponding place on the diagonal of the mass matrix.

16.2.5 Shaft Seals, Steam Whirl

In Chap. 10 it was demonstrated that, for a shaft seal, essentially the same force-
displacement relationship exists between the shaft and the seal component as for
a fluid-film bearing. Any differences occur only in the values of the coefficients.
The same applies for the forces between a blade row and its housing, giving rise
to steam whirl as discussed in Chap. 11. Accordingly, the corresponding matrices
for the two effects are found in a similar way to that given in Subsect. 16.2.4 for
the bearing support. If the seal component is assumed to be fixed to the housing,
then the relative displacements are zero and the matrices are correspondingly
reduced in size.

16.3 Reduction in the Number of Coordinates

For the mathematical model of a machine rotor, beam elements are often used
each having a constant cross-section. In the main, such rotors have many
changes of section along the axis, so that correspondingly more beam elements
and degrees of freedom need to be allowed for. In joining several elements to
form a single element with an averaged diameter the number of degrees of free-
dom can be reduced. However, it is better to follow the approximate method
described by Guyan in reference [116]. This method will apply to a system
having the following equation of motion

Mx+Bx+Kx=f(t) . (16.51 )
16.3 Reduction in the Number of Coordinates 229

Above all it is important that one should not eliminate coordinates, which de-
scribe important properties of the shaft. Hence, for example, one cannot elimi-
nate all the coordinates of rotation if one wishes to preserve the gyroscopic effect.
For a similar reason the coordinates of fluid-film bearings, seals, and their cor-
responding locations must be retained. As far as the shaft is concerned then,
in order to retain the matrix band structure, it is appropriate to reduce groups
of elements separately and not to carry out a reduction of the whole system.
Finally one must also consider whether such reduction serves a useful purpose.
It is beneficial if a greater number of repeated calculations is demanded, as for
example for resonance curves with different distributions of unbalance. After
these initial considerations we now come to the method itself.
With Xa as the master coordinates and Xb as the adjacent coordinates to be
eliminated, equation (16.51) can be used in the following form

[:: ::]{~} + [=: =:]{~}


+ [:: : : 1{ :: } ~ { ;:~:; } (16.52)
If we neglect the terms Mba, M bb , B ba , Bbb and fb(t) in the second of equations
(16.52), we are left with the equation

(16.53)

whence
Xb = -Kbb1 Kbaxa = -Axa (16.54)
The reduced equation is thus obtained,

(16.55)

in which M~a = Maa - MabA, B~a = Baa - BabA, K~a = Kaa - KabA.
This is the static reduction achieved in Sect. 16.1. The improvement of
Guyan consists in re-introducing the neglected terms into the calculations in
the following way. With the approximate solution vector

X={ xa}=[ E]xa=Txa


-AXa -A

and
5x = T 5Xa
230 16 Equation of Motion

then, from the principle of virtual work,

and the reduced equation of motion is

M*x,. + B*xa + K*xa = f*(t) (16.56)

where

The usefulness of this procedure depends upon how many and which coordinates
are eliminated. Experience obtained in the choice of adjacent coordinates allows
results to be compared for cases with and without reduction.

16.4 Discussion of the Equation of Motion

Using the procedures of Sects. 16.1 and 16.2 gives, for the model described in
Chap. 14, the equation of motion

Mi + Bx + Kx = f( t) (16.57)

where x is the vector of displacements and rotations


M is the mass matrix
B = G + D, the sum of the gyroscopic and damping matrices.
K is the stiffness matrix
and f(t) is the vector of excitation forces and moments.
The matrices have order n equal to the number of coordinates, and hence
degrees of freedom. Models usually have about one hundred degrees of freedom
but there can be more. The coefficients of the matrices are real and constant
or they depend upon the speed or upon other system parameters. In order
to solve the equations of motion, however, the coefficients are set to constant
values for each case considered. The peculiarity of rotor dynamics consists in
the occurrence of unsymmetric matrices alongside symmetric matrices.
In the sections mentioned, the equations of motion were developed in an in-
tuitive way from dynamic equilibrium, accounting for compatibility. In orthodox
mechanics, equations of motion are obtained using the Lagrangian equations or
the principle of virtual work.
Equations (16.57) are usually described as a set of linear differential equa-
tions of second order with constant coefficients. The individual terms are
inertia-, gyroscopic-, damping-, stiffness- and excitation-forces acting at the
various nodes. It will be observed that the geometry of the model in question
is expressed adequately through the elements of the matrices.
16.4 Discussion of the Equation of Motion 231

It has been seen that the finite element method consists mainly in setting out
the matrices and solving the equations of motion. How the matrices of a model
are obtained, was shown in Sects. 16.1 and 16.2. As far as the damping matrix
is concerned the following can be noted. Its structure for a single damper, as
for example at a bearing support, has been described. If one wishes to consider
distributed damping, then this can be done by the substitution

D = aM+{3K (16.58)

where a and {3 are estimated constants. Equation (16.58) is called mass- or


stiffness-proportional damping. The constants a, {3 can be calculated from the
modal damping. From equation (24.41), for the rth natural frequency Wr the
damping is

If this is separated into a term D M for the mass and a term D K for the stiffness
then
{3 = 2 DK . (16.59)
Wr

With assumed values for D M , DK and Wr one can thus calculate a and {3.
Although this is not a strictly correct procedure, it leads to realistic values
for a and {3. The solution of the equations of motion is described in detail in
Chap. 24.
17 Results

The model of Chap. 14 has many more parameters than does the simple model
of Part I. Correspondingly, its vibration behaviour is also more complex. Whilst
one can to some extent discuss the general behaviour of the simple model, it
is not possible for the complex model. Thus, we shall consider a few examples
and report results from further calculations. The essential vibration behaviour
of the rotors of turbo machines will thus be characterized.
The important basics for the understanding of vibration are the natural
vibrations. Obtaining them for a non-conservative system is by no means easy
and some generalities are described in a special section. Afterwards a rotor with
two bearings is described. On the one hand we thus have a basis for rotors with
several bearings, on the other we shall get to know how far we may use the
results of the simple model of Part I as an estimate for such rotors. After a
section on rotors with several bearings the essential results are summarized.
Part II ends with a chapter on the determination of the static loads on the
bearings.

17.1 General Discussion of Natural Vibrations

The natural vibrations of a system are described as in any vibration by the


motions at certain nodes. For our model these motions consist of displacements
perpendicular to the rotor axis and their corresponding rotations. Here it is
sufficient to discuss only the displacements, as the rotations can be found from
the bent form.
Consider the displacement i of a random node. From equations (24.20) for
the kth natural vibration

where Ctk, Wkare the real and imaginary parts of the eigenvalue Ak.
Tik, 8ik are the real and imaginary parts of the ith component of the kth
eigenvector.
Ck, "fk are constants found from the initial conditions.
The displacement thus follows an approximately harmonic motion with in-
creasing or decreasing amplitude. We neglect this change in amplitude and put
234 17 Results

ak = O. Further we make Ok = 1 and "'{k = 0, which for the following is quite


unrestrictive. Thus from equations (24.20)

(17.1)

or for displacements :1:1 and :1:2 at a certain node, omitting the index k,

:l:l(t) = Tl sinwt + 81 coswt


(17.2)
:l:2(t) = T2 sinwt + 82 coswt

The relations (17.2) describe an elliptical path, which is either forward or back-
ward in relation to the direction of rotation of the rotor. To find the sense of
rotation and the half-axes of the ellipse and their disposition, one introduces
the complex displacement

(17.3)

Using equations (17.2) and

sinwt = -2'j ( e Jwt


. - . )
e- Jwt

one obtains

z(t) = ~ [(T2 + 8t) + j (-Tl + 82)] e jwt + H( -T2 + 8t) +j (Tl + 8 2 )] e- jwt
and finally
z(t) = R' ei(wt+a') + R" e-j(wt-a") (17.4)
where
-Tl + 82
R' = ~J(T2 + 81)2 + (-Tl + 82)2 a' arctan---
T2 + 81 (17.5)
R" ~J( -T2 + 8t}2 + Crt + 82)2 a"
Tl + 82
arctan ----'--
-T2 + 81
In this form, one can recognize that the elliptical path consists of the summation
of two vectors. One vector of length R' rotates in the positive direction, the
other of length R" rotates in the negative direction. Both have angular velocity
w (Fig. 17.1).
At t = 0 the vectors are inclined to the axis Re( z) at angles a' and a",
respectively. For R' > R", then R' dominates and forward whirl prevails. Cor-
respondingly backward whirl ensues if R' < R". For R' = R" the motion is
along a straight line. From the figure it can further be seen that the following
relations hold,
a' + a"
a = R' + R" , b = IR' - R"I 0.= (17.6)
2
17.1 General Discussion of Natural Vibrations 235

Re(zl.xl-

Fig. 17.1. Elliptical path as the sum of two vectors.

To find the sense of rotation we formulate the expression,

R'-R"
u = -=---=- (17.7)
R'+R"
which we call the rotation number. Thus
for -1; u < 0 backward whirl exists
for o < u; +1 forward whirl exists (17.8)
and for u=O motion takes place along a straight line.

From relations (17.6),


b
lui = - .
a
(17.9)

The value of u is thus a measure of the breadth .of the ellipse. For u = 1 it
degenerates to a circle and for u = 0 it is infinitesimally small.
The nodes along the rotor axis describe ellipses with different half-axes at
different angular positions. The bent form of the rotor, here its natural mode
shape, consists of the inter-connection of points occurring on the individual
ellipses at the same instant in time. The natural mode shapes are spatial and
change periodically with periodic time Tk = 211"/Wk. This is true for our as-
sumption CXk = OJ otherwise, as is well-known the values increase or decrease
exponentially with time.
There are different ways in which the mode shape can be expressed. Balda
[22] gave a very graphic way. It consists in a spatial arrangement of elliptical
paths and shows the bent form at one point in time (Fig. 17.2).
Calculations show that, for a natural frequency, simultaneous forward and
backward whirls can exist on a rotor. At the position of the change over the
rotor moves along a straight line path.
236 17 Results

Fig. 17.2. Mode shapes of a rotor, from ref. [22].

17.2 Rotors with two Bearings

The machines considered in this book mostly have more than two bearings. Al-
most always assemblies consist of a driving and a driven portion. The vibrations
of a coupled assembly vary more or less from those of the individual machines,
depending on how stiff the couplings are. In any case it is useful to first consider
the behaviour of a rotor with two bearings. We thus begin with the following
example.

Example 17/1 It is required to calculate the natural frequencies and unbalance


vibration of the rotor shown in Fig. 17.3. It represents the rotor of a turbine as
a drive for a boiler feed pump. The essential data are
Power P = 16 MW
Running speed range n = 4000 - 5500 rpm
Mass m = 6690 kg
Total length L = 4 m.
17.2 Rotors with two Bearings 237

(l---..
'._ ._ ._ ._.----i._._._.---j-

Fig. 17.3. Rotor of example 17/1.

To distinguish between the individual influences on the natural frequencies,


the models of Table 17.1 will be considered. The same rotor will always be
assumed, whose special features are denoted by crosses.

Table 17.1. Models of example 17/1

Model 1 2 3 4 5 6 7
Bearings rigid x x x
Bearings flexible x x
Fluid-film bearings, rigid pedestal x
Fluid-film bearings, flexible pedestal x
Shear deflection x x x x x x
Gyroscopic action x x x x

Solution. The rotor is made up of fifteen beam elements. The bearing


pedestals are constituted as shown in Fig. 14.5. Model 7 thus has (16 x 4) + (2 x
2) = 68 degrees of freedom . Calculations were carried out using the program
MADYN.
The procedures used for calculating eigenvalues were the Householder pro-
cedure for models 1, 2 and 4, the Determinant search procedure with inverse
vector iteration for models 3 and 5 and the Hessenberg procedure for models 6
and 7.
For model 1, that is for rigid bearings, the vibrations are undamped and
the eigenvalues imaginary. Fig. 17.4 shows the first three mode shapes with
the corresponding natural frequencies, which are given in cpm to compare with
critical speeds. The mode shapes presented are denoted as of U, S or W form .
238 17 Results

f1 =3760 cpm

f2 =9285 cpm

Fig. 17.4. Mode shapes for rigid bearings (model 1).

For model 2 shear deflection is considered. Hence the natural frequencies


are lower by
5.0 % for 11
3.9 % for h
7.1 % for 13'
Any change in mode shape is minimal.
For model 3 gyroscopic action is considered. Its influence was investigated
in Chap. 4 for the simple model of a massless rotor with disc. It was found
that the natural frequencies depended upon running speed, that their number
doubled and that changing reverse and forward whirls were encountered. The
natural frequencies are shown in Fig. 17.5, in which the numbering was adapted
to the increased number of natural frequencies. Hence hand h replace 11 and
correspondingly for the higher values. The changes in natural frequencies are
very small. For example at n = 6000 rpm,

+ 1.44 %
-1.34 % } on 11 for model 2

+ 1.69 %
-1.76 % } on h for model 2

+ 0.24 %
- 0.24 % } on h for model 2.
17.2 Rotors with two Bearings 239

t cpm f6
fk fs
10000
f4
f3

5000-
f2
f1

0
0 5000 10000 rpm
n-
Fig. 17.S. Influence of gyroscopic action on the natural frequencies.

Model 4 was chosen to show the influence of bearing flexibility. For this purpose
isotropic stiffness was assumed and the flexibility hB was changed from zero
to 2h s . Here hs = 1/mwi and represents the flexibility of the rotor, which is
10.7 X 10- 10 miN. The results are shown in Fig. 17.6. The natural frequencies
decrease with flexibility, for example for hB = 0.5 hs the decrease is

15 % for iI
5.3 % for h
20 % for fa.
Using model 5 it can be demonstrated how strongly gyroscopic action influences
the natural frequencies when elastic bearings are included. Calculations gave a
small increased influence.
Model 6 has fluid-film bearings, while retaining rigid bearing pedestals. The
bearings are assumed to be circular, each with two axial grooves and LI D = 0.5.
The stiffness and damping coefficients are obtained from reference [10], page 14.
The bearing pressures are 334 NI cm 2 for bearing 1 and 165 NI cm2 for bearing
2. For both bearings, the relative clearance tjJ = 1.5 X 10- 3 and the dynamic
viscosity of the lubricant 'f/ is assumed to be 0.03 Ns/m2.
The stiffness and damping coefficients are functions of Sommerfeld number.
In [10] this is defined as
'f/n
S = ptjJ2 ' (17.10)

while in this book it is, from equation (6.34),


240 17 Results

f cpm
fk
10000

f3

f2
5000

f1

a
a 2
hB/hs-
Fig. 17.6. Natural frequencies as functions of the flexibility of the bearings.

The relation between the two is thus

S=_I_ (17.11)
211" So
The actual coefficients for this example are shown against rotor speed in
Fig. 17.7, being obtained from the non-dimensional coefficients given in ref-
erence [10]. Whilst the stiffnesses have various characteristics, all the damping
coefficients decrease monotonically with rotor speed. This is mainly due to the
fact that, in the relationships for dik , the journal speed occurs in the denomina-
tor. In the graphs showing the stiffnesses, the stiffness of the rotor, kR = mw~,
is indicated as a dashed line. It can be seen that, in this example, the bearing
stiffnesses are of the same order as the rotor stiffness.
The eigenvalues were calculated using the Hessenberg procedure. As well as
complex conjugate eigenvalues there also occur several real values which are
large and negative. These describe heavily damped motions, which have no
significant meaning as far as the mechanics of the system is concerned.
In Fig. 17.8 the first five natural frequencies are given as functions of rotor
speed, while in Fig. 17.9 the corresponding damping ratios are shown. With the
exception of h, these natural frequencies have rather flat characteristics against
rotor speed and are somewhat smaller than for the case of rigid bearings. In
the region considered the values of D3 are very large and so the third natural
frequency is heavily damped. This natural frequency is thus of little significance,
and no resonance exists. The remaining natural frequencies give critical speeds
as 2200, 3320, 8100 and 8700 rpm.
17.2 Rotors with two Bearings 241


So 0,5 0,5

o~----~------~ o~----~------~

k21 ____

-1 -1

10 6 Ns/m
d22~6NS/:
t I.

dik d11
2 d11 ,---- 2 ____

0 0

-2 d12
d 21
?::
5000 10000
n( rpm) - -
d12/
-2 d 21
5000-10000
n( rpm) - -
Fig. 17.7. Coefficients for bearings 1 and 2 and their dependence on speed.

The damping ratio D1 is negative for speeds greater than 4800 rpm and
so at this speed the system becomes unstable. In practice one would therefore
choose bearings with a different geometry so as to achieve greater stability.
From the results for the JejJcott rotor it can be expected that changing
forward and backward whirls will occur. The rotation numbers calculated from
equation (17.7) show the following behaviour:
Natural frequency 11: +, h: -, !J: +, 14: + and -, 15: + and -
where + stands for forward whirl and - for backward whirl.
242 17 Results

jRi9id bearings
fs 1/
'1
t
- 1
10000 /

~
fs
fk
cpm
f4
/yl/ f4

~4
/
f:ny/ f3
5000

4/
/
~
f2
>~
"r:
f2
f,
/

+==}
/
/
0 f,
5000 rpm 10000
n-
Fig. 17.S. Natural frequencies and mode shapes for model 6.

03
t
Ok 0.10
04

0.05 O2

05
0
5000 rpm 10000

-0.05 \ n--

Fig. 17.9. Damping ratios for model 6.

For model 7 the bearing pedestals are assumed to be flexible. Otherwise it


has the properties of model 6. Hence the masses of these bearing pedestals are
called into play. The following data, defined as in Fig. 14.5, were assumed to be
the same for both bearings.
m 1000 kg
1 X 109 N/m, d3 0.10 X 10 6 Ns/m
2 X 10 9 N/m, d4 0.14 X 106 Ns/m.
17.2 Rotors with two Bearings 243

These data give a damping ratio D = 0.05 and natural frequencies 9550 and
13505 cpm for the bearing pedestals alone in horizontal and vertical directions,
respectively.
Because of the additional degrees of freedom there are now eight natural
frequencies in the frequency range considered instead of five as for model 6
(Fig. 17.10). Natural frequencies hand /4 are heavily damped and hence have
little significance. The same applies to /5 and h in certain speed ranges, as the
damping ratios given in Fig. 17.11 indicate. The critical speeds are 2150, 3100,
7750, 8250 rpm.
The stability borderline is at 4360 rpm and so is a little lower than for
model 6.

jRigid bearings
f= n
10000 'y
/

=
f7f8
fk
cpm
f6
//1/1/ :
f4 /
/
/ fs
5000

/
/
/
O~,-~~-,-,-.-.~r-.-.--
o 5000 rpm 10000
n-
Fig. 17.10. Natural frequencies for model 7.

5000 rpm 10000

-0.05 ~ n--

Fig. 17.11. Damping ratios for model 7.


t.:)

40 -I
"'"'""
40
h
x IJ.m I h~ A
h ~ ~v
1 ......
-.J
V
~
20 ~
20 -I 11\\/ \ 20 ~ ~V \ "J M' a.....'"
'"
~ ) ~"
0-9 ..;=
oU,~ I , , , 0
0 5 10 0 5 10 0 5 10
n (1000 rpm )
Node 5 Node 9 Node 12

IJ.m

20] 20
x
~__L_~__ ) ,
0 <> j~ y , , ', ~
, , 11'1 ""-----"" ltl "iif
I
0
0 5 10
0 5 10
Node B1 B1 5 9 12 B2
Node B2

Fig. 17. 12. Model 7. Responses for a single unbalance U = 33.43 kg mm.
l~
x
50 50 50

0 '9 r. O~ 'T""'1"> i , ~ 0
0 5 10 0 5 10 0 5 10
n(1000 rpm)
Node 5 Node 9 Node 12

..-
~

,:j ~j ~
L flV J~hi'" .",i.lii ::0
0
....
0
....
en

'J
:l1l
_.
i C; --=r= , i' 'I ,~, ....
O- O- ~,
0 5 10 0 5 10 =-....
81 5 9 12 82 :l1l
0
Node 81 Node 82
t:I:l
Cb
Fig. 17.13. Model 7. Responses for an unbalance pair U = 16.72 kg mm. II>
::l.
=:0
(JI:I
en

t.:I
,;.
c:.n
246 17 Results

Finally, two unbalance cases were calculated for model 7. In the first case
a single unbalance of value U = 33.43 kg mm was assumed at node 9, that is
roughly at the centre of the rotor. For a rotor mass m = 6686 kg, this repre-
sents a mass excentricity e of 5 /Lm. Fig. 17.12 shows the calculated amplitudes
of vibration for rotor nodes 5, 9 and 12, as well as for nodes Bl and B2 on
the bearing pedestals. For this particular excitation the first two resonances
dominate at the rotor nodes.
In the second case a pair of unbalances of values U = +16.72 kg mm at
node 5 and -16.72 kg mm at node 12 were assumed (Fig. 17.13). By this means
the two higher resonances were excited, while the first two hardly appeared.
These two unbalances are the only cases considered in this example. In
practice unbalance distributions are essentially some c?mbination of these two.

The vibrational behaviour found in example 17/1 is typical of such rotors.


They can be concisely defined as rotors having a thick central portion and
thin ends (Fig. 17.3). Another type is the disc rotor, consisting of one or more
discs, which are arranged on a thin shaft. For such rotors gyroscopic action is
essentially stronger than for the former type. This feature will be elucidated in
the next example.

Example 17/2 Fan rotor, Fig. 17.14. The shaft mass is 8000 kg and its running
speed is 740 rpm. For calculation purposes the bearings are considered simply
as springs with the following stiffnesses in the horizontal and vertical directions,
respectively
k3 = 3.33 X 108 N/m, k4 = 6.67 X 108 N/m, for bearing 1
k3 = 0.83 X 108 N/m, k4 = 1.67 X 108 N/m, for bearing 2.
By comparison the rigidly supported rotor has a stiffness at the disc position
of 0.82 x 108 N /m. The calculated natural frequencies are shown in Fig. 17.15.

Fig. 17.14. Fan rotor, from [32), p. 183.


17.3 Rotors with Several Bearings 247

~/'i"'"--- _ _ _ _ f1

1000 2000 rpm


n-
Fig. 17.15 Natural frequencies ofthe fan rotor of Fig. 17.14.

Without gyroscopic action, that is at n = 0, the first three natural frequencies


are at
549, 615 and 2110 cpm.
and with gyroscopic action the critical speeds are at 461, 787 and 1900 rpm.
The differences are thus
-16 %, +28 % and -10 %, which are considerably higher than in example 17/1.

17.3 Rotors with Several Bearings

Rotors with several bearings consist of individual rotors which are either rigidly
or flexibly coupled together. To describe the general vibration behaviour of
coupled rotors would be difficult and would have no practical value. We must
therefore treat each situation as a special case, which today presents no great
problem. We confine ourselves here to the following example, from which the
essentials can be appreciated.

Example 17/3 The natural frequencies and mode shapes as well as the unbal-
ance vibrations are to be determined for the rotor system shown in Fig. 17.16.
The main data are:
Power P = 315 MW
Running speed n = 3000 rpm
Mass m = 131 t
Length L = 29.5 m.
The following results were obtained using the program MADYN. The rotor
model has 39 beam elements (Fig. 17.17). For rigid bearings it has 34 nodes,
each having four degrees of freedom and six bearing nodes, each having two ro-
tational degrees of freedom. In total the model thus has 148 degrees of freedom.
248 17 Results

HP MP LP Gen Ex
1 2 3 5 6

Fig. 17.16. Rotor of example 17/3.

Fig. 17.17. Rotor model

Calculations using the Householder procedure gave the natural frequencies


and mode shapes shown in Fig. 17.18, covering the range 0 - 5000 cpm. The
natural frequencies can be characterized to some extent by the sections of the
rotor suffering the largest vibration, as shown by the notation in the right-hand
column of the figure.
The question as to what the influence of the number of degrees of freedom
has on the results is considered by recalculating for 19 instead of 39 beam ele-
ments. The reduction of coordinates was decided as indicated in Sect. 16.3. The
natural frequencies differ but little from those of the first case. The maximum
difference occurs at the sixth natural frequency and is then only one percent.
This calculation of natural frequencies assuming stiff bearings serves only
to give a first estimate. A more realistic estimate is obtained by considering
fluid-film bearings and flexible pedestals as shown in Fig. 14.5. This is analysed
for our example by using the calculations of Eckert in [117], denoted plant E,
model RO. Data for the fluid-film bearings and bearing pedestals are also taken
from this work.
The calculated natural frequencies and damping ratios are shown in Fig.
17.19. Orders 5, 10 and 12 are omitted, because the natural vibrations in ques-
tion are very heavily damped. To some extent the strength of the damping can
be recognized by the type of curve displayed in Fig. 17.19. Thus one has
weak damping
average damping
- - - heavy damping.
Accordingly the most important natural frequencies are those numbered 2, 4, 6
and 11, whose values lie near to those of the model with rigid bearings. Hence
it can be seen that the use of the more simple model is justified, at least to
a first approximation. The mode shapes are spatial and, like the eigenvalues,
17.3 Rotors with Several Bearings 249

HP MP LP Gen Ex

f,= 917 cpm Gen U

fz= 1576 cpm LP U

MP U

f4=2604 cpm HP U

fs=2936 cpm Gen 5

LP 5

Fig.IT.IS. Natural frequencies and mode shapes. Model with rigid bearings.

depend upon the rotor speed. In Fig. 17.20 are shown these mode shapes sim-
plified as plane curves, and relate to a rotor speed of 3000 rpm. These curves
are projections at the point in time corresponding to maximum deflection. On
the right are shown elliptical orbits at the nodes having the greatest deflection.
Thus it can be seen, for example, that the first natural frequency has an essen-
tially horizontal character and the second a vertical character. By comparison
with Fig. 17.18, one can, roughly speaking assign the following mode shapes
and hence natural frequencies as being similar.

Fig. 17.18 Fig. 17.20


number 1 1,2
number 2 4
number 3 6
number 4 8
number 5 9,11.
250 17 Results

cpm r-:::: id bearings

---------- 13--
fk 3000 ~ V fk=n
~ 11~/~--
====_-:::_-_:::::::: -/- 9 - -----
/8 ---------
2000 ~ /
~ . . j=._7_6 . f - . -
---f-: 4 - - - t - -
1'.-3-'-'
1000 ~ ~2_===:-::I==
I I
Vi k
O~~-.--~.-~--~~
o 1000 2000 3000
rpm n- rpm n-
Fig. 17.19. Natural frequencies and damping ratios. Example 17/3 model with
oil-film bearings and flexible pedestals, from ref. [117].

It can be seen here also that the more simple model is very useful as a first
approximation.
To judge the unbalance vibrations of this example many unbalance distri-
butions would have to be chosen and the amplitudes of vibration at many nodes
calculated. We confine ourselves here to five cases of unbalance and to the dis-
cussion of vertical vibrations ofthe pedestals for bearings 1,3 and 4 (Fig. 17.21)
single unbalance in HP
single unbalance in MP
- -- - single unbalance in LP
- - - single unbalance in Generator
unbalance pair in Generator.
The assumed unbalances are U = me from a single unbalance and U = 0.5 me
from an unbalance pair. m is the mass of the rotor in question and e = 10 ILm
is its eccentricity from the axis of rotation. The figure shows that each unbal-
ance mainly excites only one resonance. Correspondingly the greatest vibration
amplitude occurs in the region of the unbalance position. For example a single
unbalance at the generator gives rise to an especially pronounced resonance at
bearing 4 at a speed n ::::: 880 rpm. This corresponds to natural frequency h. A
single unbalance at the LP-rotor gives a strong resonance at bearings 3 and 4
17.3 Rotors with Several Bearings 251

k fk 100 Dk HP MP LP Gen Ex
-I- cpm ./. 2 3 I. 5 6
81.5 0.51.5
1 1 A1+
2 883 0.057
1 1 1 ~+
3 1293 2.097
1~A 11+
I. 11.51 0.1.01
+=t=A -=+-1 +
6 1733 0.658
'k:Ft=l 11+
7 1771. 5.362
A 1 1 II~
8 2300 1..153
r+-l 1 11+
9 21.59 2.623
1 1 1 ~+
11 2680 0.636
1 1 1 -kA+
13 3311. 6.765 ~1/P\} ~II~
Fig. 11.20. Mode shapes corresponding to Fig. 17.19 for n = 3000 rpm. From ref.
[117].

for n = 1450 rpm corresponding to /4' Finally an unbalance pair at the genera-
tor excites an S-vibration of the generator at n = 2680 rpm, corresponding to
/n. This is demonstrated by the large amplitudes at bearing 4 in Fig. 17.21.

With the results of example 17/3 one can conclude the following for rotors
with several bearings:
- The mode shapes and natural frequencies of the model with rigid bearings are
useful in judging the vibration behaviour of rotors with fluid-film bearings
and flexible bearing pedestals.
- By considering the fluid-film bearings and pedestals one can obtain the many
natural frequencies in the operating range, of which only a few are of
importance in assessing quiet running.
252 17 Results

k- 4 6 8
I I I

1:1+---r--r-1-'I,~A;4- ah,~i
~~
Bearing 1

.:.;.a:

X 0 1000 2000 3000 rpm

4 6 8 11

I
I I I I
J'm

30

X
20
Bearing 3

10

o~~~~~~~~~==~~
o 1000 2000 3000 rpm

k-2 4 6 11

I
I I I
I'm

30
I !
X
I ."
I jl~)
~\ ; \
20
Bearing L.

....

l
10
..............

,
04---~~~'~~~'~-~~--~
~!~~ -
o 1000 2000 3000 rpm

Fig. 17.21. Example 17/3, Amplitude responses. From Eckert [117].


17.3 Rotors with Several Bearings 253

- The significance of each natural frequency can be judged by the corresponding


damping ratio. Natural frequencies with a damping ratio greater than
about 0.05 give as a rule only weak resonances. Further considerations
relating to the correspondence between damping ratio and peak vibration
at resonance are given in Sect. 21.2.
18 Static Bearing Loads

In this book, the vibrations considered are small motions about the static equi-
librium position. In most cases, the rotor is supported horizontally and the
principal load is its selfweight. The static equilibrium position and also the
static bend of a horizontal rotor depend on its mass and stiffness distribution
and the number, arrangement and stiffness ofthe bearings. For rotors with more
than two bearings, it depends also on the alignment.
A special case is provided by fluid-film bearings. These work like plane non-
linear springs, whose properties depend on speed. The static bend of a rotor
supported on many bearings can thus only be calculated in an iterative fashion.
The problem described was discussed in detail by Gasch in reference [118J.
Using his example one can appreciate the influence of fluid-film bearings on
the static bearing loads. For a turbo-set with eight bearings, the bearing loads
were calculated as functions of speed, taking into account a certain state of
alignment. The vertical loads were found to change and indeed could become
larger or smaller. Horizontal bearing loads also arise because the load loci of
individual bearings are different.

Table 18.1. Change in vertical bearing load and ratio of horizontal to vertical bearing
load for the turbo-set of Fig. 18.1.

Bearing 1 2 3 4 5 6 7 8

V3000 - Vo
% 3.6 - 25 12 -11 6.4 - 9.3 4.1 0
Vo
H3000
% - 3.1 17 - 3.4 6.9 - 2.9 4.7 - 2.2 0.3
V'tnnn

VO, V3000 vertical bearing load for n = 0 and 3000 rpm


H3000 horizontal bearing load for n = 3000 rpm.

Table 18.1 shows the essential results from this example. The first row con-
tains the percentage changes in vertical bearing load from rest to operating
speed. The maximum change amounts to - 25 % and occurs at bearing number
256 18 Static Bearing Loads

HP MP LP Gen

1 2 3 4 5 6 7 8
Fig.IS.1. Turbo-set from reference [118). P = 50 MW, n = 3000 rpm.

two. The alignment was so chosen that at standstill no horizontal loading oc-
curred. With increasing speed horizontal load arose, having variable direction.
The maximum occurred again at bearing number two and the change amounted
to 17 % at the operating speed.
The changes found in this example have little meaning in themselves when
we are only considering the stresses of the rotor and the load of the bearings,
but larger influences are possible on the stability of the system.
In the case considered there are always two bearings between two component
rotors. There are, however, other arrangements where only one bearing is used.
Here the effects described are much weaker and largely insignificant.
Part III

Rotor-foundation Coupling
In Part II a model was considered, in which the rotor support was substi-
tuted by a simple mass-spring system. By this means a bearing pedestal can be
represented very effectively by its equivalent mass, stiffness and damping. Good
results can be obtained using this model if the dynamics of the foundation or
more generally the support structure have sufficiently small influence on the
rotor dynamics. The object of this part of the book is to show firstly how one
can consider mathematically the coupled system of rotor and support structure
and secondly, with a few results, determine whether or not a coupled calculation
needs to be considered. A knowledge of the vibration behaviour of each subsys-
tem is first required. The behaviour of the rotor has already been described in
some detail and the behaviour of the foundation will be investigated in the next
chapter.
19 Foundation

Rotating machines are supported in different ways. The following figures show
some examples.
The most simple support is direct mounting in a building (Fig. 19.1). An-
other possibility is offered by a frame foundation as shown in Fig. 19.2. The
frame usually consists of sheet steel or of steel sections and is supported on
springs. Instead of a frame a massive concrete block is also often used, which
either lies directly on the ground or is mounted on springs (Fig. 19.3).

~~'i?~~~%/:/ Fig. 19.1. Direct mounting in building structure

lOp
Fig. 19.2. Frame Foundation

g F;g.l .. Foundation hl.ok


262 19 Foundation

For larger turbomachines a foundation table made from steel or from steel-
reinforced conrete is often used. The table is connected directly to the supports
(Fig. 19.4) or via suitably arranged springs (Fig. 19.5). The latter is termed a
spring foundation.
This particular kind of mounting is not shown in its complete form, but is
sufficient for a description of vibration calculations.

/.h'77777'77"77"":'77-:r:r7?"";'77TT:'777777:""~ Fig. 19.4. Foundation table

Fig. 19.5. Spring foundation

19.1 Calculations

For vibration calculations of foundations one can differentiate between rigid and
elastic foundations. The foundation block belongs to the first type, as does the
frame foundation when the frame is sufficiently stiff.
The flexibly supported rigid foundation has six degrees of freedom and can
be analysed with well-known methods. Flexible foundations can be analysed
by the method of finite elements and foundation blocks mounted directly on
the ground can be analysed with a knowledge of ground dynamics. The next
subsection deals with such dynamics and establishes certain fundamental prin-
ciples.

19.1.1 Ground Dynamics

The requirement here can be summarized as follows: A rigid massless plate


lying on a plain surface representing ground executes small displacements and
rotations (Fig. 19.6). It is required to find the forces and moments which act
on the plate and emanate from the ground.
The answers depend on the type and structure of the ground. These can
be ascertained from measurement or calculation. For calculation purposes the
ground is assumed to be an infinite half-space consisting in the simplest case
of homogeneous isotropic material with elastic properties. Further one could
19.1 Calculations 263

Fig. 19.6. Rigid plate and ground.

assume the presence of viscous damping. More important, however, is damping


due to radiation of energy. This results from waves leaving the point of excitation
and disappearing at infinity and not by reflection at a finite structure.
From these few observations it is obvious that ground dynamics is a very
complex subject. Correspondingly there are many publications, of which the
work [119) of Lord Rayleigh can be regarded as the forerunner. We confine
ourselves here on the one hand to obtaining simple relationships for estimation
purposes, and on the other to a short description of a comprehensive procedure
used to consider the ground when dealing with elastic foundations.
The following relationships emanate from the work [120) of Ehlers. In this
the ground was replaced by a model of a one-dimensional frustum of a cone
of infinite length (Fig. 19.7). At its upper end a mass m is assumed and the
effect of a harmonic excitation is determined, using the equations of continuum
mechanics.
For linear elastic material the following ratio holds between the amplitude
F of excitation force and amplitude w of displacement.

F = ~ (EA
w h -m w2 )2 + E p( A W)2 (19.1)

For the simple system shown on the right in Fig. 19.7 then, from equation (2.28),

(19.2)

Fig. 19.7. Ground model after Ehlers [120J.


264 19 Foundation

Thus, by comparison with the ground model, the stiffness and damping are

kz = EA
h '
dz=VJ!Jp
FiiE A . (19.3)

It can be seen that no damping has been assumed for the material itself. The
damping arises only from energy radiation. Ehlers called this geometric damp-
ing.
The notation is as follows:
1-v-v2
E E <;::::jE
1- v - 2v 2
E modulus of elasticity
v poisson's ratio
P density
A surface area at the upper end of the cone
h distance from ground to the apex of the cone.

The ground parameters can vary considerably but can be assumed to lie within
the following ranges.
E (5 - 80)10 7 N/m 2
v = 0.25 - 0.50
p (1.65 - 1.85)103 kg/m3

The value of h is uncertain. Ehlers recommended h = VA and so, with E = E,

(19.4)

With the same ground model, reference [121] gives, for excitation in the plane
of the ground, the coefficients

(19.5)

where G is the shear modulus.


The relationships are only to be thought of as very rough. In the litera-
ture one can find other relationships which can result in noticeably different
numerical values.
A new procedure for considering the ground, especially for elastic founda-
tions is given by Schilling [122]. The problem of coupling a discrete system,
the foundation, with the continuum of the ground is solved in such a way that
a discrete model is also assumed for the ground, as will be explained in the
following.
The contact surfaces between foundation and ground are discretised into a
number of right-angled elements, whose centres are defined as nodes (Fig. 19.8).
19.1 Calculations 265

Contact surface
Fig. 19.8. Ground model after Schilling [122].

The foundation has identical nodes. The influence of the ground is reduced to
the effects of forces lx, Iy , Iz at these nodes. These forces are assumed to be
proportional to the nodal displacements 1, v, wand to their velocities and
accelerations. The ground motion is thus represented by the following equation,
which relates the vector x of nodal displacements with the vector f of nodal
forces
Mi: + Di + Kx = f( t) . (19.6)
This equation is combined in the usual way with the equation of motion of the
foundation to obtain the equation for the combined model.
The coefficients of the matrices M, D and K of the ground are calculated,
using a procedure due to Schollhorn [123], in the following way.
For harmonic excitation with f(t) = fej"'t ,1 the solution from equation (19.6)
is ~(t) = xej"'t where

(19.7)

The coefficients of the matrix H(w) are the transfer functions defined in
Sect. 24.4.4. With excitation frequencies WI, W2, ... Wz the following system of
equations is obtained,

w~M + jW1 D + K = H(wt}


w~M + jW2D + K H{W2)
(19.8)
w~M + jwzD + K H{w z )

1Complex quantities are underlined.


266 19 Foundation

from which the coefficients of the left-hand side can be found by the least-
squares method, for a known right-hand side and a sufficient range of excitation
frequencies.
The kinetic flexibiIities of the right-hand side of equations (19.8) are calcu-
lated in ref. [122] using linear half-space theory for a visco-elastic ground. The
expenditure in calculation time can be reduced for repeated calculations with
the same ground parameters, in that one can retain a certain size of element
and store the kinetic flexibiIities.

19.1.2 Rigid Foundation Block

Consider a rigid block of general shape, which is supported on a number of


randomly arranged springs and dampers (Fig. 19.9). At the centre of gravity
lies the origin of an orthogonal z, y, z-system, which need not correspond to the
major axes.
The vibrations are described by the motion of the centre of gravity. The
equation of motion is
Mi: + Dx + Kx = f( t) (19.9)
where
x (Zt, 3:2, Za, CP4, CP5, cps)T
f (Ft, F2 , Fa, M4 , M 5 , Ms)T

M
[4] (19.10)

#. (19.11 )

5 Y X2 IPs

Fig. 19.9. Rigid foundation block with general support.


19.1 Calculations 267

and the inertia tensor

Ixx -Ixy -Ixz


1= -Iyx Iyy -Iyz (19.12)

-Izx -Izy Izz

where the moments and products of intertia are

Ixx = J(y2 + Z2) dm, Iyy = J(z2 + z2)dm, Izz = J(z2 +y2)dm
Ixy = Iyx = J zy dm, I yz = Izy = J yz dm, Izx = Ixz = J zz dm .
(19.13)
The stiffness matrix is found from the spring element matrices by transformation
to the coordinates of the centre of gravity and addition. We first transform by
rotating to coordinates parallel to the z, y, z-system and then by translation
to the coordinates of the centre of gravity. Details of these transformations are
described in Subsect. 19.1.3. With K~ as the transformed stiffness matrix for
spring i, then for the whole block K = E K~. The damping matrix is obtained
in a similar way.
In most cases the springs are so arranged that the preload plays no part. The
simple model of Fig. 19.10 shows, however, that this preload must occasionally
be considered. Without preload this model has stiffnesses

ky = 2k , kz = 0 , kip = 0

but with tensile preload Fo the stiffnesses become

(19.14)

For compression preload kz and kip are negative and the motions in z and cp are
unstable.
The system of Fig. 19.9 has six degrees of freedom, which in general are
coupled. Partial decoupling is achieved by suitably arranging the springs and
dampers. For an arrangement such as that shown in Fig. 19.11, the displacement

t~
H~H
I I bib I I Fig. 19.10. The effect of preload.
268 19 Foundation

~6 , ~6 t

, .; J~i:"
X31 m. I xx I yy I zz X31

blbl LdJ
Fig. 19.11. Symmetrical foundation block.

in the z-direction and the rotation about the z-axis are fully decoupled, while
the displacement in the x-direction and the rotation about the y-axis and vice-
versa are coupled.
In this case the equations of motion are

[: :,,]{ r;;6 }+ [~ ~]{ }~{ }


X3 X3 F3(t)

[: :J {
CP6 M6(t)
Xl
r;;5
}+[kn k"]{ }~ { } Xl F1(t)
(19.15)

[: :J {
k51 k55 CP5 M5(t)
X2
r;;4
}+[~" k" ]{ }~{ }.
k42 k44
X2
CP4
F2(t)
M4(t)

where lxx, lyy and lzz are the principal moments of inertia corresponding to
axis x, y and z.
The configuration shown in Fig. 19.11 is not the only possibility for de-
coupling using equations (19.15). It is implicitly assumed that the supporting
system has a certain symmetry in relation to the central major axes of the block.
The four springs of Fig. 19.11 must be equal. For one spring we have from
Fig. 19.12,

(19.16)

The essential stiffnesses are k/ and kt , since the coefficients kep and ktep = kept
may usually be neglected in the x, y, z system. With further approximations
obtained by neglecting the torsional stiffness of the springs, the stiffness coeffi-
cients for the x, y, z system are
19.1 Calculations 269

Fig. 19.12. Spring coordinates.

k33 = 4k, , 1%6 = 4(a 2 + b2)kt


kll = 4kt , k15 = k51 = -4 c kt
k55 = 4 (a 2k, + c2kt ) (19.17)

k22 = 4kt , k24 = k42 = 4 c kt


k44 = 4 (b 2k, + c2k,)

19.1.3 Flexible Foundation

We now consider foundations such as those shown in Figs. 19.4 and 19.5, but
also structures as in Fig. 19.1 and frames as in Fig. 19.2. These are space or
plane structures, whose vibrations can be calculated using the finite element
method. It is necessary to set up and solve an equation of motion oftype (24.1)
and a major requirement is in establishing the matrices. In the following it will
be shown how one can proceed.
The structure can be built up with one-, two- or three-dimensional ele-
ments. We shall confine ourselves here to beam elements and rigid-body ele-
ments. Each beam element will have a constant cross-section and in general un-
equal bending stiffnesses in two mutually perpendicular directions (Fig. 19.13).
Its motion will be described by the displacements of nodes A and B. The dis-
placement coordinates lie in the major axes of the element. Each node has six
degrees of freedom, and so the element matrices have order (12, 12).

Fig. 19.13. Beam element for space structures.


270 19 Foundation

Frequently at a corner of a structure the centre lines of the connecting beams


do not intersect as for example in Fig. 19.14. In this case one can construct a
mechanically representative connection using a rigid-body element.
Rigid-body elements can also be used as approximations for regions which
are significantly stiffer than the rest of the structure.
With these elements a calculation model can be built from the real structure
and its matrices can be ascertained by the following procedure:

Establish the system of coordinates for the model

Establish and assign numbers to suitable nodes

Specify boundary conditions

Match the elements

Transform the element matrices to global coordinates

Build up the matrices of the structure.

Fig. 19.14. Corner with rigid-body element.

In order to describe further details of these procedures let us consider the exam-
ple of Fig. 19.15. This figure shows on the left the actual structure and on the
right its model. The x, y, z-system of coordinates is positioned at the centre
of a cross member, although it could be placed anywhere. The model has 21
nodes. This number is arbitrary and represents the nodes required to obtain the
necessary accuracy. Nodes at the meeting points of different elements and at
the support points must be specified. Further nodes are then specified in order
to allow the vibration of the structure to be ascertained to sufficient accuracy.
For this purpose nodes are nominated for example which are half way up the
vertical supporting members in Fig. 19.15. Experience has to be used to choose
these arbitrary nodes. One must be able to imagine how the area in question
will vibrate and which vibrations can essentially influence the results.
19.1 Calculations 271

6f
3f /s

z
*~
"-....
1Y,

Fig. 19.15. Example of a space structure

The nodes can be numbered in a random fashion. Since one specifies the
matrices in relation to this numbering, this will influence the matrix structure.
Because one wishes to obtain, as far as possible, a band structure, the nodes
are numbered accordingly. Many programs have an automatic means of band
structuring.
The coordinates of nodal displacements, so-called global coordinates, are
taken parallel to the coordinates x, y, z and in the same positive directions.
The boundary conditions are specified by the choise of the displacement co-
ordinates. Extreme cases are the completely free and completely built-in nodes.
In the first case all six coordinates must be assumed and in the second case they
must all be omitted.
To determine the positions of the elements in the system we establish a
matching list. We give to each element a number and assign to it the numbers
of the corresponding nodes of the system. For beams there are two numbers and
for point masses just one number.
The element matrices relate to certain coordinates. The originally specified
coordinates, for example the coordinates 1 to 12 in Fig. 19.13, are called local
coordinates. In the structure these local coordinates do not usually agree with
the global coordinates. In order to be able to set up the structure matrices
from the element matrices in a simple way, these element matrices must be
transformed to global coordinates. The necessary steps are made clear using
the beam element of Fig. 19.16.
Begin with the transformation of local displacement coordinates Xl, X2, X3
of node A to the global coordinates x~, x~, x;. With angles aI, (31, III

Xl = x~ cos al + x~ cos (31 + x~ cos 11 .


272 19 Foundation

xl
y

x
Fig. 19.16. Directional transformation

Using corresponding angles for and X3,

r}
X2

eo, {3, c'"' 7' 1 x'I


X = X2 = [co,a, cos cos
COS<l!2 {32 /2
{ x'2 } ~ To'" (19.18)
X3 cos <l!3 cos {33 cos /3 x'3

Similarly for the forces

FI

F3
} ]{
r~ { F2 ~ [ To
F'1
F'2 }
F'3
~Tor (19.19)

Further, using matrix K,

f = Kx and Tof' = KTox' .

Hence by multiplying the left-hand side by TO- I = T[ one obtains the relation-
ship
f'=K'x' ,
where K' is the transformed matrix, that is

K' = T~KTo. (19.20)

Rotations are assumed to be small, so that they, like the displacements, have
the properties of vectors and so can also be transformed using To. The beam
is straight, so that the coordinates of node B are parallel to those of node
A. Hence for these nodes the transformation using To is valid. The complete
19.1 Calculations 273

transformation matrix of the beam is finally

To
To
T total = (19.21)
To
To

With K total as the stiffness matrix of the beam element with reference to its
local coordinates, the matrix for the global coordinates is given as

(19.22)

The mass and damping matrices are transformed in a similar way.


The elements of To are the direction cosines of coordinates :1:1, :1:2, :1:3. Using
the space coordinates of nodes A and B of the beam of length 1,
YB -YA ZB - ZA
, cos /31 = 1 ,cos '}'1 = 1 . (19.23)

In a similar way the remaining cosines could be calculated using the space
coordinates of further points on the :1:2- and :l:3-axis. In general, however, it is
easier to proceed as follows. Assume a random point C(:l:C,YC,zc), which lies
in the :l:ll :l:2-plane, although not on the :l:l-axis. The slave coordinate :1:0 thus
defined has, analogous to equations (19.23), the direction cosines
Zc -ZA Yc - YA Zc - ZA
cos a o =
10
, cos /30 = 10 ,cos '}'o = 10 (19.24)

where
(19.25)

the distance between A and C. To continue further we use the fact that the
direction cosines are the components of the unity vectors in question.
el = (cos all cos /311 cos '}'d
eo = (cosao, cos/3o, cos'}'o) .
The unity vector for :1:3 is the cross-product of e1 and eo, that is
e3 = e1 x eo = (cosa3' COS/33, COS'}'3)
where
cos /31 cos '}'O - cos '}'1 cos /30
cos /33 cos '}'1 cos ao - cos a1 cos '}'O (19.26)
cos '}'3 cos al cos /30 - cos /31 cos ao .
The remaining direction cosines are found using the product
e2 = e3 X el = (cos a2, cos/32, COS'}'2) .
274 19 Foundation

In this way all the required values at the space coordinates of points A, B, C
are found. It should be noted also that apart from lengths I and 10 the only
important factors are the differences in coordinates.
After the matrices of all elements are transformed into global coordinates,
the coefficients of the structure matrices result simply from addition of the
element coefficients belonging to their own coordinates. This was described in
detail for a simple plane example in Sect. 16.1. In the present case one has the
analogous extension to a space structure.
Let us now return to the rigid body element. This is defined as an un-
changeable configuration of two or more nodes. Fig. 19.17 shows an example
with four nodes. In general the element is massless and can move freely in space.
Its movement is specified by the six displacements and rotations of one of its
nodes. The displacements and rotations of the remaining nodes depend on those
of the reference node. Consider small displacements and rotations, so that the
relationships are linear. For displacements and rotations of nodes P and Q in
Fig. 19.17,
:VI 1 ( -'TJ :v'1
:V2 1 -( e :v'2

Xp =
:V3 1 'TJ -e :v'3
=TxQ (19.27)
CP4 1 cp~
CP5 1 cP~
CP6 P 1 cP~ Q

e,
where 'TJ, ( are the coordinates of P relative to Q. For the transformation of
forces and moments from P to Q we have

F'1 1 FI
pI 1 F2
2
pI 1 F3
fQ = 3
= TTfp (19.28)
M'4 -( 'TJ 1 M4
M'5 ( -e 1 M5
M'6 Q -'TJ e 1 M6 P
or inversely

If one further puts


fp = Axp
then, using equation (19.27),
19.1 Calculations 275

1) - - -
/4 y 2 - -5
x

Fig. 19.17. Rigid body element with four nodes.

and after multiplication by TT

fQ = TT ATxQ = A'xQ

A' =TTAT .
where
(19.29)
This is the matrix transformed to the coordinates of Q. From equation (19.27)
the transformation matrix is

T=[~] , L = [ -( (-~ 1 (19.30)


1] -~

It must be noted that the origin of the ~,1], (-system lies at Q and that the
coordinates ~,1], ( in general are either posi ti ve or negative from P to Q.
If the rigid-body element is a structure having mass, for example a stiff
housing, then one usually describes its motion by the motion of its centre of
gravity. The mass matrix relative to its centre of gravity is, from equation
(19.10),

M= [4].
For the coordinates of a different node Q, the transformed matrix is, from
equations (19.29) and (19.30)

M' = [-ifh-] [4] [~] [*]


276 19 Foundation

where

[m
J [-~ -m']
m(
A m , B-
- me (19.31)
m", -me

C
[ I" + m(,'
-Iyx - m",e
+ (') -Iw - m" -I" -
Iyy + m((2 + e) -Iyz - m",(
m,e ]
.
-Izx - m(e -Izy - m(", Izz + m(e + ",2)

It can be seen that the linear transformation employed is only valid for a trans-
formation proceeding from the centre of gravity. From a different point there
are additional terms in expressions (19.31).
If a beam element is connected to a rigid-body element P, Q, then the beam
is expanded to include this element. With P as the connecting node and Q the
reference node of the rigid-body element, the new element has either the nodes
A, Q or Q, B (Fig. 19.18). To transform the matrices of the beam element, T
must be expanded to a (12,12) matrix using expressions (19.30). Thus

Ttotal for element A, Q


( 19.32)
Ttotal for element Q, B.

Using K as the stiffness matrix of the beam element, then for the expanded
element,
(19.33)
The matrices M and D are transformed in a similar way.

y
x
Fig. 19.18. Beam with rigid-body element.
19.2 Results 277

19.2 Results

The vibration of foundations and similar structures as described at the begin-


ning of the present chapter can be determined using the methods of Sect. 19.1.
In the following the most important results from such calculations will be sum-
marized and discussed.

19.2.1 Rigid Foundation

A rigid foundation mounted on flexible supports has six degrees of freedom and
hence six natural frequencies. For general support conditions usually each nat-
ural vibration contains all of these six displacements and rotations to a greater
or less degree. The number of parameters is too large to allow a meaningful
general discussion of the natural modes. We shall therefore consider instead the
important special case of the symmetrical foundation block as previously shown
in Fig. 19.11.
The particular arrangement of springs in this case gives rise to extensive
decoupling of the various motions. From equations (19.15) the vertical displace-
ment X3 and the rotation CP6 about the vertical axis are uncoupled. The displace-
ment Xl is coupled with rotation CP5 (rocking in the z, X plane) and displacement
X2 is coupled with rotation CP4 (rocking in the y,z plane).

The corresponding natural frequencies are

We,d = / C =f J C2 - D
where
C = mk55 - Iyyk11 D = k11 k 55 - k15 k 51
(19.34)
2mIyy mIyy

We,! =/ E =f J E2 - F
where
E = mk44- Ixxk22 F = k22k44 - k24k42
2mIxx mIxx
The orders of the natural frequencies depend upon the values of the parameters
and are different from case to case. The indices a, ... f were therefore chosen so
that in an individual case indices 1, ... 6 can be reserved for successive natural
frequencies.
In the following example more general conditions are given.
278 19 Foundation

Example 19/1 A prismatic foundation block is given, which carries a machine


assembly which is assumed to be rigid. The foundation stands directly on the
ground (Fig. 19.19). The centres of gravity of the three main components are
indicated as A, Band C. It is assumed that the principal axes of these three
components are parallel to the axes :/;, y, z. The following moments of inertia
relate to the assembly. The rotor masses are given in brackets.

A B C
m t 69 10 (3) 6,5 (2)
I",,,, tm 2 230 12,1 5,7
IIJIJ tm 2 59 10,6 3,2
I zz tm 2 243 14,7 5,7

For the ground,

E = 3.0 X 108 N/m2 , G = 1.15 X 108 N/m 2 , p = 1750 kg/m3

y x

I. 6.0m
L B
W
Fig. 19.19. Example 19/1. Foundation block with rigid machine.

It is required to determine the natural frequencies and the unbalance response


of the block with and without the machine assembly.
Solution. To calculate the stiffness coefficients of the ground we use equa-
tions (19.4) and (19.5). With A = 2.5 m X 6 m = 15 m 2 ,
kz = 11.62 X 108 N/m , k", = klJ = 4.47 X 108 N/m .
These coefficients are valid for evenly distributed support conditions. To be
able to use equations (19.15), the ground is replaced by four equivalent springs
19.2 Results 279

as shown in Fig. 19.11. Using the sides B, L, H of the foundation block the
following relationships are obtained,

H (19.35)
kll = kx, k1S = kSl = -kx2 '
H
k22 = k y , k24 = k42 = kY2

The calculated natural frequencies of the system without the machine assem-
bly are shown in Fig. 19.20. At the first and second natural frequencies, the
block executes rocking motions about an axis below the block and for the fifth
and sixth natural frequencies about an axis above the block. The third nat-
ural frequency has a rotating motion about the z-axis and the fourth has a
translational vibration in the z-direction.

cpm 535 702 769 1239 1288 1391

1>-----~-~----,L--..L-L-~------.-----~~(
o 500 1000 cpm 1500
fk -
Fig. 19.20. Example 19/1. Natural frequencies, foundation block without machine
assembly.

For the block with its machine, symmetry no longer holds and so the sim-
ple equations (19.15) are no longer valid. More precise calculations (using the
Jacoby-procedure) gave the values shown in Table 19.1.
As the machine assembly is assumed to be rigid, the system has six degrees
of freedom as before. Because of the increased mass (85.5 t instead of 69 t), all
natural frequencies are lower than before and since A is proportional to V1/m,
the influence of the increased mass can be estimated. We thus find that the

difference = (J 69 -
85.5
1) .100 = -10 % .
280 19 Foundation

Table 19.1. Example 19/1. Natural frequencies

Block without Block with Difference


machine machine
cpm cpm %
II 535 413 -23
h 702 598 -15
h 769 687 -11
14 1239 1110 -10
Is 1288 1156 -10
16 1391 1227 -12

The approximation can thus be seen to be acceptable.


In calculating the unbalance response the following questions need to be
answered.

- Which natural frequencies must be noted as important?

- What significance has ground damping?

The damping coefficients of the ground are determined using equations (19.3)
and (19.5) and give the values,

dz = 10.9 X 10 6 Ns/m , dx = dlJ = 6.8 X 106 Ns/m.

The elements d11 , d66 of the damping matrix are found from equations (19.35)
by substituting dx , dy , dz for kx , klJ, kz We assume an unbalance at each node
Band C, firstly in phase and secondly in antiphase, putting these as

UB = 30 kg mm , Uc = 20 kg mm .

They correspond to an eccentricity of 10 p,m and calculations gave the ampli-


tudes shown in Figs. 19.21 and 19.22.
For in phase unbalances, a resonance at n i::::! /1 appears in the horizontal
direction.
We assess the damping from the half power points, and using equation (2.38)
for a simple oscillator, we obtain a damping ratio D = 0.3.
In the vertical direction the amplitudes grow with rotational speed in a
monotonic fashion and a resonant peak is not observed.
For antiphase unbalances (Fig. 19.22) the horizontal resonance also disap-
pears and the amplitudes are very small. For this example it can be concluded
that
19.2 Results 281

f ~m Node B f ~m Node C

x x
2 2

horizontal horizontal

vertical vertical

rpm rpm
n- n-
Fig. 19.21. Example 19/1. Amplitudes for in phase unbalance.

~m Node B ~m Node C
f f horizontal
x 0.4 horizontal x 0.4

0.3 0.3

0.2 0.2

0.1 0.1

n- n-
Fig. 19.22. Example 19/1. Amplitudes for antiphase unbalance.

- there is a noticeable resonance at n ~ /1 and

- there is very heavy damping due to the ground.

For the calculations so far carried out the machine assembly was assumed
to be rigid. One can imagine that a flexible rotor will be little influenced by
ground damping around its own resonance. To investigate this more precisely,
we assume that machine B has a flexible rotor having the following properties
282 19 Foundation

- Jeffcott rotor with a point mass

- m = 3000 kg, k = 118 X 106 N/m.

The natural frequency of the rotor on rigid bearings is obtained from

60 118 X 106
ik=-
211" 3000 = 1894 cpm .

If we assume that the rotor is stiff in the axial direction, the system now has
eight degrees of freedom. Calculations show that the first six natural frequencies
can be assigned to the foundation block and the seventh and eighth to the rotor.
These latter and the corresponding damping ratios are

h = 1937 cpm, D7 = 0.021


and is = 1994 cpm, Ds = 0.011 .
The natural frequencies h and is vary only a little from the natural fre-
quency ik on a rigidly mounted block, the differences being 2.3 % and 5.3 %.
The damping ratios D7 and Ds are very small in relation to D = 0.3 for the rigid
system. Even so the ground damping makes itself felt on the rotor in spite of
the large foundation block. In order to calculate unbalance vibrations, a value
e = 10 I'm, that is an unbalance UB = 30 kg mm was assumed. It gave the
resonance amplitudes

:C1 = 16 pm, :C2 = 238 pm for n ~ b


2)1 = 50 pm, 2)2 = 422 pm for n ~ is

where 2)1 is the amplitude at the left bearing of rotor Band 2)2 is the amplitude
at the rotor centre (node B). At the first of these resonances the rotor vibrates
essentially in the vertical direction and at the second in the horizontal direction.
This can be ascertained from the mode shapes.
For the Jeffcott rotor the resonance amplitude is
~ e
x=-. (19.36)
2D
This relationship gives with D = D7 and D = Ds good approximate values for
the maximum amplitudes in our example. For the resonance in the first case,

2) = 10/(2 x 0.021) = 238 pm == 2)2

while for the second,

2) = 10/(2 x 0.011) = 455 I'm = 1.08 ~2


19.2 Results 283

19.2.2 Frame Foundation

To calculate the vibration characteristics of a flexibly-supported frame foun-


dation for small and medium sized machine assemblies we usually assume the
frame and the machine to be rigid bodies. We calculate the machine assem-
bly separately and assume the bearings to be either rigid or flexible. Usually
measured values or values from experience are used for the bearing stiffnesses.
Fundamental to this type of calculation is the procedure adopted in the present
section. We should, however, be aware just how useful this procedure is and
when a more exact model needs to be chosen. To answer these questions exam-
ple 20/1 in Chap. 20 will later be considered.

19.2.3 Table Foundation

When computers were not available, table foundations were replaced by a simple
model, which was so chosen that the essential vibrations could hopefully be
exhibited. With a computer one can of course calculate for much better models,
having many degrees of freedom. Thus one obtains many natural frequencies
and as a result one has the difficulty of distinguishing the most important
vibrations in order to ultimately modify the foundation design. The significance
of a given natural frequency can be recognized from its mode shape, given the
requisite experience. It is easier to judge the dynamic stiffnesses (or alternatively
receptances) at the bearings. The best, although with the largest expenditure of
time, is the calculation of the coupled system of foundation and machine which
can be undertaken as in Chap. 20. We confine ourselves here to the foundation
and consider a representative example.

Example 19/2 A table foundation is given for a turbo-generator set running


at 3000 rpm and producing a power of 125 MW. It consists of steel-reinforced
concrete and has a mass of 1210 t (Fig. 19.23). The rotor mass is 81 t, and the
mass ratio is thus 1/15. It is required to calculate the natural frequencies and
to discuss the mode shapes.
This problem was solved as a student project [124]. The mathematical model
is shown in Fig. 19.24. It consists of the foundation and the bearing housings
and has
79 nodes
76 beam elements
30 rigid bodies and
218 degrees of freedom.
The ground plate was assumed to be rigid and the supports to be rigidly re-
strained.
The natural frequencies were calculated using the program MADYN, from
which, amongst others, the following results were obtained.
284 19 Foundation

E
N
+-~----------------~

E
Q

Fig. 19.23. Example 19/2. Table foundation


27m of a 125 MW turbo generator set.

Fig. 19.24. Example 19/2. Mathematical model.

11 = 76 cpm /32 = 2497 cpm


12 = 80 cpm /33 = 2655 cpm
/34 = 2741 cpm I'll = 4110 cpm
/35 = 3247 cpm 142 = 4218 cpm

In order to judge the natural frequencies, we display the mode shapes in


different projections. We can roughly split them up into:

- Rigid body vibrations

- Table vibrations
19.2 Results 285

- Cross-member vibrations and

- Vertical-support vibrations.
The mode shapes are rarely in a form appropriate to one category alone. They
are mostly mixed together and frequently a rearrangement is diffucult or im-
possible.
For our example, using an operating speed of 3000 rpm the natural frequen-
cies 33, 34 and 35 are of special interest. The corresponding mode shapes are
shown in Fig. 19.25.

f33= 2655 cpm

f34 =2741 cpm

f35 =321.7 cpm

Fig. 19.25. Example 19/2. Some mode shapes.

More severe vibrations of the turbo generator set are to be expected if large
excursions of the bearings take place in a transverse direction for any particular
natural frequency. Accordingly for speeds in the region of n ~ /33, more severe
vibrations are to be suspected in the forward area of the set (left in Fig. 19.25).
The speed range around /35 may well be harmless as far as the foundation is
286 19 Foundation

concerned, because then essentially axial vibrations of the cross members occur
and no excitation is present in this direction.
To illustrate and judge the mode shapes requires a considerable expenditure
of time, as this example has shown. Simpler, and in the opinion of the author
quite sufficient, is the calculation and assessment of the dynamic flexibility
(or receptance) at the bearing positions in horizontal and vertical directions.
Generally, the receptance is given as hik(W) = Zik/ f,.. It is the amplitude of
vibration at coordinate i as a result of harmonic excitation at coordinate k,
divided by the amplitude of this excitation force. In general extensive coordinate
coupling is present and so many values of hik(W) exist. We need, however, only
consider the direct values hii(W), that is the receptances at the position of the
excitation. For z bearings we thus have only 2z characteristics to consider.
Fig. 19.26 shows as an example the calculated receptances at the positions
of the generator bearings of a set producing power P = 350 MW at speed
n = 3000 rpm. The natural frequencies are marked by short lines. It can be
seen that only a few resonances are evident from the many natural frequencies
and that those responsible for such resonances can be easily recognized.

Vertical
t
h 44 .2
1111 III II II III
h
t 0.2 III
f18
1111 II II III

55
f18

JVt--
Ilm Ilm
kN 0.1 kN 0.1

o I Iii I i I I I I I o
o 2000 3000 cpm o 2000 3000 cpm
f-- f--
Horizontal
t
h 44 0.2
III I I 11/ II" III
h
t 0.2 IIII 1111 II II III
55
Ilm Ilm
kN 0.1 kN 0.1 f21

o
o
~",:::
2000 3000 cpm
o
o
~A~
cpm
2000 3000
f-- f--

Bearing 4 Bearing 5
Fig. 19.26. Receptances of the foundation of a turbo-generator set at the positions
of the generator bearings. After reference [125].
19.2 Results 287

19.2.4 Spring Foundation

For a spring foundation vibration calculations can be confined to a system of


table and springs. Because vertical supports are not present, there are much
fewer natural frequencies, and so there is little essential difference from a ta-
ble foundation. Fig. 19.27 shows the results of calculations for a foundation
made of steel-reinforced concrete having dimensions 12 X 32 X 3.5 m and mass
1400 t. There exist about half as many natural frequencies as for a similar table
foundation.

)7. I

~III' i I i Ii Ii i All frequencies

I ., Rigid body
6
I I Bending vertical
8 12 18
I Bending horizontal
910 14 2023
Torsion
7 11 16 24
Cross member, bending
13 21 vertical
Cross member, bending
151719 22 longitudinal
I I I

0 20 40 cps 60
Frequency f -
Fig. 19.27. Natural frequencies of a spring foundation (from [32]).

19.2.5 Reinforced Concrete and Steel

Table foundations and spring foundations are mostly made of reinforced con-
crete. However, a welded construction of steel is often preferred. To decide which
to use, we need to know amongst other things, what is the vibration behaviour of
the turbo-generator set with a steel foundation. To answer this question many
288 19 Foundation

details must be noted. This is clearly shown by the investigations of Weinmann


[126], -whose results are discussed in the following.
A turbo-generator set was investigated having n = 3000 rpm, P = 215 MW.
It has a steel foundation and the set has been in operation for over twenty
years. For comparison, a suitable foundation made of reinforced concrete was
envisaged. Unbalance vibrations for both foundations were conducted, when
supporting the turbo-generator set.
The two arrangements are shown in Fig. 19.28 and the various masses are
as follows.

Steel Reinforced concrete


Table 246 t 2129 t
Casing 530 t 530 t
Bearing pedestals 39 t 39 t
Stationary mass 815 t 2698 t
Rotor 127 t 127 t
Ratio 1/6.4 1/21.2

The following unbalance vibrations were calculated:


Rotor with oil-film bearings, rigidly supported.

35200

a
:5
N
Steel
~

"
o
o
00
N

a
a Reinforced
2
~
concrete

"
Fig. 19.28. Turbo-generator set of investigation [126].
19.2 Results 289

Both foundations with bearing pedestals.


The complete coupled systems.
The individual systems employed have the following numbers of degrees of free-
dom:
96 for the rotor
280 for the steel foundation and
202 for the reinforced concrete foundation.
Calculations gave 68 natural frequencies for the steel foundation and 51 natu-
ral frequencies for the reinforced concrete foundation in the range from zero to
4200 cpm. The static flexibilities of the bearing supports for the steel foundation
are:

40 to 100 p,m/kN horizontally and 0.6 to 1.7 p,m/kN vertically

and for the reinforced concrete foundation:

4 to 9 p,m/kN horizontally and 0.2 to 0.3 p,m/kN vertically.

Similar ratios are obtained for the receptances. For the steel foundation larger
amplitudes are found at resonance with a larger separation between resonances.
Proportional damping was assumed (Eq. (16.58)) for the calculation of re-
ceptances and unbalance vibrations. The values of Q and f3 were chosen so that
the assumed values of damping ratio were obtained for the frequencies 2700 and
3300 cycles/min for the foundations in question (Fig. 19.29).
The unbalance vibrations depend upon the value and distribution of un-
balance as well as upon the damping. Representative values were assumed as
follows: For each of the HP, LP1, LP2 and Gen components, a possible pair
of in phase and antiphase unbalances due to 10 p,m mass eccentricity. The am-
plitudes of horizontal and vertical vibration of the bearing housings and of the

f 0.03
o

0.02

'\
Reinforced concrete
0.01

Steel

04---~--.---~---r----
o 2000 4000 cpm
f-
Fig. 19.29. Assumed damping ratios for comparison of two foundations.
290 19 Foundation

shaft at the bearing centre were determined. The frequency range was confined
to 2700 to 3300 cpm. These gave 512 graphs which were displayed in 256 figures
in an easily-assimilated manner.
A discussion of all the many results would be out of the question here. We
therefore consider only those shown in Figs. 19.30 to 19.32, which feature the
main essentials. These figures show amplitudes at bearing 6 (LP2 aft) and at
bearing 8 (Gen aft).
At bearing 6 the amplitudes with a steel foundation are greater than with
a reinforced-concrete foundation. We might expect this result since the steel
foundation is the lighter and more flexible. A similar behaviour is displayed by
most of the other results not shown here. There are exceptions, however, which
are shown in Fig. 19.32. At bearing 8 the amplitudes are roughly the same

Bearing Shaft
ilm ilm
f f
40 ,. 40
x x

20 20

o o
o 2800 3200 rpm o 2800 3200 rpm
n- n-
Fig. 19.30. Amplitudes at bearing 6 in vertical direction for in phase unbalance.

- Steel, - - - Reinforced concrete

Bearing Shaft
11 m 11m
f
,. 80 f
,. 80
x x

40 40

- - - , / ........
... __ .... "
0 0
0 2800 3200 rpm 0 2800 3200 rpm
n- n-
Fig. 19.31. Amplitudes at bearing 6 in vertical direction for anti-phase unbalance.

- Steel, - - - Reinforced concrete


19.2 Results 291

~m ~m

x'"
[,0
.- .- ----- Shaft x [,0
20 20
---,- --__ Shaft

0
----- -----~ Bearing
0
--___- Bearing
0 2800 3200 rpm 0 2800 3200 rpm
n- n-
Fig. 19.32. Amplitudes at bearing 8 for anti-phase unbalance.

- Steel, - - - Reinforced concrete

for both foundations and indeed for both directions and also for the bearing
housing and the shaft.
This example shows how complex is the problem under discussion. On the
one hand it confirms that a turbine-generator set having a steel foundation
vibrates more violently and over a wider range than when reinforced concrete
is used. It does show, however, that there are regions in which practically no
difference exists. One can thus generally conclude that by careful tuning quiet
operation can be achieved using a steel foundation. The great number of such
foundations in practical use testifies to this.
20 Rotor and Foundation

After the publication of the transfer matrix method [6, 7] it was possible to
determine the vibration characteristics of rotors supported on several bearings.
Furthermore, it became desirable to include the foundation and the casings.
Before that time their influence could only be estimated or judged from the
results from simple models.
A part solutiori to this problem was achieved by Weber [127] by extending
the transfer matrix method to a model having two beams. One beam was the
rotor and the other represented the table foundation. Both beams were coupled
by springs at their supports and the foundation beam was also supported on
springs representing the vertical supports. As a result a new appreciation of
their combined vibration behaviour was achieved, although the lower beam was
only a rough representation of the foundation.
An essentially better and indeed sufficiently accurate model of the founda-
tion can be assumed in using the finite element method, which is described in
Chap. 19. In principle the coupling with the rotor brings nothing new. One sim-
ply adds a further structure, which is catered for by the well-known addition of
corresponding elements of the matrices. Further details of this will be presented
in Sect. 20.1.
The model of the foundation has symmetric matrices of high order. The
rotor model, on the other hand, has few degrees of freedom by comparison, but
by its nature its matrices are non-symmetric. For the coupled system there exist
many degrees of freedom, as well as non-symmetric matrices, and this demands
a considerable expenditure in computing time.
Publications in this field are above all characterized by the way this expen-
diture can be reduced. A very effective method was put forward by Pons [128]
as component mode synthesis. By this means the solution is approximated by
a linear combination of a sufficient number of eigenvectors of the subsystems
representing rotor and support structure. The order of the matrices is thus re-
duced from n to m, where m is the number of substitution vectors. In reference
[128] an example is given in which n = 6000 and m = 70.
The solution is made easy for the case of unbalance vibrations by consider-
ing the foundation as represented only by its dynamic stiffnesses at the bearing
positions. These are obtained by inversion of the separately determined recep-
tances. This method is described in references [129] and [130] and in the next
section. A similar method, which is not confined to unbalance vibrations is de-
294 20 Rotor and Foundation

scribed by Wang and Lund [111]. In this the rotor and the foundation are first
calculated separately and then combined by "impedance matching".
Calculations involving the combined system of rotor and casing encounter
the same problems as for the rotor and foundation combination. Occasionally
the casing model has more than ten thousand degrees of freedom, and so the
reduction of expenditure in computer time is even more urgent. Vinsonneau
and Lemant give an example of such a system in ref. [131]. They suggest the
reduction of the stationary components after the manner of Guyan and the
completion of the calculations after making the matrices symmetrical, following
an averaging procedure.

20.1 Calculations

Using the rotor model of Chap. 14, the total structure outside the bearing shells
is replaced by spring and damper combinations which are supported rigidly
(Fig. 14.5). In considering the dynamics of the foundation and in general the
supporting structure the rigid bearing support is replaced by the model of this
structure. The connection can be done directly to ground or by interposing
a rigid body to represent the bearing pedestals (Fig. 20.1). The equation of
motion of the coupled system is comprised of the equations of motion of the
rotor and the support structure by adding the corresponding matrix elements.
It is assumed that the corresponding coordinates are parallel and have the same
positive directions.
With order nR of the rotor and nF of the support structure the equation
of motion of the coupled system has order nRF = nR + nF. We need this
equation of motion in order to calculate the natural frequencies. The calculation
of unbalance vibrations or vibrations resulting from general harmonic excitation
is then achieved with considerably fewer coordinates. This will be illustrated in
the following.

Pedestal (rigid body)

Fig. 20.1. Model of a bearing pedestal.


20.1 Calculations 295

From Sect. 24.2 one obtains, for harmonic excitation, the particular solution
in the form of the equation,
K(w}& =f 1 (20.1)
with complex vectors for the amplitudes

f of the excitation forces and .& of the displacements


as well as the dynamic stiffness,

(20.2)

After this general consideration, we now take the separate systems of rotor
and support structure. The positions of the separation can be arbitrary but at
certain boundaries. We can separate the subsystems as shown in Fig. 20.2, for
example. Separated nodes are points A' and A" with connecting coordinates
indicated with arrows. Before separation A' and A" would be coincident.
The rotor model is free and the model of the support structure is supported
in a way depending upon the type of structure. For each of the two subsystems,
equation (20.1) holds with the dynamic stiffness obtained from equation (20.2).
After sorting the two equations in terms of connecting coordinates and the
remaining coordinates, the equation of the rotor becomes

(20.3)

and of the support structure

(20.4)

I~,I /j -Pedestal
i "I
t /1

y
I A"
~ 1 /cross member

Rigid body

Fig. 20.2. Example of a point separation (stiff bearing housing, supporting structure,
stiff part of supporting structure).
lUnderlined parameters are complex
296 20 Rotor and Foundation

In these equations,

XA' X~ are the amplitudes of the displacements


'"
fA, fA the amplitudes of the forces at the nodes at which the sub-
-II

systems are separated and


fR are the amplitudes of the excitation forces.
A = A(w), ... H(w) are the dynamic stiffness matrices.
On reconnecting the subsystems the compatibility conditions must hold, namely

(20.5)

As a result the following equation is formulated for the coupled system by the
use of equations (20.3) and (20.4)

(20.6)

Elimination of ~F from the third row finally gives the reduced equation

(20.7)

where
(20.8)

Equation (20.7) only contains the coordinates of the rotor and connecting nodes
and so is the most significantly reduced in contrast to the original equation.
It follows from equation (20.4) that KA = KA(W) is the dynamic stiffness of
the support structure. With HA(w) as the corresponding receptance we have,
KA(w) = H A1(W) . (20.9)
The coefficients of HA (w) can be obtained with relatively little effort from modal
considerations, so that it is more beneficial to calculate KA(w) from equation
(20.9) than from equation (20.8).

20.2 Results

The simplest model of a rotor-foundation system is shown in Fig. 20.3. From


the equations of motion,
20.2 Results 297

Fig. 20.3. Simple model of a rotor and foundation.

mR XR + kR XR -kR XF = 0
(20.10)
mFXF-kRxR +(kR+kF)XF =0

we obtain the natural frequency ratios

(20.11 )

where
(20.12)

and the natural frequencies

WR = J
kR
mR
of the rotor and
(20.13)
WF = J
kF
mF
of the foundation.

In spite of the simplicity of the model its natural frequencies readily show im-
portant features which are also found from complicated models. From Fig. 20.4
we conclude:

The natural frequency Wl is smaller than the smaller natural frequency of


the subsystems of rotor and foundation and W2 is larger than the larger
natural frequency.

- For a relatively large foundation mass (mF ~ mR) the natural frequencies
of the coupled system are about equal to the natural frequencies of the
rotor and the foundation.

Equal tuning (WF = WR) is not dangerous. Coupling only produces a split in
Wl < WR and W2 > WR
298 20 Rotor and Foundation

2+---------+3~--~~~-

O~~~~_r-+~~-,~~r_~~

a 2
~-
WR
Fig. 20.4. Natural frequencies of the model of Fig. 20.3.

After this introduction we shall now consider more complicated models with the
object of ascertaining the influence of the foundation on vibration behaviour.
We shall also be able to judge, under what conditions the complete calculation
of coupled systems appears to be necessary. Hence we have to decide between
small machine sets with frame foundations and large turbo-alternator sets. In
the main we calculate the first type as if uncoupled, while the second is nowadays
often calculated as a coupled arrangement.

20.2.1 Small Machine Sets with Frame Foundations

A small machine set is, according to our classification, a set consisting of two
individual machines, one driving and the other driven. For such a set vibration
calculations are carried out in a simple manner. This because, on the one hand, a
more comprehensive study is not regarded as necessary, and on the other because
any greater expenditure of time is not warranted in relation to the objectives
required. A typical example of such a composite system is the ventilator or fan
shown in Fig. 20.5.
Experience has however shown that it would sometimes be better to make
greater use of modern vibration methods here. To explain this the following
example will be considered. Results and recommendations will be summarized
at the end of this subsection.
20.2 Results 299

5160

Fig. 20.5. Ventilator. Rating 550 kW, speed 1480 rpm. a

Example 20/1 For the ventilator of Fig. 20.5, the vibration characteristics will
be calculated first in the usual way, and then with consideration given to the
foundation. The ventilator is driven by an electric motor. Both machines have
rolling-element bearings and the coupling is flexible. The foundation frame is
of welded steel plate construction and is supported on springs.
Usual calculation. Usually natural frequencies and hence critical speeds
are calculated for the two shafts and for the foundation with an assumed stiff
machine casing.
We begin with the ventilator rotor. For this purpose the model of Fig. 20.6
was assumed. The shaft was made up of six beam elements. The model of the
rotor wheel corresponds to Fig. 16.7, in which the stiffness between hub and

Fig. 20.6. Model of ventilator rotor.

BThe drawing and data were kindly presented by Schiele GmbH of Eschborn.
300 20 Rotor and Foundation

5000
cpm

4188
4000

3000

24B1-

2000 f3
1752- fz
n3= 2074

1000
f,
n,=1216 n z=1952
0
0 1000 2000 rpm
no
n-
Fig. 20.7. Natural frequencies of rotor in Fig. 20.6.

rotor was assumed to be kcp = 2.09 X 10 7 Nm. The stiffness of each bearing was
taken to be 3 x 108 N 1m, both horizontally and vertically. The total mass of
the rotating assembly was 1800 kg, of which 1080 kg was allocated to the rotor
wheel. The model has thirty degrees of freedom. Calculations gave the natural
frequencies shown in Fig. 20.7. Their behaviour through the speed range is as
discussed in Chap. 4. For the operating speed no = 1480 rpm the following
natural frequencies are found.

1117, 1934, 2106 and 4188 cpm.

The critical speeds are

nl = 1216 rpm = 0.82 no


n2 = 1952 rpm = 1.32 no
n3 = 2074 rpm.

The natural frequencies can be easily found experimentally when the rotor is sta-
tionary and were used for checking purposes with the calculated values 1752 cpm
and 2481 cpm for n = O.
The first natural frequency of the motor rotor is 5530 cpmj the higher orders
are not of interest here.
The foundation was calculated as a rigid body as shown in Subsect. 19.1.2.
The mass of the foundation frame with machine assembly totals 15420 kg.
The support consists of 20 springs each having a stiffness of kl = 0.7 X 106 N I m.
20.2 Results 301

For calculation purposes for this study, a value of hi = 4.9 X 106 N/m was
employed and ht = 0.5 hi was assumed.
The calculated natural frequencies are:

/1 = 344 cpm /4 = 745 cpm


/2 = 417 cpm /5 = 918 cpm
h = 643 cpm /6 = 1000 cpm
The highest of these natural frequencies lies 32 % below the operating speed
and so the foundation is still sufficiently low-tuned using the stiffer springs.
Detailed calculation. In the following calculation we shall investigate

- The influence of the coupling between the ventilator and the motor shaft.

- The influence of foundation flexibility.

- The behaviour of the coupled system of machine and foundation.

The first requirement is to complete the machine assembly by adding the motor
shaft to the ventilator model of Fig. 20.6. The coupling was assumed to be
flexible and the motor bearings to be stiff. The masses are
ventilator shaft 1800 kg
motor shaft 1010 kg
total rotating assembly 2810 kg
The number of degrees of freedom is increased from 30 to 52.
Coupling of the two shafts showed only a slight influence in the frequency
range of interest. The natural frequencies differ only marginally from the values
for the individual shafts.
For the next requirement - the influence of flexibility of the foundation -
a model was assumed, consisting of the flexibility-supported base, the bear-
ing pedestals and the ventilator and motor casings. The base and the bearing
pedestals were modelled as beam elements and the casings were assumed to be
rigid bodies, connected to the rest of the assembly by beam elements. It is clear
that this is a very rough representation, which permits some arbitrary choice.
This is particularly true of the casings. To appreciate the range of possibili-
ties, the stiffness of the beams carrying the casing masses was assumed to be
multiplied by either 0.2 or 5.
The model has a mass of 13270 kg and 138 degrees of freedom. The calcu-
lated natural frequencies are:

/1 = 374 cpm /4 = 784 cpm


h = 446 cpm /5 = 940 cpm
h = 655 cpm /6 = 994 cpm
and h = 1111, 1235, 1274 cpm
/8 = 1667, 2272, 2708 cpm ,
302 20 Rotor and Foundation

for h and i8 the first and third values refer to stiffnesses multiplied by 0.2 and
5 respectively.
The first six natural frequencies correspond to the natural frequencies of the
rigid body model, and the values differ only slightly from those found previously.
The flexibility of the model first manifests itself in h.
The large variation of stiffness produced by the factors 0.2 and 5 gives rise
to only a slight change in h while for i8 the differences are -27 % and +19 %
respectively.
To investigate the behaviour of the coupled system the models described
were coupled together. The mass totals 16080 kg and the model has 190 degrees
of freedom.
The natural frequencies depend on speed. The first twelve which occur for
the operating speed no of 1480 rpm are:

il = 343 cpm h = 1105 cpm


i2 = 406 cpm i8 = 1191 cpm
fa = 620 cpm fg = 2020 cpm
i4 = 733 cpm ilO = 2170 cpm
i5 = 883 cpm in = 2269 cpm
i6 = 994 cpm i12 = 2407 cpm
To give an overview of these natural frequencies they are displayed in Fig. 20.8.
This figure shows at A the natural frequencies of the coupled rotors and at B
the natural frequencies of the foundation with bearing pedestals and casings.
At A + B is shown the natural frequencies of the coupled combination. The

,,
A
t 23

B
t " ,,
1 3
IIJ;aJ
6 7
~~~
8

A+B
t II
1
, , , , II
3 678
,
"'
910 12
11

!
0
I
1000
I
2000
I
3000 cpm
no
fk -
Fig. 20.S. Example 20/1. Calculated natural frequencies.
A Ventilator and motor shaft coupled
B Foundation with pedestals and casings
A + B Systems A and B coupled.
The values in A and A + B relate to the operating speed.
20.2 Results 303

natural frequencies are numbered according to their size. Thus some similar
natural vibrations have different order numbers, which can be identified with
little trouble from the mode shapes. We shall not consider this in any great de-
tail, but answer the main questions using the following results from calculations
of unbalance vibration.
For these calculations it was assumed that there was a mass eccentricity
of 30 I'm at the impeller, corresponding to an unbalance of 32.4 kgmm. To
restrict the resonance some damping was assumed to exist at the impeller.
To take a sensible value, relation (2.19) for a simple oscillator was assumed.
With the assumption D = 0.03, m = 1080 kg and Wn = 200 rad/s, a value of
d = 1.3 X 10 4 Ns/m was found to be appropriate. The horizontal and vertical
amplitudes of vibration of the shaft at the impeller and at the second bearing
were determined.
Fig. 20.9 shows the amplitudes of vibration of the coupled rotating as-
sembly without the foundation. The amplitudes are equal in horizontal and
vertical directions, because the bearings were assumed to be isotropic. In the
frequency range considered only one resonance appears, having a frequency
n2 = h = 2034 cpm, being in a forward mode. From the half power points of
each response curve the damping ratio would be 0.018 or 0.020. This compares
with ,the previously assumed value of D = 0.03.
Fig. 20.10 shows the amplitudes of the coupled system of machine and
foundation for which the same values were used for the unbalance and damping
at the impeller. In addition damping having d1 = 1.2 X 10 4 Ns/m and dt =
0.5 d1 was assumed in parallel with the foundation springs. Fig. 20.10 shows the
horizontal and vertical amplitudes of vibration of the shaft at the position of the
impeller and at bearing number two. The highest resonances are horizontally at

600
" 11 m
510
X
Impeller
400

200
111
04-----~~~~~~-
o 1000
n--
Fig. 20.9. Vibration amplitudes of the coupled rotating assembly without the foun-
dation.
304 20 Rotor and Foundation

t
Impeller

t
600 11 m
f f
,.
x 11 m S82 x 300 342
400 2028 2170
200

~
f9

200
100

0 0
0 1000 no rpm 3000 0 1000 no rpm 3000
n- n---
fk-

t II 111111
1 3 68 9
I III
12 t II 111111
1 3 68
1111
9 12

Bearing 2

t
f
11 m 11 m
,.f ,.f
f
66

T
10
186
x 150 x 60
2028 2170
100 40

~
f9
f7
50 20 f4

0 0
0 1000 no rpm 3000 0 1000 n rpm 3000
0
n- n-
Horizontal Vertical
Fig. 20.10. Vibration amplitudes of coupled system of machine and foundation.

n = 2028 rpm corresponding to f9 and vertically at n = 2170 rpm corresponding


to flO. It is fairly obvious that these resonances are as depicted in Fig. 20.9 and
are recognizable as shaft resonances. It is worth noting that the peak amplitudes
of Fig. 20.9 lie between the corresponding values of Fig. 20.10. These values are
at the impeller 510 I'ID, 582 I'ID and 342 I'm, respectively and at the bearing
20.2 Results 305

111 I'm, 186 I'm and 66 I'm, respectively. The weak resonances at It and /4 are
so-called rigid body resonances. Resonances hand /12 arise from the flexibility
of the foundation and its construction.
The results of the calculations given in this example can be summarized as
follows:

- The operating speed is free from resonance.

The main resonances are the shaft resonances 19 and 110. They differ only a
little from resonance b for the shaft model.

For the coupled model of machine and foundation there appear near the shaft
resonances further weak resonances, due to the foundation.

- In this example the usual simple calculations are sufficient.

As a result of this example and many similar calculations the following


general conclusions can be drawn.
There is no doubt that it is sensible in some cases to determine more ac-
curately than usual the vibration responses of machines which have here been
described as small. Appropriate computer programs are available for such pur-
poses.
In most cases, however, simple calculations are sufficient as for example
when the foundation is relatively heavy and rigid or when the natural frequen-
cies are situated a long way from the running speed. A simple calculation is
also sufficient when one is preparing to modify the machine, if this should be
necessary.
Conversely a comprehensive calculation has to be considered when,

the foundation frame has to be as light and cheap as possible,

a larger number of machines of the same type are planned,

- the vibrations have to be especially low, for example when the machines
operate in living or office accommodation,

and in similar circumstances. If in doubt it is recommended to undertake a


comprehensive calculation, since future cases can be judged better from such
results.

20.2.2 Large Turbine Sets

Large turbine sets are supported on table foundations or spring foundations.


The question as to whether to treat the models as coupled or separate is more
306 20 Rotor and Foundation

important here than for the small machines considered in the previous subsec-
tion. This is because the two subsystems have many natural frequencies in the
significant frequency range and furthermore they usually lie close together.
A coupled model is required if it behaves significantly differently from the
subsystems. Differences can occur in respect of

- natural frequencies and hence critical speeds,

- resonance amplitudes and

- stability borderlines.

These questions were investigated amongst others in references [117, 124, 125,
128, 130, ... 133]. The results and consequences of these and other unpublished
work are summarized in Sect. 20.2.3. Details are elucidated in the following two
examples.

Example 20/2 We consider the question of the vibration at resonance and


the change of peak value due to coupling a turbine with its foundation. For
this purpose we consider the turbine assembly shown in Fig. 20.11. This was
investigated in references [117] and [124].
We begin with the rotor. This was built up of a string of shafts with fluid-
film bearings and bearing pedestals as shown earlier in Fig. 14.5. 153 beam
elements were first used and then reduced to 22, giving 106 degrees of freedom.
The mass of the shaft system amounts to 81 t and for the bearing pedestals
17 t. The damping of the fluid-film bearings was taken into consideration, but
no further damping was allowed for.

27m

Fig. 20.11. Turbine assembly for example 20/2. P = 125 MW, n = 3000 rpm.
-----
20.2 Results 307

Pd~+
fk=n
---14
""-- 13 1 1 1
fk 3000
/
cpm 11 /

--- A
~
/ 9 1 ~ 11+
2000 I

11-+
h5
I
I
I
I I I
1000 -t:-2

I
I 1
/ / ~I 1 I~~
/ k
a
a 1000 2000 3000 rpm
n-

~~
Dk 0.05 9-<>- '-

5
10 1
2
a
a 1000 2000 3000 rpm
Fig. 20.12. Example 20/2. Natural frequencies, damping ratios and mode shapes of
the rotor model.

Fig. 20.12 shows the calculated eigenvalues and mode shapes. Results for
heavy damping were omitted. Thus cases of Ik are not shown where Dk > 0.08
and mode shapes are not shown for Dk > 0.03. Using this model we must expect
to run near resonance during operation, as 113 is at 3093 cpm only 3.3 % greater
than the operating speed of 3000 rpm.
The dynamics of the foundation have already been considered in Chap. 19,
example 19/2. The foundation is made of steel-reinforced concrete and has
a mass of 1210 t. The chosen model has 218 degrees of freedom. Especially
noteworthy is the close collection of natural frequencies - up to a frequency of
3300 cpm there exist 35 natural frequencies.
On combining model (R+F) from rotor (R) and foundation (F), the casings
were considered as rigid bodies with estimated masses.
We shall dispense with a comparison of natural frequencies and consider
only critical speeds.
Fig. 20.13 shows these for models Rand R + F and as in Fig. 20.12 only
values for Dk < 0.08 are given. A measure of the severity of the resonance is
the inverse of the damping ratio. We introduce the expression
1
Rk = - (20.14)
2Dk
308 20 Rotor and Foundation

Genv HP v LP v Gen 'V MP v


1 2 5 9 1011 1314
R

R+F
7 10 13 17 19 35 41 43 50 54
23 36 51
80 nB
60 T I

1-- ~r-r If-I~t-r


40 I
Rk
20

10 ----
I
6
4
Ti:
I
I
I
I
I I

~ I I I
0 1000 2000 3000 4000 rpm
nk -
Fig. 20.13. Example 20/2. Critical speeds and resonance factors.

Model R
Model R+ F

which corresponds to the Q-factor after equation (2.37). It is shown in the lower
part of Fig. 20.13 on a logarithmic scale, dashed for model R and in full lines
for model R + F.
The connecting lines between the critical speeds for model R and model R+
F serve to show which resonances correspond. The problem of assigning numbers
to the resonances has already been addressed. We shall not go any further into
a comparison of resonance factors, but consider straightaway vibrations due to
unbalance.
In order to calculate unbalance vibrations proportional damping in the foun-
dation in addition to damping from the fluid-film bearings was assumed. The
factors a and f3 were determined so as to give a damping ratio of 0.02 for
n = 1200 and 3000 rpm.
The unbalance vibrations depend upon the value and distribution of un-
balance. To judge these vibrations we usually show the amplitudes of vibration
in horizontal and vertical directions at the nodes. Another more elaborate way
is to show the ellipses of vibration at individual nodes. In either case there is
a great deal of information and hence the problem of coming to a considered
conclusion from the many results.
Numerous cases were calculated for our example. These and their results
are described in references [117] and [124]. We shall confine ourselves here to
20.2 Results 309

the question as to whether the apparently serious resonance /13 for the rotor
model is also dangerous for the combined model.
For resonance /13 the generator rotor vibrates in an S-form, and mainly
in the vertical direction, as can be seen in Fig. 20.12. In order to excite this
resonance, a pair of unbalances was assumed to act on the generator, having
values corresponding to a 10 I'm eccentricity of a half-rotor mass.
The most important result is shown in Fig. 20.14, in which the vertical
vibration amplitudes at bearings 5 and 6 are displayed. The resonance of the
rotor model is shifted by about 10 % and drastically reduced as a result of the
coupling with the foundation. Further results are contained in the summary in
Subsect. 20.2.3.

Example 20/3 As to the question of the influence of the foundation on the


stability borderline, we consider the results of reference [132) as an example.
In this reference a turbine set was investigated which has P = 315 MW and
n = 3000 rpm. The rotor of this set has already been considered in Chap. 17, as
example 17/3. The turbine set is supported on a spring foundation measuring
32 X 12 x 3 m. Amongst other investigations the onset speed for oil-film insta-

HP Gen
I
2
I
3 4
I I
5
!
I I
6 7

1351

Ilm Ilm II
f 30
f 30
R
A
A
X X R
20

10 J!!+F
20

10
ljR+F
0 0 I I I I f
0 2000 3000 rpm 0 2000 3000 rpm
n- n-

Bearing~ Bearing~

Fig. 20.14. Example 20/2. Amplitudes of vibration at bearings 5 and 6 in the vertical
direction as a result of the application of an unbalance pair.

R Rotor model
R +F Model of rotor and foundation
310 20 Rotor and Foundation

bility and the performance at the boundary of instability for steam whirl were
determined for different models. The assumed models were

A Rotor with fluid-film bearings and rigidly supported bearing shells.

B Rotor with fluid-film bearings and bearing pedestals as shown in Fig. 16.12
(our standard rotor model).

C Rotor as in B, but coupled to the foundation.

First consider the boundary for oil-film instability. The significant damping
ratios are plotted against speed in Fig. 20.15, the stability borderline occurring
for D = O. The speeds at the onset of instability (nth) are seen to be
5310 rpm for model A
5880 rpm for model B
4290 rpm for model C.
These onset speeds show the strong influence of the model on predicted
stability. The trend in this example is not uniform. The more realistic model B
(as compared with model A) has a higher stability--onset speed, but the even
more realistic model C has a lower onset speed than does model A.
For steam whirl the system will become unstable at a certain power
(Chap. 11). In reference [132] steam whirl was assumed in the HP portion.
To determine the power at the boundary of stability the degree of steam whirl
was varied and the eigenvalues of models A and C were calculated. This showed
that the significant real parts, that is damping ratio Ds for model A and D I9 for

0.003
n-

3000

-0.003
4290 5880
{
nth 5310

Fig. 20.15. Example 20/3. Relevant damping ratios for stability (models A, B and
C).
20.2 Results 311

model C had a very flat behaviour over the power range, so that the boundary
value could only be roughly determined. But it could be recognized that the
boundary power for model C was essentially smaller than for model A. The
ratio lay between one and two.
The conclusion from this example is that the stability borderline depends
a great deal on the model chosen. This statement is supported by practical
experience, and often only trivial changes to the machine or its bearings will
have a significant influence on stability.

20.2.3 Conclusion

The problem of a suitable choice of computer model depends on the objective


and is very complex. We confine ourselves to the machines considered here
and the following statements are concise and simplified. Exceptions and other
numerical values are also possible.
For small machines having frame foundations one can conclude:

The usual separate calculations for the machine and the foundation are mostly
sufficient.

In special cases - some are mentioned in Subsect. 20.2.1 - it is necessary to


calculate for the coupled system.

If the machine has a range of operating speeds, and if we suspect after some
rough calculation that natural frequencies can occur in this range, then we
should calculate unbalance vibrations using the coupled model. Possible
positions and severities of resonances can be found by assuming different
models and unbalances, so that we can judge whether these resonances
are acceptable.

For large turbine sets, investigations have shown that:

The coupled model of rotor and foundation has many natural frequencies, of
which only a few lead to noticeable resonances. Which natural frequencies
are strongly excited depends on the unbalance distribution.

The natural frequencies of the coupled system are above or below those of
the uncoupled model. Their ordering is possible with a few provisos by
comparison of mode shapes. The variation of natural frequencies due to
coupling can be assumed to be a maximum of 12 %, for machines with
a steel-reinforced concrete foundation.

The peak values of the resonances are usually lower for the coupled system
than for the rotor model.
312 20 Rotor and Foundation

After investigation the following recommendations can be made for vibration


calculations:

Fluid-film bearings and bearing pedestals should be considered in the model


used for calculating rotor vibrations. This model suffices in most cases,
if the turbine set is supported on a heavy foundation made of steel-
reinforced concrete and if some experience has been acquired in its design.

A model coupled with the foundation should be used for new designs, if the
rotor model alone suggests the presence of resonances. When employing
a steel foundation one should preferably use a coupled model.

Finally it should be noted that, by using the coupled model it can be as-
certained whether the foundation can be made lighter, as far as vibrations are
concerned. In practice this assessment is usually made for steel foundations. It
would also be interesting to consider this for reinforced concrete foundations,
because these are often over-designed.
Sometimes a machine and its foundation are not mounted on the ground but
on a floor of a building. If there is some doubt as to whether separate vibration
calculations suffice, then one should determine the forced vibration response
for the coupled system in a simple way by using equation (20.7). The dynamic
characteristics of the building are completely catered for by the matrix KA(W),
that is by the dynamic stiffnesses for the connecting coordinates. These can be
found by separate calculation or by measurement.
Finally, something should be said about the accuracy of present-day vi-
bration calculations in comparison with reality. The question can be posed as
follows:
For values of
- natural frequencies,
- resonance peaks and
- stability borderlines
how big are the differences found by calculation from those found by measure-
ment on site?
Let us begin with stability. As has been already remarked, even with today's
calculations we often cannot account for all possible influences or cannot achieve
results with sufficient accuracy. The calculated values for the stability boundary
can therefore differ markedly from values found in practice. The real use of
calculated stability criteria consists essentially in the determination of trends
and has little to do with precise numerical values.
In order to achieve a comparison between calculated and actual resonance
behaviour it is obvious that vibrations at the same spatial position 'should be
selected. In practice this is often difficult. Furthermore the values and distribu-
tion of unbalance should be the same. This requirement is again very difficult
to realize in practice. A possible compromise is to use the results of trial unbal-
20.2 Results 313

ances. Further details on this problem are discussed in reference [117], where
four turbine sets were thoroughly investigated, having outputs of between 125
and 660 MW. It was found that the calculated speeds for resonance in the main
differed less than 5 % from measured values. The damping assumed in [117] -
listed values for oil-film bearings and D = 0.02 for the foundation - gave more
pronounced resonances than those measured.
The accuracy of the calculated resonance speeds are unlikely to be improved
in the near future. However, the peak values can be more accurately computed,
if more realistic damping values are assumed.
Part IV

Further Features
21 Rough Calculations

The desire to calculate the important quantities of system dynamics in a sim-


ple and rapid manner still exists in spite of the increasing use of computers.
Amongst others the following are especially desirable:

- Natural frequencies and critical speeds

- Resonance amplitudes

Stability boundaries

Stresses.

Above all, it is essential to make rough calculations at the beginning of a project,


to act as a check on computed results and to clarify unusual phenomena likely
to be encountered during running.
In the following two sections, reference to the approximate calculation of
natural frequencies and resonance amplitudes will be made. Stability boundaries
can also be determined but only approximately even by computer, because only
inaccurate values of excitation and damping coefficients will be available. To
estimate stability parameters, the contents of Chapts. 7, 10, 11 and 12 will be
utilized. To estimate stresses in rotors, use will be made of the statements given
at the end of Chap. 22.

21.1 Natural Frequencies and Critical Speeds

To calculate the natural frequencies of the systems considered here, a computer


is needed. However, one usually wishes to know first of all the value of the lowest
natural frequency ft. This can be calculated reasonably accurately for rotors
with two bearings without a computer, as will be shown in this section. From
known values of /kl ft, the value of the higher natural frequencies /k can also
be calculated, at least approximately.
Critical speeds nck are obtained when the rotational speed coincides with a
natural frequency Jk:
(21.1)
This is true for steady operating conditions. During acceleration or deceleration,
the critical speeds are displaced, as is described in Sect. 3.2 for the Jeffcott rotor.
318 21 Rough Calculations

This effect is, however, only noticeable for very rapid acceleration. From Fig. 3.9
for example, a difference in resonance of only about five percent is obtained, if
the resonance is reached after one hundred revolutions from stand-still.
For the leffcott rotor

where
k = 48 EI and I = 7r d4
P 64
Thus

11
/E2d
= 0.244 V;;z3 . (21.2)

With units N, m and kg, 11 is in cycles/so The influence of l, d and m on 11


is illustrated in Fig. 21.1. Here units were chosen so as to conform to the usual
notation found in practice.

f 100000
fl cpm

10000

1000

100

0.1

10

Fig. 21.1. Natural frequencies of Jeffcott rotor.

E = 20.6 X 10 10 N/m 2
21.1 Natural Frequencies and Critical Speeds 319

A real rotor with two bearings can be assumed to approximate to a Jeffcott


rotor, if its mass is concentrated towards the centre of the rotor and about
midway between the bearings. For a mass centre lying off centre or out board of
a bearing, the models I, II and I I I of Fig. 4.1 can be used as approximations.
They have two natural frequencies at zero speed, as was indicated in Sect. 4.2
with suffixes 20 and 40.
The natural frequencies normalized by fo = wo/27r are

and

and are obtained from equation (4.27) and from Figs. 4.13, 4.14 and 4.15. If one
takes fl of the Jeffcott rotor as a reference frequency, then

(21.3)

The ratio fo/ fl = WO/WI is given as follows, with Wo from Table 4.3, and WI =
/48 EI/mP.
1 12
for model I
4 ab
fa 11{;f for model II (21.4)
fl 1+ c
4 c

~ Gf /2 for model III.

When rotation takes place gyroscopic moments come into play and four natu-
ral frequencies result, two forward and two backward whirling. They arise as
solutions of the characteristic equation (4.18) or can be deduced from Figs. 4.6,
4.10, 4.11 and 4.12.
Critical speeds in forward whirl are obtained from equation (4.38) and from
Figs. 4.20 to 4.22. In these figures the mass has been assumed to be in the
form of a thin disc for which Id = 0.5 Ip. For other ratios of I d/ Ip, the natural
frequencies are given by the above equations together with the equations of
Chap. 5, for the short and long rotor.
If the mass of a rotor is not lumped, the first natural frequency may be
computed with good accuracy considering energy terms, which will be shown
in the following.
For a rotor with arbitrarily distributed mass and stiffness (Fig. 21.2), having
natural frequency Wk and natural mode shape Yk(X), the equation of natural
vibration is
(21.5)
320 21 Rough Calculations

Fig. 21.2. Calculation of WI by Eq. (21.11)

At Yk(X,t) = 0 the kinetic energy is a maximum

Tmax 2
= 1Wk2 J1'( )2 x )dx
X Yk( (21.6)

where 1'( x) is the mass per unit length.


Equation (21.6) may also be written as

(21.7)

or
(21.8)
where

(21.9)

The right hand side of equation (21.8) is the work of the maximum inertia
forces, because ak( x) is the function of maximum acceleration. Equating (21.6)
and (21.8) the maximum kinetic energy is equal to the work of the maximum
inertia forces and this is equal to the maximum potential energy Urnax Thus
Trnax = Urnax , which is the theorem of energy for natural vibrations of undamped
systems.
For the first natural frequency (k = 1) we now assume
ak( x) = a, where a is an arbitrary constant acceleration,
YI (x) = y( x), where y( x) is the static deflection due to the loads
dF = a dm.
Hence equating (21.6) and (21.7) one obtains,

(21.10)
21.1 Natural Frequencies and Critical Speeds 321

Replacing the integrals by sums finally yields

WI = aJ a ,
Ymax
(21.11)

where
EmiYi
a= '" Ymax
(21.12)
Yi2
L.J mi

and
Ymax = max(Yi) .
It is found that equation (21.11) is a good approximation. The error is usually
lower than one percent.
With a lumped mass, stiffness k = F/y at the centre of the mass, a = g
and Ymax = mg/k = Y.,
the static deflection due to weight W = mg. Putting
a = 1 in equation (21.11) one obtains the well-known formula

= {i . (21.13)
yY.
WI

If the deflection Ymax due to the loads Fi = ami is known, the first natural
frequency It may be found from equation (21.11) with sufficient accuracy by
a
assuming a value of with the aid of Table 21.1. Then It = nl = wd27r.
The natural frequencies of rotors with several bearings should preferably be
determined by computer. From the mode shapes one usually finds that a certain
correlation is possible with a component rotor. For example, in Fig. 17.18 the
first to the fourth coupled vibrations correlate with the individual mode shapes
of the Gen, LP, M P and H P rotors respectively. The ratios of the coupled

Table 21.1. a-values after Eq. (21.12)

c
IS 0 h 1

102 0
0.54
!02~1
1.02

Machine rotors 1.04 +1.11

A A 1.13

~ 1.24
322 21 Rough Calculations

natural frequencies to the corresponding individual natural frequencies in this


example are
Gen 917/805 1.14
LP 1576/1347 1.17
MP 1906/1506 1.27
HP 2604/2160 1.21 .
If such ratios are known from numerous calculations for fairly similar rotors,
then for a new design one can estimate the lowest natural frequency of its
coupled rotors, using the rough values of II for the uncoupled component rotors.
Also, one can obtain a rough check on the results of a computer calculation.

21.2 Resonance Amplitudes

To calculate approximate values of maximum displacement amplitude at res-


onance first consider the oscillator with one degree of freedom. From Chap. 2
equation (2.37), it was found, with sufficient accuracy, that

(21.14)

where z. is the static displacement due to a static force equal to the amplitude
F of the excitation force.
Correspondingly, from Chap. 3, for unbalance excitation due to U = me it
was found that
~ U 1 1
z max -
-
--
m 2D-- e2D'- (21.15)

For models with several degrees of freedom one obtains as an approximation for
the amplitude at resonance an analogous relationship, using a shortened modal
calculation, as shown in the following.
For excitation at coordinate k by a force Fk cos wt the amplitude of the
displacement Zi is given by
(21.16)

The receptance hik(W) is found using a modal calculation for a model with n
degrees of freedom, as the sum of n terms. At resonance, hik can be approxi-
mated by equation (24.60), using certain assumptions. For w = W r , where Wr is
a natural frequency,
(21.17)

where m*r is the generalized mass


are components of the eigenvector CPr
and is the modal damping ratio.
21.2 Resonance Amplitudes 323

For excitation by an unbalance Uk at coordinate k of a rotor for which n = Wr,


the excitation force is

and so
(21.18)

For example consider the simple case of a cylindrical shaft with two bearings
and a single unbalance U at its centre and let us find the maximum vibration
amplitude at the shaft centre for the first natural frequency.
At the first natural frequency the shaft has the form of a half sine wave.
With !Pil = !Pkl = 1 and the shaft mass m = pAl, the generalized mass is
mi = 0.5 m and so, from equation (21.18),
~ U 1
Zmax ~ 2 . m . 2Dl . (21.19)

If the mass is concentrated at the shaft centre, equation (21.15) is valid and the
amplitude is only half as big as for a uniformly distributed mass. This can be
explained by the fact that, in the second case, the natural frequency and the
unbalance force are smaller than in the first case. With

(wdc = )48mP
EI for a concentrated mass

and
{Ei
(wd d = 9.871y;;;p for a distributed mass
the ratio of unbalance forces is
48
9.871 2 = 0.493 ~ 0.5 .

To calculate the resonance amplitudes using equations (21.17) or (21.18), the


assigned modal values must be known. In cases where they are not available,
they can be estimated from values obtained by earlier calculations. One can,
however, also use the relationships

(:cik )max = C

Z 1
2Dr
resp. _ C Uk _1_
- u m 2Dr
(21.20)

where z. is the static displacement at k due to Fk and m is the reference mass,


for example the mass of the shaft in question. It is assumed that one has from
experience suitable values of C. and hence Cu. One can obtain these from back-
calculation using computer results or from measurements.
324 21 Rough Calculations

The usefulness of these approximation formulae can only be assessed by


carrying out parallel calculations. As a rule good results can be obtained for
isolated, well defined resonances.
22 Bending Stresses in Rotors

The solution of the equation of motion consists in finding the displacements at


the nodes of the structure. With these displacements, the internal forces and
moments are also ascertained. These latter have, like the displacements, space
fixed coordinates. The stressing of a rotor consists in finding the stresses in its
cross-section, whose time-dependent behaviour follows from the displacements
in a coordinate system fixed to the rotor. Thus a coordinate transformation is
necessary, which can be done effectively by using complex quantities.
The rotor consists of beam elements, whose mechanics are described in
Subsect. 16.2.2. With the node displacements as shown in Fig. 16.9,

then, using the element matrices the internal forces and moments are

fs = =Mi:+(D+G) x+Kx. (22.1 )

The matrices M, G and K follow from equations (16.34, 16.38, 16.31) and for
proportional damping, the damping matrix is,

D = aM +,BK.

Let us now consider the left-hand cross section of the beam element of Fig. 16.9
having bending moments M3 and M 4 With axes Re and 1m representing sta-
tionary coordinates and Re*, 1m* coordinates fixed in the rotor, having angular
velocity il, the complex bending moments are, from Fig. 22.1,

M = M 3+J. M 4
_ M*= M*3+J. M*4=_e
an d _ M -jilt . (22.2)

The stationary bending moment comprises a constant and a harmonically vary-


ing term
(22.3)
326 22 Bending Stresses in Rotors

1m

~0*

Re

Fig. 22.1. Moments of the rotor cross-section.

where
M3(t) = M3c cos wt + M3 sin wt
(22.4)
M4(t) = M4c cos wt + M4.sinwt.
The alternating moment M(t) can be considered as the sum of a forward and
a reverse moment, as discussed in Sect. 17.1.

(22.5)

where

M' = ~v'(M3C + M4.)2 + (M4c - M3.)2 ,


(22.6)
Mil = ~v'(M3C - M4.)2 + (M4c + M3.)2 ,

If one further puts


(22.7)
where
M 40
{3 = arctan--
M30
then, using equation (22.2) the required moment for the rotating system is

(22.8)

The cross-section thus experiences

- a primary moment of amplitude M o, changing at frequency n


22 Bending Stresses in Rotors 327

a forward-rotating moment at frequency (w - ill and amplitude M' and

a reverse-rotating moment at frequency (w + ill and amplitude M".


For unbalance excitation w = il, the moment
(22.9)

The forward-rotating portion is constant in this case and the reverse-rotating


portion has a frequency equal to twice the rotating frequency.
U sing these results, one can carry out a stress analysis for a given rotor. We
shall do this in the following, by estimating the stress due to deflection in the
rotor of Fig. 22.2 as a result of gravity loading and unbalance loading.

A B
I I I I
w
~-----~
I I
Fig. 22.2. Example of stress estimation.

The gravity load gives a displacement w in plane A of Fig. 22.2, and the
unbalance forces give the elliptical orbit, having half-axes a and b. The stress
in the rotor in any plane B arises from moments Mo, M' and M" (equation
(22.9)). With W as the section modulus of the cross-section, the maximum
values of bending stresses are

Mo M'
un= W ' Uo = W ' (22.10)

in which the suffix indicates the frequency associated with the stress. In general
these values relate to different positions of the cross-section, which for our
estimate is unimportant. We assume as an approximation that the moments at
B behave like the displacements at A, and so

Mo : M' : M" = w : R' : R" . (22.11)

For forward whirl, from expressions (17.6) we have,

, a+b b
R =-- R " =a-
--
2 2
328 22 Bending Stresses in Rotors

and for reverse whirl,

R,_a-b R"=a+b.
- 2 ' (22.12)
2
Finally, using expressions (22.10) the relation between the stresses is

un : Uo : U2n = w : R' : R" . (22.13)

Bending stresses are especially noticeable in heavy slender rotol;s, for which the
following comments can be made:
In commercial operation the size of the elliptical orbit, that is the values of R'
and R", are much smaller than the static deflection w, so that, from relation
(22.13), Uo and U2fJ are correspondingly much smaller than Un. In other words,
in service, alternating stresses due to gravity loading dominate.
In passing through a resonance the dynamic deflections, that is R' and R",
achieve values of the order of the displacement w. Hence the stresses Uo and U2fJ
also have values of the order of un. Nevertheless, for resonance-free operation,
the life of heavy slender rotors depends essentially on Un, since high values of Uo
and U2fJ only occur during run-up and run-down, that is during comparatively
short periods of time.
23 Cracked Rotors

Cracks in rotors are quite rare, but they must be considered very carefully.
Cracks can occur in old rotors with long life histories, but sometimes they also
occur in new rotors after only a few hours of operation. If the crack increases
to the extent that the bending stiffness is altered (a transverse crack) and if it
has the requisite depth, then it will influence rotor vibration. Details relating to
this will be considered in Sect. 23.1. With present knowledge and corresponding
measurements, one can almost always recognize a transverse crack in time to
avoid a catastrophic failure.
In Sect. 23.2, we shall summarize information taken from various reports
on the subject of rotor cracks. The last section will be concerned with a concise
set of rules by which cracks can be diagnosed and/or avoided.

23.1 Vibrations of a Rotor with a Transverse Crack

Since about 1976 investigations of the dynamics of rotors with transverse cracks
have been discussed in the literature [134, ... 153]. They vary in the type of
rotor model, the crack model and in the method of solution. The following are
essentially a repeat of the results of Schmied [153] and represent the current
state of knowledge on the subject.

23.1.1 J effcott Rotor

The Jeifcott rotor of Fig. 23.1 has a crack at the disc position. Without a crack
its equation of motion would be

[ : :]{::} + [: :]{ :: } + [: :]{:: }

= men 2 { cos(nt + ex) } _ { o} (23.1)


sin(nt + ex) mg
330 23 Cracked Rotors

r
X2

~cm'k
J}l; k Xl

I. .I
Fig. 23.1. leffcott rotor with a transverse crack.

Because of the crack the stiffness of the shaft IS reduced and is also time-
dependent when the shaft rotates. Thus

K -- [ok Ok] becomes K = K(t) .

To determine K(t) consider the shaft shown in Fig. 23.2. The crack takes the
form of a segment of a circle, is infinitely thin and of depth a.
Under compressive load the shaft behaves as ifit has no crack. This is termed
a closed crack. When the crack is open it is treated as in Fig. 23.2, with the
shaded area only being operative. In general during running the shaft constantly
changes between these two extremes and the crack is said to be breathing.
When the crack is open the bending stiffness is reduced. Let the maximum
stiffness be kl and the minimum k 2 These stiffnesses will depend on the depth
of the crack as Fig. 23.3 shows. Both decrease progressively with depth. For
depths of a < 0.1 r the influence is very small so that vibration characteristics
are hardly influenced. For a breathing crack the stiffnesses change between ko
and kl and between ko and k 2 The breathing action depends on the prevailing
static and dynamic forces, and the equation of motion is non-linear. Linearity is
maintained, however, if one assumes a cyclic variation of breathing. A reasonable

112 112

Fig. 23.2. Position and geometry of a crack.


23.1 Vibrations of a Rotor with a Transverse Crack 331

O~~~~~~~~-r
a
--
0.5 1.0
a
r
Fig. 23.3. Stiffnesses and their dependence on depth of crack for I = 10 r, from [153].
- from [136], - - - from [142]

assumption is harmonic breathing as shown in Fig. 23.4, having a period T =


211' / a. The stiffnesses then become

kl(t) = kml + Llkl cos at (23.2)


k2(t) = km2 + Llk2 cos at

where
1
kml = "2 (leo + kt)
(23.3)
1
km2 = "2 (ko + k2)
To a good approximation such a cyclical variation occurs in the stiffnesses of a
horizontal, slowly rotating shaft, due to its own weight. However, in reference
[153] it is shown that this representation gives useful results for heavy rotors
even at resonance. It is also valid for vertical rotors, when loaded by a radial
force, constant in direction and with time. The stiffnesses kl(t) and k2(t) are
valid for coordinates fixed in the rotor and designated as ZlO and Z20. In order
to write the equation of motion, however, one needs these stiffnesses relative to
space fixed coordinates Zl and Z2. These can be found by writing

(23.4)

ZlO} [ cos at sin at 1{ Zl } { ZI }


(23.5)
{
Z20 = - sin nt cos nt Z2 = T Z2
332 23 Cracked Rotors

~~~~~
closed open closed

ko

~1(t)
k1
k 2(t)
k2

~ I I
a T/2 T
t---
Fig. 23.4. Variation of stiffnesses for a harmonically breathing crack.

and

K(t) = T T KaT = [ km + Llk cos 2{}t Llk sin 2{}t 1 (23.6)


Llk sin 2{}t km -.ilk cos 2{}t

where
1 1
km = 2[k1(t) + k2(t)] , .ilk = 2[k1(t) - k2(t)] . (23.7)

If we take k1(t) and k2(t) from equations (23.2), we obtain after some manipu-
lation the expression,

K(t) = Kc + Kt(t) = [k:" 0 1+ .ilk' [J1(t) h(t) 1 (23.8)


o k:" h(t) /2(t)

where

k' = k 2 + "'1 + "'2


m a 4

.ilk'
23.1 Vibrations of a Rotor with a Transverse Crack 333

and
!l(t) C cos nt + 2 cos 2nt - cos 3nt
!2(t) (2 + C) cos nt - 2 cos 2nt + cos 3nt
(23.9)
fa(t) - sin nt + 2 sin 2nt - sin 3nt
4 - 3ltl - 1t2
C

The stiffness matrix of the shaft with crack thus consists of constant and pe-
riodically varying terms. The periodic terms consist of harmonic components
with frequencies n, 2n and 3n. The values of these components depend on k:",
Llk' and C. The non-dimensional values k:" I ko, Llk'i ko and C depend on kd ko
and k2/ko, which themselves depend on the relative crack depth alT. Fig. 23.5
shows this dependency for I = 10 T. It can be seen that for small crack depth,
k:" ~ ko and C ~ 1. The quantity Llk' is small in comparison with k o. For
example, for a = T, the ratio is only 5 %. Hence, especially for an initial small
crack, the periodic term in k( t) is small in relation to the constant term.
In reference [153] the matrix K(t) was also established for a non-circular
shaft with crack. It was shown that, for a breathing crack, the same rules were
obtained as for a circular shaft. For small lack of circularity any error could be
neglected. With stiffnesses ku and kv of the non-circular shaft, the following
comments can be made regarding equation (23.8).

- Instead of ko, the average value (ku + kv)/2 should be used,


- there are more terms with frequency 2n. Their coefficients are Llku cos 2{3,
Llku sin 2{3 or Llku sin{3 where Llku = (kv - ku)/2 and {3 is the angle

f 1.0-r---_
c k~

3 k. 0.9

0.06 0.8

2 0.04 0.7

0.02

o o~..:....;::::c""---'-'--r---r---.----r----.-.--.--.---
o
..Q.-
r
Fig. 23.5. Values of k'"./ko, ..::::lk' /ko and C for I = 10 r. After table 2.1 in reference
[153].
334 23 Cracked Rotors

between the major axis of the non-circular shaft and the cracked cross-
section and

- the terms of frequency [} and 3[} remain unaltered.


Having considered K(t), we now come to the solution of the equation of motion.
Using equations (23.8) and (23.1) we obtain

Mi + Di + [Kc + Kt(t)]x = f(t) . (23.10)

As already mentioned, the matrix Kt(t) is small in comparison to Kc for all but
the larger values of crack depth. Hence the following perturbation calculation
gives a satisfactory solution.
We first neglect K t ( t) and calculate from

(23.11)

the particular solution xc(t) and substitute Kt(t)x c into equation (23.10) as an
approximation to Kt(t)x. Hence we obtain the equation

(23.12)

whose solution
(23.13)
which consists of xc(t) from equation (23.11) and Xt(t) from the following equa-
tion
(23.14)
We thus obtain a final solution, which is discussed in reference [153]. The es-
sential results are shown in the figures in Subsect. 23.1.3.

23.1.2 General Rotor

We now consider a general rotor on general bearing supports, whose equation of


motion is known. If the shaft has a crack, then the appropriate beam element is
designated accordingly. It will differ from a standard element only in its stiffness
matrix. Instead of K (expression (16.31)) we have a matrix K(t), which is
obtained as was equation (23.8). For details, reference should be made to [153].
Naturally, the crack affects the stiffness of the whole shaft, but does so through
the aforementioned change of stiffness matrix. An equation of motion of the
same type as equation (23.10) is then set up. In reference [153] an investigation of
stability is first undertaken by solving the homogeneous equation. Calculations
showed that the stability is markedly influenced only by a fairly deep crack. We
shall not consider this further, as stability plays no special role in the detection
of a crack. For the determination of forced vibration, reference [153] postulates
23.1 Vibrations of a Rotor with a Transverse Crack 335

a solution of the form


N
X(t) = Xo + E (:iCer cos rnt + xsr sin rnt) . (23.15)
r=l

The unknowns Xo, Xc1, Xsll Xe2' Xs2 are obtained by substitution into the
equation of motion and hence by comparing coefficients. This procedure indi-
cates that N = 3 gives sufficiently accurate results.

23.1.3 Results

The following results come of reference [153] where a Jeffcott rotor was inves-
tigated, followed by a generator rotor. The substitution (23.15) was used to
solve the equation of motion for each model. Comparative calculations for the
Jeffcott rotor using perturbation theory gave good agreement with the results,
using series solutions.
The dynamic displacements were determined at the node positions in both
horizontal and vertical directions for a range of values of speed n = n /271". With
the assumptions used these displacements consist of three harmonics, which
we can call the n-, 2n- and 3n-vibrations and the harmonic amplitudes were
designated :e(n), :e(2n) and :e(3n). The maximum value of their summation was
defined as :e(L)' In general the three harmonics had different phase angles and
therefore
:e(E) ~ :e(n) + :e(2n) + :e(3n) . (23.16)
Jefcott rotor. A model such as that shown in Fig. 23.1 was investigated.
For all calculations 1= 10 r and the damping ratio D was assumed to be 0.05.
The crack depth was a = 0.25 r and the influences of gravity load, unbalance
and lack of circularity were investigated.
Fig. 23.6 shows the maximum values of horizontal and vertical vibration
of the disc centre as a result of gravity excitation and the amplitudes of the
harmonics.
Let us consider first the harmonics. For resonance the n-vibration occurs
at no, the 2n-vibration at no/2 and the 3n-vibration at no/3. Accordingly, the
total vibration has resonances at no/3, no/2 and no, where

no=~ Iko.
271" V;;;
(23.17)

This is the natural frequency and hence critical speed of the rotor without
a crack (the assumed crack depth has no noticeable influence on the natural
frequency). The excitation force in this case is the gravity force and the vibration
results are proportional to static displacement
mg
:e. =-. (23.18)
ko
336 23 Cracked Rotors

t 0.2xs

x([) 0.1xs x([)


0.1 Xs

0
no no no no
n-- n-
32
0.2x s
f
x(rn) O.1x s x(rn)
0.1 x s

n--

Horizontal Vertical
Fig. 23.6. Dynamic displacements due to gravity excitation. I = 10 r, D = 0.05,
a= 0.25 r .

The maximum values are, in the horizontal direction, 0.10 :c. at n = no/2 and,
in the vertical direction, 0.17 :c. at n = no.
Any static radial force, for example, a fluid force acting on the disc will have
the same effect. With F. as the resultant of the radial forces, one can in general
write
F.
:c. = 11:0 . (23.19)

"Vertical" is then regarded as the direction of the resultant and "horizontal"


the direction at right angles to this.
Fig. 23.7 shows the displacements obtained by a combination of gravity
load and unbalance. The ratio of eccentricity e to static displacement :c. is a
significant parameter here and in Fig. 23.7 it was assumed to be 1/28.6 = 0.035.
The largest resonance occurs at n = no and the position of the unbalance
in relation to the crack has a strong influence on the peak value. This is 15 e
in the vertical direction, if the unbalance lies on the same side as the crack and
on the axis of minimum stiffness. Its value is 5 e, if the unbalance lies on the
opposite side. For other positions of the unbalance the peak values lie between
these two extremes. Without a crack the peak value would be 10 e (the dashed
line). In the horizontal direction the differences are smaller.
For exclusive gravity excitation the peak value :c(E) occurs at n = no and
has a value 0.17 :cs = 0.17/0.035 . e = 4.9 e. This shows that the peak value is
23.1 Vibrations of a Rotor with a Tranl!verse Crack 337

t t
x(r) x (I:)
10e
10e

O~--~r----.-------- O~--.-~~--.--------
n_ n_

Horizontal Vertical
Fig. 23.7. Dynamic displacements due to unbalance and gravity load. e = 0.035 2.,
1= 10 r, D = 0.05, a = 0.25 r .

governed mainly by gravity excitation. Comparison with Fig. 23.6 further shows
that the resonances at no/3 and no/2 are also governed in the main by gravity
excitation.
A non-circular, horizontally supported rotor has, as a result of gravity ex-
citation, a 2n-vibration with a resonance at n ~ no/2 (see Chap. 13). If a crack
occurs, then the vibrations change, as Fig. 23.8 shows for our example. This
change depends upon the position of the crack in relation to the principal axes
of the cross-section. The peak value increases if the degree of non-circularity in-
creases and vice versa. In the example these changes amount to between +41 %
to -36 % of the value without a crack.
The unbalance vibrations of a non-circular shaft with crack are very com-
plex, because they are influenced by the position of the crack as well as by
the position of the unbalance. We shall not proceed further with this problem,
since in diagnosis, unbalance vibrations are less important than gravity-excited
vibrations. Reference [153) should be considered for further reading.
The results of Figs. 23.6 to 23.8 are valid for constant crack depth a = 0.25 r.
For diagnosis purposes it is further important as to how the crack-depth changes
such figures. These changes depend, amongst others, on rotor speed. We shall
confine ourselves, in the following three figures to constant speed n = 1. 7 no
and show the change in vibration due to a growing crack during continuous
operation.
Fig. 23.9 shows the vertical amplitudes of the 2n- and 3n-vibration due to
gravity excitation. Both amplitudes increase with crack depth. To get an idea
of the order of magnitude we consider a leffcott rotor for which no = 600 rpm.
From equation (21.13) its static deflection z. is equal to 2.48 mm. If it has a
338 23 Cracked Rotors

x(2n)
0.4 Xs /.craCk
-$- No crack

O.2xs
f Crack

O~=-~~~~~
o no no n -
2
Fig. 23.8. 2n-vibration of a non-circular shaft (nIl = 1.05 nI) with crack. Vertical
amplitude.
1= 10 T, D = 0.05, a = 0.25 T.

t 0.005xs
x(2n) x(2n)
x(3n)

x(3n)
O~~=F~~-.--~~~

o 0.5
..Q.-
r
Fig. 23.9. Influence of crack depth on the vertical2n- and 3n-vibration due to gravity
excitation for n = 1.7 no, 1= 10 r, D = 0.05 .

crack of depth a = 0.5 r, then the amplitude of the 2n-vibration is given by

x(2n) = 0.00375 x. = 9.3 I'm


and of the 3n-vibration

x(3n) = 0.00075 x. = 1.9 I'm .


In Fig. 23.10 are shown the amplitudes of the n-vibration as a result of
unbalance and gravity excitation. For these it was assumed as before that
e = 0.035 x . For the rotor without a crack we obtain from equations (2.41,
2.42) x(1.7 no) = 1.52 e. The amplitudes increase or decrease with crack depth
23.1 Vibrations of a Rotor with a Transverse Crack 339

x(n)
2e
x(n)
2 2
22.
~1~1
2
1

.1~1
2e
e
2
-$-1
o 0.5 o 0.5
.9.- Q.-
r r
Horizontal Vertical
Fig. 23.10. Influence of crack depth on the n-vibration due to unbalance and gravity
excitation for n = 1.7 no, 1= 10 r, D = 0.05, e = 0.035 z .

and position of unbalance. This effect is especially strong in the vertical direc-
tion. For a = 0.54 r in one case this vibration disappears.
A similar behaviour is observed in the amplitude of the 2n-vibration of
a non-circular shaft with crack. These amplitudes become smaller or larger
with crack position (Fig. 23.11). A crack making the non-circularity smaller
compensates it completely when its depth a is equal to 0.4 r.
Figures 23.6 to 23.11 show the character of the vibrations. Just as impor-
tant is the behaviour of the phase angle of the harmonics. The corresponding
investigations in reference [153] can be summarized as follows:

n-vibration: The phase angle changes in general with the crack depth.

1 0.010 xs
x (2n)
.1 2

O.OOSxs

o 0.5
Q.-
r
Fig. 23.11. Influence of crack depth on the vertical 2n-vibration of a non-circular
shaft nIl = 1.05 nI for n = 1.7 no, 1= 10 r, D = 0.05.
340 23 Cracked Rotors

2n-vibration: The phase angle changes in general with the crack depth if the
shaft is non-circular.

3n-vibration: The phase angle, theoretically, does not change.

Generator rotor. The usefulness of the results for a Jeffcott rotor was proved
for a realistic model in reference [153]. This is reproduced in the following.
The rotor model is shown in Fig. 23.12. It has 12 nodes and 44 degrees
of freedom. It is supported on oil-film bearings and pedestals as shown in
Fig. 16.12. The coefficients for the oil-film bearings were calculated from short
bearing theory and the data for the pedestals are
Mass m = 20 t
Stiffness ~or = 5.00 X 109 Nlm, kvert = 6.67 X 109 Nlm
Damping dhor = 2.83 X 105 Nslm, dvert = 3.27 X 105 Nslm.

1 2.11

m=71t 1.1m /

I. 12.6m ..I
Fig. 23.12. Model of a generator rotor. P = 735 MW, n = 3000 rpm.

45004----+~L-~--~-

f=:!==:!t=7===t====~ - ~~----=il
1500 -+-+-I-~----t------t----

~+----+-_ ~ --- rk=::-1~


O~~~~~~~~~
o 1500 3000 4500 rpm
n-
Fig. 23.13. Natural frequencies and mode shapes of the generator rotor.
23.1 Vibrations of a Rotor with a Transverse Crack 341

In Fig. 23.13 are shown the calculated natural frequencies against speed.
The mode shapes are equal in pairs and change only a little with speed as do
the natural frequencies. For further calculations a crack of depth 95 mm was
assumed in cross-section 7, corresponding to a = 0.25 T. Fig. 23.14 shows the
calculated dynamic displacements at the right hand pedestal, node 12, as a
result of gravity excitation. In spite of the rather deep crack the resonances
are not very sharp. At their largest the peak values are 4.5 fm horizontal and
8.2 fm vertical for a speed of 2000 rpm corresponding to n4. The harmonics show
resonances at nt, n2 and n4, at a half and a third of these values respectively, as
well as horizontal at ns/3and vertical at ng/3. It can be seen that resonances
with higher natural frequencies occur, but not all natural frequencies in the
speed range considered contribute to the overall vibration.

5 f 11m
11m
x([) x([) 5

2000 rpm rpm


n-

,L ':l----""Jk--'---=?"' ~&=--r-----,-
1
f ~4
X(2n)Il:~
o
3
I
n,
2000 2000
I
n4

I
o I I 2000 I 2000

t 1l~1 ~ ~ ~
x(3n) ~
oA.~
oI I 2000
n, ~
I
ns
"3 3
"3
Horizontal Vertical

Fig. 23.14. Generator rotor with crack of depth a = 0.25 T at section 7.

Displacements at the right hand pedestal (node 12) as a result of gravity exci-
tation.
342 23 Cracked Rotors

Calculations with unbalance and non-circular shafts gave essentially similar


results as for the Jeffcott rotor. There are however too many phenomena to dis-
cuss them here. In practice it is best to carry out such calculations appropriate
to a given rotor, when a crack is suspected.

23.2 Case Studies

At least as important as theoretical knowledge is practical experience, and again


there are many publications in the literature. Some of these are to be found
in references [154, ... 169], in which other work is also cited. The designer,
researcher and the user thus have at their disposal a wealth of experience. We
cannot analyse these fully here and so some noteworthy characteristic features
will be described to give a first appreciation.
Cracks occur in the shafts of steam turbines, generators and pumps. Most
frequently they appear in the LP- and generator rotors of large turbine sets.
This is mainly because such rotors undergo relatively large alternating stresses
due to their weight.
Cracks are mostly discovered through the observation of a change of vi-
bration characteristic. From a first suspicion to ultimate confirmation that a
significant crack had occurred a variable length of time is often stated. This
could be between a few days and several years but with today's knowledge and
improved vibration surveys we should be able to identify a crack in time to
prevent a disaster. Indeed, in the past in most cases, cracks were discovered in
time for the machine to be shut down without a great deal of damage being
done.
With regard to the initiation of a crack, there are two possibilities. On the
one hand a small initial crack can already be present before operation, and can
grow during service; on the other hand a crack can arise only after a certain
period of operation. In this respect most reports of cracks give insufficient in-
formation. We can, however, assume that in most cases an initial crack was
present, because the cracks could not be explained using conventional stress
calculations. From our knowledge of fracture mechanics it is found to be suffi-
cient for the growth of a crack if it has an initial depth of only a few tenths of a
millimeter, which could not have been discovered without special measurements
anyway.
The operating time up to the point when a crack is suspected, has also been
very variable. Sometimes it was a few hundred, but mostly a few ten thousand
of hours. Indeed, it sometimes amounted to several years and once thirty years.
This shows very clearly how complex the problem is and how difficult it can be
to give a prognosis.
Cracks occur at the following locations amongst others:
Step changes in shaft diameter (Fig. 23.15)
23.2 Case Studies 343

Fig. 23.15. Shaft with step change in diameter.

- The ends of shrink fits (Fig. 23.16)

Fig. 23.16. Shrink fit.

- Grooves for keyways (Fig. 23.17)

-~---3---
, Fig. 23.17. Shaft with groove for keyway.

- Cross cuts (Fig. 23.18)

~~ I
- - -+--- --- - -- -- Fig. 23.18. Generator shaft with cross cuts.
I

- Sharp edges (Fig. 23.19)

/
/\
// I
l' J
\ /
I /
/ /
/ /
Fig. 23.19. Generator shaft with winding grooves.
----
,1-/
344 23 Cracked Rotors

- Grooves for conical keyways (Fig. 23.20)

~
-+- Fig. 23.20. Shaft with conical keyway (diagrammatic).
I

- Radial borings, especially with thread (Fig. 23.21).

: ~ F;g.".21. Shaft wlth,adW bolt (,""",~t").


The broken surface always has the appearance of a fatigue fracture and finally
amounts to between 20 % and 90 % of the cross-section.
The root cause of cracks has always been stress variation, which falls into
two categories:

- Large stress amplitudes and a small number of load reversals (such as occurs
at startup and run down)

- Low stress amplitudes and a large number of load reversals (such as occurs
in normal operation).
23.3 Summary 345

Damage has always been the result of a combination of the two, sometimes with
the first and sometimes with the second predominating. The extent of the stress
variation is an important quantity. Its nominal value at the crack position has
amounted to between 17 and 35 N/mm 2
As well as the above causes the following conditions and prerequisites have
been identified:

- Initial crack (due to manufacture, residual stress)

- Stress concentration (due to geometry)

- Fretting

- Reduced resistance to stress variation due to moisture or corrosive media.

Some consequences of such damage are given in the next section.

23.3 Summary

The discussions in Sects. 23.1 and 23.2 show that the problems associated with
a cracked rotor are very complex. To obtain a basic understanding some main
points on diagnosis and on the avoidance of cracks are summarized in the fol-
lowing. However, it should be remembered that these points represent simplifi-
cations and that other possibilities can occur.
- The observation of vibration is an appropriate means of discovering a crack.
When a transverse crack is present the vibration has 2n- and 3n-components
as well as the fundamental of frequency n. The 3n-component is weaker
than the 2n-component and frequently does not occur.
Changes of vibration are particularly striking for larger crack depths (from
about a> 0.15 r).

A one-sided crack is better observed than a partial or completely circular


crack. However, the latter kind also ultimately changes the vibration sig-
nature.

Vibrations usually increase with increasing crack depth. They can, however,
also decrease, as Figs. 23.10 and 23.11 show.

For diagnoses purposes the phase angle is as important as the vibration dis-
placement. In general, the phase angle changes with increase in crack
346 23 Cracked Rotors

depth. It can, however, also change if the crack causes a thermal curva-
ture of the shaft, and so in general the angle of the unbalance can change.
- If a crack is suspected it is recommended that the machine be shut down
and at the same time the vibration be exhaustively measured and anal-
ysed. From experience, there should be sufficient time to prepare special
measurements and to await a suitable point in time to carry them out.

- When a crack is present, resonances at nk/3 and nk/2 occur in part alongside
the usual resonances at n = nk, as, for example, is shown in Fig. 23.14.

- A crack reduces the stiffness ofthe shaft. Hence the resonance speeds are lower
and the amplitudes of vibration greater. Fig. 23.22 shows an example. In
this case the vibration increased over a period of thirty days by about
36 %, whilst the measured resonance speed remained roughly constant.
After a further nine days the vibrations increased markedly and a shift in
resonance of about -4.4 % could be detected.

1 400
xl!:)

200

04-----.-------~------._------~------_.--
800 900 rpm 1000
n-

Fig. 23.22. Measured rotor vibration during run down in the region of n2/2 for a
735 MW turbo-set with a crack in the generator rotor. From reference [154], p. 44.

The avoidance of cracks.


- In areas of changing load and from a nominal stress of about 30 N/mm 2 one
should avoid at the design stage anything that can lead to a crack. Above
all, things to be avoided are sudden changes of cross-section, keyways,
radial borings and the possibility of fretting.
23.3 Summary 347

Changes in cross-section should be shaped so that the increase in stress is as


little as possible.

Beware of residual stresses.

With dampness or corosive media, beware of the sharp reduction in endurance


limit.

Areas of high stress after fabrication should be inspected carefully for micro-
cracks.
24 Solution of the Equation of Motion

The vibration problems dealt with in this book are represented by the equation
of motion,
Mi+Bx+Kx=f(t). (24.1 )
Its complete solution consists of the solution of the homogeneous equation to-
gether with the particular solution corresponding to the right-hand side. The so-
lution of the homogeneous equation leads to the eigenvalue problem, which gives
the eigenvalues and eigenvectors, and hence the natural frequencies (Sect. 24.1).
Particular solutions are described in the following sections. Sect. 24.4 concerns
the modal method and is especially important not only for calculation purposes
but also for the understanding offorced vibration. The chapter ends with a short
section on the Rayleigh-quotient and a reference to the approximate calculation
of natural frequencies for small changes in a system.

24.1 Homogeneous Equation, Natural Frequencies

With f(t) = 0, equation (24.1) reduces to the homogeneous equation,

Mi+Bx+Kx=O (24.2)

whose solution describes the natural vibrations of the model in question.


We begin with the special case B = 0, that is with negligible gyroscopic
effect and damping. Thus equation (24.2) then further reduces to

Mi+Kx=O. (24.3)

With the substitution


x = cp sinwt (24.4)
this leads to the equation

(24.5)

Solutions cp f:- 0 exist only when


350 24 Solution of the Equation of Motion

With n as the order of the matrix, this leads to a polynomial of nth degree in
w 2 If we put /. = w 2 , then we obtain the equation,

(24.6)

The following definitions are useful:


(K - w 2 M) the characteristic matrix
det (K - w 2 M) the characteristic determinant
Equation (24.6) , the characteristic equation or secular equation.
/. = w 2 , the eigenvalue
cp , the eigenvector.
The roots /'1, /'2, . /.n of equation (24.6) are real and positive or zero provided
that
M and K are symmetric
M is positive definite
K is positive definite or semi-definite.
With these assumptions we obtain,

that is pairs of eigenvalues, which from substitution (24.4) have the interpreta-
tion of frequencies and are called natural frequencies.
For completeness, substitution (24.4) should be extended by a cosine term.
Thus the kth eigensolution is

(24.7)

where A k , Bk are constants and

Ak
/k = arctan Bk . (24.8)

If all the eigenvalues /.k are different and none is zero, then the general solution
of equation (24.3) is
n

Xh(t) = L Xk(t) . (24.9)


k=1

The constants Ak and Bk are obtained from the initial conditions of the com-
plete solution. With xp( t) as the particular solution corresponding to f( t), the
complete solution
(24.10)
With the initial conditions x(t = 0) and :ic(t = 0) a system of 2n equations is
obtained from equation (24.10), with which the 2n constants are found.
In the general case, when B i- 0, the substitution

(24.11)
24.1 Homogeneous Equation, Natural Frequencies 351

leads to the equation


(A2M + AB + K)!p = O. (24.12)
The condition that !P 1= 0 leads to the requirement that

and hence to the characteristic equation

(24.13)

where m = 2n and n is the order of the matrices. Equation (24.13) has m roots,
the eigenvalues, which as a rule are complex conjugate. With

(24.14)

and using equation (24.12),

(A%M + AkB + K)!pk = 0


(24.15)
(Ai2M + AiB + K)!pi = 0
giving the solutions
(24.16)
These are complex eigenvectors, whose components Tlk, T2k, ... Tnk and Slk,
S2k, ... Snk are determined up to an arbitrary factor. With constants Pk , P: it
follows from substitution (24.11) that the kth solution is

(24.17)

This consists of displacements and rotations, which must, of course, be real and
results from the fact that, in equation (24.17), Pk , Pi are complex conjugate.
Thus if we put
(24.18)
where A k , Bk are real, then after transformation using the above formulae and
from the Euler equation, we obtain the solution

Xk(t) /~.kt [rk (Ak cos wkt + Bk sinwkt) + Sk (Bk coswkt - Ak sinwkt)]

eak t Ck [rk sin (Wkt + 'Yk) + Sk cos (Wkt + 'Yk )] (24.19)

where Ck, 'Yk are given in expressions (24.8).


Accordingly, for an arbitrary coordinate i,

(24.20)
352 24 Solution of the Equation of Motion

or
(24.21)
where
8ik
eik = arctan - . (24.22)
Tik

From equation (24.21), a general displacement or rotation for ak < 0 consists


of a decaying vibration and for ak > 0 of an increasing vibration of frequency
Wk
A measurement of the damping of the kth natural frequency is the quantity

(24.23)

This is called the modal damping ratio and is analogous to the damping ratio
(equation (2.18)) ofthe simple vibrator.
The relationship shown in (24.23) is only of significance for values of lakl
up to about Wk. Dk changes very little for higher values, approaching either +1
or -1 asymptotically. Expression (24.23) is also not of any great use for real
eigenvalues, as Dk is then always either +1 or -l.
The values of displacements Zlk, Z2k, Znk are determined by the factor
CkUik and in general different phase angles eik exist between them. For the
general solution and for determining the constants, these can be regarded as
equivalent to the case when B = O.
Thus we have the equations of the eigensolutions. These assume the de-
termination of eigenvalues - the solution of the eigenvalue problem. For this
purpose there exist many numerical methods, whose foundation is the so-called
theory of eigenvalues.
We shall differentiate between the special and the general eigenvalue prob-
lem. The special eigenvalue problem is posed when B = 0 and the matrices
M and K are symmetric and positive definite. If we multiply equation (24.5)
by M- 1 and put A = M- 1 K, then we have the following equation, in which
w2 = p
(A - pI) If' = 0 . (24.24)
The solution consists of n real eigenvalues and eigenvectors.
The generalized eigenvalue problem is posed by equation (24.12), in which
the matrices M, Band K may now be quite general.
Some solution procedures assume a standard form corresponding to equa-
tion (24.24). This gives, from the equation of state (Subsect. 24.4.5),

(C - .H)r = 0 (24.25)

with C from equation (24.61). The problem is of order 2n and 2n eigenvalues


exist, which are mostly complex conjugate. The eigenvalues can, however, also
24.2 Harmonic Excitation 353

be real or completely imaginary. In addition there exist 2n corresponding eigen-


vectors. A description of solution procedures for eigenvalue problems will not
be dealt with here. More detailed considerations are given amongst others in
references [170, 171, 172].

24.2 Harmonic Excitation

For harmonic excitation we use complex notation. The equation

Mi + Bi + K~ = fe iwt (24.26)

has the particular solution


(24.27)

When substituted, this gives the equation

(24.28)

whose solution gives the vector ~ of the displacements.


The expression in brackets in equation (24.28) is the matrix of dynamic
stiffnesses
K(w) = K - w2 M + jwB . (24.29)

Its reciprocal is the transfer function and its definition and determination are
dealt with in Subsect. 24.4.4.

24.3 General Excitation

For general excitation h(t), ... fn(t) we can calculate the solution using one of
the many well-known direct integration procedures (see, for example reference
[171]). In program MADYN (Chap. 25) the modal solution is determined and
the polygon procedure used, as described in Sect. 2.4.

24.4 Modal Method of Calculation

The system of equations of motion is in general coupled. The expenditure in


computation time to obtain a solution can, because of this, become a problem
when a large number of degrees of freedom are to be considered. The modal
354 24 Solution of the Equation of Motion

method allows uncoupling and hence a reduction in computer time. For this
purpose, however, we still need to know the eigensolutions, but less computer
time will be required in total, if the same particular solution is needed very often,
as for example in the calculation of forced amplitudes of vibration. The modal
method is advantageous for solutions where great accuracy is not demanded
when we can make do with only a few eigensolutions or maybe even one solution.
In any case, the study of modal analysis is to be recommended, because it
provides a good understanding of the vibration of large structures.

24.4.1 Symmetric Matrices

The matrices of the equation of motion are unsymmetric if one takes account
of gyroscopic action, oil-film bearings, steam whirl, seals and similar influences.
In many investigations one can neglect these or make the matrices symmetric
by taking the average values of the coefficients in question.
In the following we assume that M and K are symmetric. Furthermore, we
make B = D and D proportional to M and/or K. Using constants a and /3,

D = aM+/3K (24.30)

With these assumptions the equation of motion (24.1) can be uncoupled. We


replace the displacement coordinates :1:17 :l: n , using the linear transformation

x=~y, (24.31)

by the coordinates Yl, .. Yn. Hence,

(24.32)

This is the matrix of eigenvectors tpk, which are found by solution of the eigen-
value problem Mi: + Kx = o. ~ is called the modal matrix.
Substitution of equation (24.31), multiplying by ~T and putting

M* = ~TM~, D* = ~TD~, K* = ~TK~, *(t) = ~T(t) (24.33)

gives the equation


M*y + D*y + K*y = *(t) . (24.34)
The matrices of this equation are diagonal (for a proof see reference [32),
amongst others). Hence equation (24.34) represents a system of decoupled equa-
tions
m'kih + d'kiJk + k'kYk = f;(t) where k = 1,2, ... n. (24.35)
The quantities

(24.36)
24.4 Modal Method of Calculation 355

are described as the kth generalized mass, damping, stiffness and force, respec-
tively.
Equations (24.35) each correspond to an equation of motion of a simple
oscillator. The coordinates Yk and the motions Yk(t) can, however, only have
kinematic meaning in very simple cases. One calls YI .. Yn the principal coor-
dinates, because the matrices are made diagonal by the transformation (24.31).
Equations (24.35) can then be solved in well-known fashion (see Chap. 2). Using
equation (24.31) one finally obtains the displacements XI (t), ... xm(t), ... Xn( t),
where n

xm(t) = L 'PmkYk(t) = L Ymk(t) . (24.37)


k=1

Yk(t) is described as the principal vibration. By contrast, Ymk(t) is called the


modal principal vibration. From equation (24.37), xm(t) is thus equal to the sum
of n modal principal vibrations Ymk(t). In other words, xm(t) is developed from
the principal coordinates Yk( t), using as coefficients the components 'Pmll ... 'Pmn
of the eigenvectors.
When the system has uniformly distributed damping then proportional
damping from equation (24.30) can be assumed as an approximation. This
assumption is also useful in the calculation of response curves, for example,
to obtain finite resonance peaks. Here one proceeds from assumed magnifica-
tion factors or damping ratios, as described in the following. Dividing equation
(24.35) by m;" gives the equation

(24.38)

where
(24.39)

and
(24.40)

The last is the so-called kth modal damping ratio.


The natural frequencies Wk are known already from the solution of the eigen-
value problem and so there is no need to calculate them from equations (24.39).
By putting d;" = am;" + (:Jk'k, one obtains the expression

(24.41)

This is shown in Fig. 24.1.


If one puts Dk = Da when Wk = Wa and Dk = Db when Wk = Wb, then from
expression (24.41),
WbDa - WaDb , {:J -_ 2WbDb - waDa
a = 2WaWb 2 2 (24.42)
Wb - Wa W~ - W~

By this means all the required Dk values are found.


356 24 Solution of the Equation of Motion

Wk-
Fig. 24.1. Modal damping ratios for proportional damping.

One is not restricted, however, in the choice of Dk to the use of expression


(24.41). Instead it can be done quite freely. Using expression (24.40), one can
obtain the elements die and hence the diagonal matrix D* from assumed values
of D k Then by back transformation one can if needed find the accompanying
damping matrix.

24.4.2 Harmonic Excitation

For harmonic excitation with force


f()
_ t = f -fn ) T eiwt
(f-1' LJc' 0 -
- _e
f iwt (24.43)

then the kth generalized force is

E.(t) = rpI f. eiwt = 1~ eiwt (24.44)

and the particular solution from equation (24.35) is

(24.45)

Also from equation (24.37) the mth displacement is


on
t
1l.m () = '"'
L..J 'Pmk1!.k eiwt = 'L..J1l.
" mk eiwt (24.46)
k=1
24.4 Modal Method of Calculation 357

with

Jamk
-'l<mk 'PmkJl..k = Xmk e
~ j(wt + amk) j(wt + am)
L...J Xmk e = Xm e (24.47)
k=l

The amplitude Xm and the phase angle am depend on the frequency w.


The form of Xm for a simple example is evident from Fig. 24.2. This figure
shows the amplitudes and phases of the three modal vibrations and of the
displacement X2 for a system with three degrees of freedom and for the special
case of equally phased excitations and equal modal damping ratios. This will
be investigated more closely in the next sub-section.

X2
X2k

w, W2
w--

Fig. 24.2. System with three degrees offreedom. Forms of:l: 2 and 0.2 from the modal
components. 0.21(0) = 0.22(0) = 0.23(0) = 0, Dl = D2 = D3 = 0.07.

In expression (24.43) the amplitudes of the excitation forces are assumed


to be constant. For unbalance excitation at angular velocity [} of the rotor the
excitation force is

(24.48)
358 24 Solution of the Equation of Motion

where IL ... U z are unbalances and the generalized force is

(24.49)

Putting w = {}, the solution can be obtained as for the case of excitation forces
of constant amplitude.

24.4.3 Reduced Modal Calculation

The sum of the n modal vibrations constitutes the exact solution of the equa-
tions of motion with the given assumptions. Approximate solutions are obtained
if one neglects terms which contribute little to the result. From equation (24.31)
it is obvious that one does not consider all of the eigenvectors. One speaks of
reduced modal calculation.
For general excitation rules can hardly be given as to which eigenvectors
should be considered and which not. For harmonic excitation it is, however,
possible, as will be shown in the following.
For harmonic excitation, at a resonance the appropriate modal vibration
dominates. This is evident in the example of Fig. 24.2 at the second and third
resonances. The peak values of :1:2 and :1:22 differ only by 3.5 % and 2.7 %
respectively from the values of :1:2 and :1:23. At the first resonance :1:22 and :1:23
contribute strongly to the peak value, as well as does :1:21.
The relationship of the peak value to the dominating component depends,
amongst other things, on the null phase angles (that is, when time equals to
zero) of the modal components. Fig. 24.3 shows this for the second resonance of

t 1.2

(X2)mQx 1.1
(X22)mQx

1.0

0.9

0.8
r---r-~?~--~--~
-It o It
1X21(0) - - -
Fig. 24.3. Influence of null phase angle on the ratio of peak values at the second
resonance, example of Fig. 24.3. Q~~) = o.
24.4 Modal Method of Calculation 359

the example of Fig. 24.2. The relationship between the ratio of peak values and
the values of the two other null phase angles is indicated here for 022(0) = O.
This ratio lies in the region between 0.86 and 1.14.
Consider this case generally. For a system with n degrees of freedom, we
shall attempt to calculate roughly the amplitude Zm at the 1'th resonance, for
W =Wr
For equally phased excitation and for zero null phase angles, the complex
amplitudes of the modal vibrations have the positions shown on the left Argand
diagram in Fig. 24.4. The amplitudes for k = 1 to r - 1 lie on the left and for
k = l' + 1 to n lie on the right of '!!<'mr.
With the summations
r-1 n

1!.ma=L .!!<.mk' 1!.mb = L .!!<.mk and 1!.me = 1!.ma +1!.mb (24.50)


1 r+1

then
.!!<.m = 1!.mr +1!.me (24.51)
For relatively small values of1!.me, it can be seen that 1!.mr is a good approximation
to 1!.m. 1!.me is small when the neighbouring natural frequencies W r -1 and W r +1
are well spaced from Wr and the damping is weak. With these assumptions the
resultants 1!.ma and 1!.mb have directions roughly parallel to the real axis (the
angles'Y and e are small) and they change more or less in a mutually reciprocal
fashion.
It can be seen that the reduced modal calculation requires caution. If one
does not know which terms are needed, then it is recommended that one calcu-
lates twice, once reduced and once with no reduction. In this way the necessary
experience can be acquired to consider further cases with some confidence. Ex-
amples can be found for example in references [173] and [174].

1m 1m

Re Re
~ml

~mr-l ~mr+l

~mr

I
Fig. 24.4. The formation of ~m.
360 24 Solution of the Equation of Motion

24.4.4 Transfer Function

The ratio of the response at coordinate i to an excitation at coordinate k is


called a transfer function.
For harmonic excitation with
T . .
f(t)= ( 0, ... 1k 0) eJwt=feJwt (24.52)

the response at coordinate i, from expression (24.46), is


n

.!l<.i(t) = L 'Pir'!!.r e jwt , (24.53)


r=1
in which the index k has been replaced by 1'.
From expression (24.45),

y = 1; (24.54)
-r k; - m;w 2 + jd;w .

Also, using expressions (24.44) and (24.52)

1; = "": f = 'Pkr 1k ' (24.55)


and thus
=L
n
.!l<.i(t) 'Pir 'Pkr f jwt (24.56)
r=l
k*r - m*w2
r
+ J'd*w
r
-k e .
The transfer function is thus

(24.57)

and so, putting


Wr
2 k;
== - , Dr --~ (24.58)
m*r 2m;wr
then
~ 'Pir 'Pkr 1 _ h. ( ) jaik(W)
Hik () - (24.59)
w = L.J
r=1 m; 2
wr - W
2
+ J'2DrWrW ,k w e .

In many cases the 1'th term of the sum dominates at W = Wr and so

h ik (W == Wr ) 'Pir 'Pkr 1
~ * 2 (24.60)
mr wr 2 Dr

24.4.5 Unsymmetric Matrices

If one has a system of unsymmetric matrices M, Band K, the individual equa-


tions of motion can be uncoupled by the application of so-called left eigenvec-
24.4 Modal Method of Calculation 361

tors, as will be shown in the following. Further details can be found in reference
[32], amongst others.
Multiplication of the left hand side of equation of motion (24.1) by M- I
leads, with the identity x = x, to the equations

x 0 + IX + 0
i: = - M-IKx - M-IBx + M-If(t).
With

Z~{:} , c- - (24.61 )

one arrives at the so-called state equation


z= Cz+ b(t) (24.62)
in the state vector z = z(t).
If n is the dimension of the equation of motion the equation of state has
dimension 2n. The eigenvalue problem
z- Cz = 0 (24.63)
gives, with the substitution
z = cP eAt

the eigenvalues Ak and the eigenvectors CPk, where k = 1 ... 2n. The eigenvalues
are the same as those obtained from equation (24.1).
Transposing C gives the second eigenvalue problem
z- C T Z = o. (24.64)
With the substitution
z = tP eAt

the eigenvalues Ak and the eigenvectors tPk are obtained. The eigenvalues Ak are
equal to those from the corresponding eigenvalue problem of equation (24.63).
The eigenvectors tPk are, however, not equal to CPk. CPk and tPk are designated
as right and left eigenvectors.
The linear transformation
z = tPw, (24.65)
with tP = (CPk) as the matrix of right eigenvectors normalized to unity gives,
from equation (24.62) the system of the following 2n decoupled equations.
w= Aw + tPT b(t) . (24.66)
Here A = diag (AI ... A2n) and tP = (tP k) is the matrix of left eigenvectors
normalized to unity. The kth equation of (24.66) is
(24.67)
362 24 Solution of the Equation of Motion

where
(24.68)
Solving equation (24.67) for k = 1, .. . 2n and substitution into equation (24.65)
gives the solutions
2n
Zm(t) = L CPmk Wk(t) (24.69)
k=l
in which the required solutions xm(t) are to be found. This method is described
as bimodal, to distinguish it from the usual modal method.

24.5 Rayleigh Quotient

The influence of small changes to the system on its natural frequencies can be
estimated by using the Rayleigh quotient.
We shall neglect the gyroscopic effect, damping and similar effects and re-
strict ourselves to the equation

Mx+Kx=O.

With the assumptions made in Sect. 24.1, we can find n natural frequencies
Wk and eigenvectors 'Pk. Inserting these values into equation (24.5) gives the
equation
K'Pk = w% M'Pk
and, after multiplication by 'PI the Rayleigh quotient

(24.70)

or with expressions (24.39)


2 kZ
W
k -
--
m* (24.71)
k

Changes in stiffnesses and in masses yield LlkZ, Llmk and give the new natural
frequency
12 kZ + LlkZ
(24.72)
W
k
-
- m*k + Llm*k
together with new eigenvectors.
If the parameter changes are small, then we can use the old eigenvectors to
calculate these eigenvalue and eigenvector changes. With LlK and LlM as the
matrix changes, the additional terms become

(24.73)
24.5 Rayleigh Quotient 363

and using expressions (24.72) and (24.73) the new natural frequencies are

(24.74)
1 + l1mk/m'k

In this way we can avoid the need to carry out a new calculation to find the
revised eigenvalues.
25 Program MADYN

H. D. Klement

In the foregoing chapters, reference was made to the program MADYN in the
solution of several examples of rotor vibration. Here we shall describe in a
concise way the capabilities of this program.

25.1 Review

MADYN, in its function, stands between special rotordynamics programs, which


in most cases can only calculate for a simple rotor in oil-film bearings and gen-
eral FE-programs, like for example NASTRAN, ASKA, ANSYS, STARDYNE,
ABAQUS etc., which on the other hand (with the exception of ANSYS), do not
meet the special needs of rotordynamics. These needs include gyroscopic action
and the inclusion of unsymmetrical coefficients in the stiffness and damping
matrices, e. g. for the oil-film bearings.
As is shown in Fig. 25.1, the complete program system consists of a group of
independent programs, which exchange between each other essential information
relating to various items of data.
The model preparation is carried out in the main part of MADYN with
coordinate reduction (after Guyan) and eigenvalue calculation, as well as struc-
ture and eigenform plots. Further eigenvalue algorithms are available in FKRIT
and PARVAR for the determination of critical speeds of conservative systems.
In PARVAR a chosen parameter can be varied, such as rotor speed, damping
or the stiffnesses and damping coefficients of oil-film bearings and springs. The
natural frequencies and dampings are presented as functions of these parame-
ters. In RAYLEI, with the help of the Rayleigh quotient, analysis is carried out
to assess how the natural frequencies of a system are affected by variations in
the stiffnesses and masses of individual elements.
Different kinds of loading are handled in further program blocks.
STATIK serves to calculate static loads, weights and load due to tem-
perature differences. It then calculates the displacements, element forces and
stresses. In addition the load conditions can be subsequently superimposed.
Different solution algorithms are available to calculate steady-state har-
monic vibrations. A complex Gaussian decomposition of the matrices is pro-
vided in DHARM. This makes it possible to vary the fluid-film bearing data as
functions of rotor speed. Furthermore, substructures, whose kinetic stiffnesses
366 25 Program MADYN

Model preparation
MADYN Real, complex eigen-
values

direct- bimodal-
FKRIT Critical Speeds
modal
PARVAR Parameter variation
RAYLEI Sensitivity analysis

MHARMC MHARM
Harmonic vibration
HKMAT

TRANSC TRANS Linear Transient


}
Non-linear vibration
NOLINC NOLINR

ASPEK
Earthquakes
BEBEN

RANDOM Random vibration

IADPLOTI I AGPLOT I I AVPLOT I General graphics

Fig. 25.1. Structure of program system MADYN.

were calculated with the program HKMAT can be subsequently added to the
main structure. In typical cases the main structure of the rotor is associated
with its fluid-film bearings and the foundation substructure. MHARM provides
a modal solution with the real eigenvectors obtained from the main portion of
MADYN. This process is decidedly faster, but does not allow the gyroscopic
effect to be considered nor frequency-dependent fluid-film bearing data and
substructures. The bimodal variant in MHARMC provides complex eigenvec-
tors, so that the gyroscopic effect is taken into account in the calculation at least
for a constant rotor speed. Furthermore, the coefficients of the fluid-film bear-
ing matrices can be unsymmetric and local damping can occur in the model.
The loads are, in each case, harmonic forces, unbalances and base excitations.
Transient vibrations are calculated in a modal fashion with TRANS or in
bimodal fashion with TRANSC, with the help of the polygon procedure. Excita-
25.2 Building the Model 367

tion is from forces or base accelerations, and both are functions of time. General
initial conditions can be provided as displacements and velocities, which, in the
given circumstances, are the end results of a previous calculation.
The programs NOLINR and NOLINC allow in addition the application of
local non-linearities, as from fluid-film bearings, springs with special character-
istics or Coulomb friction elements. Additionally, general non-linear forces can
be defined in sub-programs, which have been prepared by the user. The restric-
tion to local non-linearities also allows modal calculations to be performed. In
most cases this is considerably faster than the direct-integration method, which
would otherwise have been necessary.
The programs ASPEK and BEBEN are available for earthquake calcula-
tions. Response spectra, frequently also displayed as shock spectra, are calcu-
lated in ASPEK from acceleration wave forms, which originate from transient
calculations or are provided directly. These are then the input data for the pro-
gram BEBEN, which provides the probable maximum responses in terms of
displacements, velocities, accelerations, element forces and stresses.
Random-vibration calculations are carried out in RANDOM for which
Fourier- or power density spectra are required as excitations. These spectra
can be obtained from wave forms by the use of an integrated analysis portion of
the program. The results are RMS values for the same quantities as in BEBEN.
All programs have their own plotting routines so as to be able to plot the
results in graphical form. Independently ofthis, there is, however, an additional
general plotting program, which allows the user to decide the kind of display
in a very detailed fashion. AD PLOT is intended for diagrams with linear or
logarithmic axes, with or without grid lines, numbering and legends. In addition,
the input data can be manipulated in many different ways. AGPLOT shows
the geometry of the structure or its stiffness and mass distribution. If needed,
parts of the structure can be shown displaced in space (exploded). A section
is contained, with which shafting can be shown in a simply optimal fashion.
AVPLOT finally shows the deformed structure in a similar way to AGPLOT.

25.2 Building the Model

The input of the model takes the usual form found in FE-programs, that is by
the definition of nodes and elements. In order to keep expenditure to a mini-
mum, the possibility of generalization is provided. The positions of the nodes
can be described in a system of cartesian or cylindrical coordinates as well as in
auxiliary systems and in relative coordinates. Each node has initially six degrees
of freedom, which by the provision of boundary conditions, are reduced in num-
ber. Here also, different coordinate systems are available and all elements can
368 25 Program MADYN

be orientated in space in a general way. The general straight beam is described


by its cross-sectional data, such as surface, moment of inertia, shear factor and
mass. A bar is a special form of the general beam and differs only in having a
shortened set of inputs and outputs. The shaft is also only a special case of a
beam which differs by virtue of its rotational speed which leads to its gyroscopic
matrix. At zero speed the shaft is just a rod or tube, which mechanically is no
different from any other structure. A maximum of twenty rotor cross-sections
can constitute a so-called reduced rotor. This type of element is used in the
main, when many stepped sections occur in the line-out, where the normal
representation would lead to an exaggeratedly fine sub-division. The bent tube
element takes account of any internal pressure, which might lead to an increase
in stiffness. The spring/damper element can be orientated globally or locally. It
combines together all the six degrees of freedom of two nodes. For rotordynam-
ics, the fluid-film bearing is indispensable and its matrix can have unsymmetric
coefficients. As well as additional masses and mass moments of inertia, there is
the possibility for the theoretically versed user to include generalized elements,
whose coefficients can be given explicitly, as well as direct matrix addition.
The rigid element which serves to bridge a spatial gap is not an element
usually considered in FE-theory. The kinematic coupling is a general rigid el-
ement, with whose help gears, joints, difference coordinates and other special
features can be defined.
From the element matrices, the global matrices of the structure with variable
width of band can be built up, that is each row or column can be given an
individual length.

25.3 Eigenvalue Algorithm

There are four algorithms at our disposal for calculating the real eigenvalues of
the system defined by the equation

Mi+Kx =0 (25.1 )

The particular algorithm used depends on the size of the model and the desired
eigenvalues. The Jacobi procedure is used for full matrices, and hence is very
calculation intensive and demanding of storage space. For this reason it is only
suited to small models of up to about one hundred degrees of freedom.
The Householder method needs only about one quarter of the memory and
is faster by a factor of between about three and ten. It is thus used for models
of up to about 500 degrees offreedom (in special cases up to 1000). With poorly
conditioned matrices, it is numerically sensitive and this can be a disadvantage.
Both methods can also be employed for reduced systems.
The iterative procedure is better suited for larger models, for which only
the lower eigenvalues are needed. Here we use either the simultaneous vector
25.3 Eigenvalue Algorithm 369

iteration of Corr and Jennings [175] or the subspace iteration of Bathe and
Wilson [171]. Both procedures depend on the band structure of the matrices. In
implementing simultaneous vector iteration the external memory was not used
in the removal of the intermediate vectors, whereas it is retained for 'subspace
iteration. For this reason the choice of possible eigenvectors for simultaneous
vector iteration is trivial. On the other hand this procedure is somewhat faster
and more reliable in regard to ill-conditioned eigenvalues.
For the general eigenvalue problem

Mi + (D + G) x + Kx = 0 (25.2)

with real but unsymmetrical matrices, there are two algorithms available.
The Hessenberg procedure replaces the Jacobi- or Householder algorithm.
Because of the complex eigenvalue calculation, it is, however, slower than a real
solution with Householder, by a factor of around 20. The inverse vector iteration
method is better suited for larger systems. First an approximate solution is
sought with the help of subspace iteration, which is then used to find the exact
solution. Because of possible extended bimodal calculations both the right and
also the left eigenvectors are determined.
In program FKRIT the critical speeds are calculated as eigenvalues of the
system
Mi+Gx+Kx= 0 (25.3)
in which the stiffness matrix K is symmetric. This is done by searching for
the determinant. In this case the eigenvalues are real, but the eigenvectors are
complex. By confining ourselves to a symmetric stiffness matrix the calculation
times are only a little greater than for a real eigenvalue calculation.
In program PARVAR, the Hessenberg algorithm, inverse vector iteration or
determinant search can be selected. In regard to the special boundary conditions
(the approximate solution can be found by changing only one parameter), these
procedures can, however, be so .modified that computation time increases less
proportionally to the number of parameters. Furthermore for the determinant
search, there is the additional special case of isotropic bearings. If one can
make use of this, special time advantages can be achieved. In this special case
PARVAR concurs with many of the old rotordynamics programs, which used
the transfer matrix method, and the determinant search method is now widely
used.
References

[1] Foppl, A.: Das Problem der Lavalschen Turbinenwelle. Der Civilingenieur, 4,
(1895) 335-342
[2] Jeffcott, H. H.: The lateral vibration of loaded shafts in the neighbourhood of
a whirling speed. Phil. Mag., 6:37, (1919) 304-314
[3] Rankine, W. J. M.: On the centrifugal force of rotating shafts. The Engineer,
27, (1869) April 9
[4] Dunkerley, St.: On the whirling and vibration of shafts. Phil. Trans. of the Royal
Soc., A, 185, I, (1895) 279-360
[5] Stodola, A.: Dampf- und Gasturbinen. 4. Aufl., Berlin: Springer (1910)
[6] Myklestad, N. 0.: A new method of calculating natural modes of uncoupled
bending vibrations. Int. Aeron. Sci., 11, (1944) 153-162
[7] Prohl, M. A.: A general method of calculating critical speeds of flexible rotors.
J. Appl. Mech. Trans. ASME Ser. E, 67, (1945) 142-146
[8] Newkirk, B. L.: Journal bearing instability: A review. Inst. Mech. Conf. on
Lubr. and wear, (1957) October
[9] Lund, J. W.: Review of the concept of dynamic coefficients for fluid film journal
bearings. Journ. of Trib. 109, (1987) 37-41
[10] Someya, T.: Journal-bearing data book. Berlin ... Tokyo: Springer (1989)
[11] Lomakin, A. A.: Feed pumps of the SWP-220-280 type with ultra-high oper-
ating data. Energomashinotroenie, 2, (1955)
[12] Workshop at Dniv. College Station, Texas A & M: Rotordynamic instability
problems in high-performance turbomachinery. NASA (1980) CP-2133, (1982)
CP-2250, (1984) CP-2338, (1986) CP-2443, (1988) CP-2409, (1990) CP-3122
[13] Thomas, H. J.: Instabile Eigenschwingungen, angefacht durch die Spaltstro-
mungen in Stopfbiichsen und Beschauflungen. Bull. de l'AIM 71: 11,12, (1965)
333-344
[14] Newkirk, B. L.: Shaft whipping. General Electric Rev. 27, (1924) 169-178
[15] Kimball, A. L.: Internal friction theory of shaft whirling. General Electric Rev.
27, (1924) 224-251
[16] Kimball, A. L.: Measurement of internal friction in a revolving deflected shaft.
General Electric. Rev. 28, (1925) 554-558
[17] Henning, G., Schmidt, B., Wedlich, Th.: Erzwungene Schwingungen beim Re-
sonanzdurchgang. VDI-Ber. 113, (1967) 41-46
[18] Hassenpflug, H. L., Flack, R. D., Gunter, E. J.: Influence of acceleration on the
critical speed of a Jeffcott rotor. Trans. ASME 103, (1981) 108-113
[19] Markert, R., Pfiitzner, H., Gasch, R.: Biegeschwingungsverhalten rotierender
Wellen beim Durchlaufen der kritischen Drehzahlen. Konstruktion 29:9, (1977)
355-365
372 References

[20] Gasch, R., Markert, R., Pfiitzner, H.: Acceleration of unbalanced flexible rotors
through the critical speeds. Journ. of Sound and Vibr. 63:3, (1979) 393-409
[21] Hasselgruber, H.: Zur Berechnung der elastischen und diimpfenden Lagerung
hochtouriger Wellen. MTZ 15, (1954) 373-376
[22] Balda, M.: Dynamic properties of turbo set rotors. In: Dynamics of rotors. Symp.
Lyngby. Berlin: Springer (1975) 27-55
[23] Lund, J. W.: Stability and damped critical speeds of a flexible rotor in fluid-film
bearings. Trans. ASME, J. Eng. Ind. 92:2, (1974) 509-517
[24] Saito, S., Someya, T.: Study of damped critical speeds and damping ratios of
flexible rotors. Trans. ASME, J. Vibr. 106:1, (1984) 62-71
[25] Springer, H.: Optimale Lagerdiimpfung fUr hochflexible Rotoren. ZAMM 65:4,
(1985) 105-107
[26] Kriimer, E., Knoblauch, J.: Optimale Lagerdiimpfung eines einfachen Rotors.
Konstruktion 41, (1989) 103-108
[27] Biezeno, C. B., Grammel, R.: Technische Dynamik, Bd. 2. Berlin, ... : Springer
(1953)
[28] Gasch, R., Pfiitzner, H.: Rotordynamik. Berlin, ... : Springer (1975)
[29] Magnus, K.: Kreisel. Theorie und Anwendung. Berlin, ... : Springer (1971)
[30] Hupfer, H.: Beitrag zum kinetischen Verhalten von anisotrop gelagerten Wellen
unter Beriicksichtigung der Kreiselwirkung der Rotormasse. Techn. Univ. Berlin,
Diss. (1974)
[31] Kellenberger, W.: Erzwungene Biegeschwingungen einer anisotrop gelagerten
Scheibenwelle mit Kreiselwirkung und Drehtriigheit, iiufierer und innerer Diimp-
fungo Forschg. Ing.-Wes. 48:3, (1982) 65-96
[32] Kriimer, E.: Maschinendynamik. Berlin, ... : Springer (1984)
[33] Someya, T.: Stabilitiit einer in zylindrischen Gleitlagern laufenden, unrunden
Welle. Techn. Hochsch. Karlsruhe, Diss. (1962)
[34] Someya, T.: Das dynamisch belastete Radial-Gleitlager beliebigen Querschnitts.
Ing. Archiv 34:1, (1965) 7-16
[35] Lund, J. W.: Evaluation of stiffness and damping coefficients for fluid-film bear-
ings. Shock and Vibr. Digest 11:1, (1979) 5-10
[36] Dubois, G. B., Ocvirk, F. W.: Analytical derivation and experimental evaluation
of short-bearing approximation for full journal bearings. Cornell Univ. Rep.
1157, (1953)
[37] Holmes, R.: The vibration of a rigid shaft on short sleeve bearings. Journ. Mech.
Eng. Science 2:4, (1960) 337-341
[38] Smith, D. M.: Journal bearings in turbomachinery. London: Chapman and Hall
(1969)
[39] Glienicke, J.: Experimentelle Ermittlung der statischen und dynamischen Eigen-
schaften von Gleitlagern fiir schnellaufende Wellen - Einfiufi der Schmierspalt-
geometrie und der Lagerbreite. Fortschr. Ber. VDI 1:22, (1970)
[40] DIN 31657. Plain bearings. Hydrodynamic plain journal bearings under steady-
state conditions. Calculation of multi-lobed and tilting pad journal bearings.
Berlin, Koln: Beuth (Draft 1990)
[41] Glienicke, J.: Feder- und Diimpfungskonstanten von Gleitlagern und deren Ein-
flufi auf das Schwingungsverhalten eines einfachen Rotors. Techn. Univ. Karls-
ruhe, Diss. (1966)
References 373

[42) Merker, H. J.: Uber den nichtlinearenEinflufi von Gleitlagernauf die Schwingun-
gen von Rotoren. Fortschr. Ber. VDI, 11:40, (1981)
[43) Lund, J. W., Nielsen, H. B.: Instability threshold of an unbalanced, rigid rotor
in short journal bearings. In: [176), (1980) pp. 91-95
[44) Hori, Y.: A theory of oil whip. Trans. ASME Ser. E, 26:2, (1959) 189-198
[45) Merker, H. J.: Zusammenhang zwischen Zapfenkraft und Zapfenbewegung beim
Kreislager und Kippsegmentlager. Fortschr. Ber. VDI, 11:33, (1980)
[46) Ott, H. H.: Kippsegment-Radiallager bei stationarer Belastung. Forschg. im
Ing. Wes., 40:5, (1974) 133-144
[47) Hertz, H.: Uber die Beriihrung fester elastischer Korper. Ges. Werke, I, (1895)
155-196
[48) Palmgren, A.: Grundlagen der Wiilzlagertechnik. 3. Aufi., Stuttgart: Franck
(1964)
[49) Meldau, E.: Die Bewegung der Achse von Wiilzlagern bei geringen Drehzahlen.
Werkstatt und Betrieb, 84:7, (1951) 308-313
[50) Tamura, H., Gad, E. H., Sueoka, A.: Computer study of radial vibrations in a
rotor/ball bearing system. In: [177), (1986) pp. 553-560
[51) Lomakin, A. A.: Calculation of critical speeds and securing of the dynamic
stability of hydraulic high-pressure machines with reference of the forces arising
in the gap seals. Energomashinostroenie 4.1, (1958)
[52) Black, H. F.: Effects of hydraulic forces in annular pressure seals on the vibra-
tions of centrifugal pumps. J. M. Eng. Sci. 11:2, (1969) 206-213
[53) Black, H. F., Jensen, D. N.: Dynamic hybrid properties of annular pressure
seals. Proc. 1. Mech. Eng. 184, (1970) 92-100
[54) Black, H. F., Jensen, D. N.: Effects of high pressure ring seals on pump rotor
vibrations. ASME Paper 71-WA/FF-38, (1971)
[55) Black, H. F.: The effect of inlet flow swirl on the dynamic coefficients of high
pressure annular clearance seals. Univ. of Virginia, Charlotteville, (Aug. 1977),
unpubl.
[56) Childs, D. W., Dressman, J. B., Childs, S. B.: Testing of turbulent seals for
rotordynamic coefficients. In: [12), (1980) pp. 121-138
[57) Childs, D. W.: Dynamic analysis of turbulent annular seals based on Hirs' lu-
brication equation. Trans. ASME, 82-Lub-41, (1982)
[58) Childs, D. W.: Finite-length solutions for rotor dynamic coefficients ofturbulent
annular seals. Trans. ASME, Journ. of Lubr. Techn. 82-Lub-42, (1982)
[59) Alford, J. S.: Protecting turbomachinery from self-excited rotor whirl. Trans.
ASME, Journ. of Eng. for Power 87, (1965) 333-344
[60) Spurk, J. H., Keiper, R.: Selbsterregte Schwingungen bei Turbomaschinen in-
folge Labyrinthstromung. Ing. Archlv 43, (1974) 127-135
[61) Benckert, H., Wachter, J.: Studies on vibrations stimulated by lateral forces in
sealing gaps. AGARD Conf. Proc. No. 237, (1978) pp. (9-1)-(9-11)
[62) Benckert, H., Wachter, J.: Flow induced coefficients oflabyrinth seals for appli-
cation in rotordynamics. In: [12), (1980) pp. 189-212
[63) Benckert, H., Wachter, J.: Flow induced spring constants of labyrinth seals. In:
[176), (1980) pp. 53-63
[64) Iwatsubo, T.: Evaluation of instability forces of labyrinth seals in turbines or
compressors. In: [12), (1980) pp. 139-167
374 References

[65J Childs, D. W., Chang-Ho Kim: Testing for rotordynamic coefficients and leak-
age: Circumferentially grooved turbulent annular seals. In: [177J, (1986) pp.
609-618
[66J Scharrer, J. K.: Rotordynamic coefficients for stepped labyrinth gas seals. In:
[12J, (1988) pp. 177-195
[67J Dietzen, F. J., Nordmann, R.: Calculating rotordynamic coefficients of seals
by finite-difference technique. Trans. ASME, Journ. of Tribology 109, (1987)
388-394
[68J Dietzen, F. J., Nordmann, R.: A 3-dimensional finite-difference method for
calculating the dynamic coefficients of seals. In: [12J, (1988) pp. 211-227
[69J Nordmann, R., Dietzen, F. J.: Finite-difference analysis of rotordynamic seal
coefficients for an eccentric shaft position. In: [176J, (1988) pp. 379-386
[70J Weiser, H. P.: Ein Beitrag zur Berechnung der dynamischen Koeffizienten von
Labyrinthdichtungssystemen bei turbulenter Durchstromung mit kompressiblen
Medien. Techn. Univ. Kaiserslautem, Diss. (1989)
[71J Schmied, J.: Rotordynamic stability problems and solutions in high pressure
turbocompressors. In: [12J, (1988) pp. 395-413
[72J Kanki, H. et al.: High stability design for new centrifugal compressor. In: [12J,
(1988) pp. 445-459
[73J Kirk, R. G., Miller, W. H.: The influence of high pressure oil seals on turbo-
rotor stability. ASLE Trans. 22:1, (1979) 14-24
[74J Kirk, R. G., Nicholas, J. C.: Analysis of high pressure oil seals for optimum
turbocompressor dynamic performance. In: [176J, (1980) pp. 125-131
[75J Alford, J. S.: Protecting turbomachinery from self-excited rotor whirl. Trans.
ASME, Journ. for Power 87, (1965) 333-344
[76J Thomas, H. J., Urlichs, K., Wohlrab, R.: Liiuferinstabilitiit bei thermischen
Turbomaschinen infolge Spalterregung. VGB Kraftwerkstechnik 56:6, (1976)
377-383
[77J Urlichs, K.: Die Spaltstromung bei thermischen Turbomaschinen als Ursache
ftir die Entstehung schwingungsanfachender Querkriifte. Ing. Arch. 45:3, (1976)
193-208
[78J Pollmann, E., Schwerdtfeger, H., Termuehlen, H.: Flow excited vibrations in
high pressure turbines (steam whirl). Journ. of Eng. for Power 100, (1978) 180-
189
[79J Hauck, L.: Vergleich gebrauchlicher Turbinenschaufeln hinsichtlich des Auftre-
tens spaltstromungsbedingter Kriifte. Konstruktion 33:2, (1981) 59-64
[80J Thomas, H. J.: Thermische Kraftanlagen. 2. Aufi., Berlin ... : Springer (1985)
[81J Baumgartner, lid.: Berechnung von Querkriiften an Turborotoren verursacht
durch die Stromung in Labyrinthdichtungen. Techn. Univ. Miinchen, Diss.
(1989)
[82J Kramer, E.: Selbsterregte Schwingungen von Wellen infolge von Querkriiften.
Brennst.-Wirme-Kraft 20:7, (1968) 307-312
[83J Pollmann, E.: Stabilitiit einer in Gleitlagem rotierenden Welle mit Spalterre-
gung. Fortschr.-Ber. VDI 1:15, (1969)
[84J Wohlrab, R.: Einflufi der Lagerung auf die Laufstabilitat einfacher Rotoren mit
Spalterregung. Konstruktion 28, (1976) 473-478
References 375

[85) Thomas, H. J.: Zur Laufstabilitat einfacher Turborotoren, besonders bei Spalt-
erregung. Konstruktion 30:9, (1978) 339-344
[86) Thomas, H. J.: Zur Spalterregung bei Mehrlager-Wellensystemen. Konstruktion
40, (1988) 417-420
[87) Robertson, D.: Hysteretic influences on the whirling of rotors. Proc. of the Inst.
of Mech. Eng. 131, (1935) 513-531
[88) Lazan, B. J.: Damping of materials and members in structural mechanics. Ox-
ford: Pergamon Press (1968)
[89) Bert, C. W.: Material damping. Journ. of Sound and Vibr. 29, (1973) 129-153
[90) Harris, C., Crede, Ch. E.: Shock and vibration handbook. 2nd Ed., New York:
McGraw-Hill (1976)
[91) Nashif, A. D., Jones, D. I. G., Henderson, J. P.: Vibration damping. New York:
John Wiley & Sons (1985)
[92) Lund, J. W.: Destabilization of rotors from friction in internal joints with mi-
croslip. In: [177), (1986) pp. 487-491
[93) Tondl, A.: Some problems of rotor dynamics. London: Chapman & Hall (1965)
[94) Kellenberger, W.: Die Stabilitiit rotierender Wellen infolge innerer und iiufierer
Diimpfung. Ing. Archiv 32:5, (1963) 323-340
[95) Gunter, E. J., Trumpler, P. R.: The influence of internal friction on the stability
of high speed rotors with anisotropic supports. ASME Paper No. 69-Vibr. 2,
(1969) 1-13
[96) Shiraki, K., Umemura, S.: On the vibrations of two-rotor system connected by
gear couplings. Mitsubishi Techn. Rev. 22-33, (1970)
[97) Pedersen, P. T.: On self-excited whirl of rotors. Ing. Arch. 42, (1973) 267-284
[98) Muszynska, A.: Motion of a shaft with nonlinear elastic and internal damping
properties. In: Nonlinear Vibration Problems, Warszawa 17, (1976) pp. 189-224
[99) Williams, R. Jr., Trent, R.: The effects of nonlinear assymmetric supports on
turbine engine rotor stability. SAE Trans. 79, (1970) 1010-1020
[100) Morton, P. G.: Aspects of Coulomb damping in rotors supported on hydrody-
namic bearings. In: [12], (1982) pp. 45-57
[101] Marmol, R. A., Smalley, A. J., Tecza, J. A.: Spline coupling induced nonsyn-
chronous rotor vibrations. ASME Journ. of Mech. Design 102:1, (1980) 168-176
[102] Nataraj, C., Nelson, H. D., Arakere, N.: The effect of a Coulomb spline on
rotor dynamic response. In: Bently. Symp. on instability in rotating machinery,
Carson City (1985)
[103] Biihlmann, E. T., Luzi, A.: Rotor instability due to a gear coupling connected
to a bearingless sun wheel of a planetary gear. In: [12], (1988) pp. 19-39
[104) Prandtl, L.: Beitriige zur Frage der kritischen Drehzahlen. Dinglers Polytechn.
Journal 333, (1918) 179-182
[105] Smith, D. M.: The motion of a rotor carried by a flexible shaft in flexible
bearings. Proc. Roy. Soc. A142, (1933) 92-117
[106J Taylor, H. D.: Critical speed behavior of unsymmetrical shafts. Trans. ASME,
J. Appl. Mech., Vol. 62, A-71 (1940)
[107J Kellenberger, W.: Biegeschwingungen einer unrunden rotierenden Welle in ho-
rizontaler Lage. Ing. Arch. 26:4, (1958) 302-318
376 References

[108] Foote, W. R., Poritzky, H., Slade, J. J.: Critical speeds of a rotor with unequal
shaft flexibilities, mounted in bearings of unequal flexibility. Journ. of Appl.
Mech. 10, (1943) A77-A84
[109] Black, H. F.: Parametrically excited lateral vibrations of an asynunebic slender
shaft in asynunetrically flexible bearings. Joum. Mech. Eng. Sce. 11, (1969)
57-67
[110] Iwatsubo, T., Tomita, A., Kawai, R.: Vibrations of asymmetric rotors supported
by asymmetric bearings. Ing. Arch. 42, (1973) 416-432
[111] Wang, Z., Lund, J. W.: Calculation oflong rotors with many bearings on flexible
foundation. In: [176], (1984) pp. 13-16
[112] Wang, Z.: The singularity of the Riccati transfer matrix method and a method
for its being eliminated. In: [176], (1986) pp. 81-85
[113] Pestel, E. C., Leckie, F. A.: Matrix methods in elastomechanics. New York:
McGraw-Hill (1963)
[114] Marguerre, K., Uhrig, R.: Berechnung vielgliedriger Schwingerketten I. Das
Ubertragungsverfahren und seine Grenzen. ZAMM, 44:1, 2, (1964) 1-21
[115] Marguerre, K., Wolfel, H.: Mechanics of vibration. Alphen aan den Rijn: Sijthoff
& Noordhoff (1979)
[116] Guyan, R. J.: Reduction of stiffness and mass matrices. AIAA 3, (1965) 2
[117] Eckert, L.: Biegeschwingungen von Dampfturbogruppen - Vergleich zwischen
Berechnung und Messung. Fortschr. Ber. VDI 138:11, (1990)
[118] Gasch, R.: Berechnung der Lagedasten mehrfach gleitgelagerter Rotoren unter
Beriicksichtigung des nichtlinearen Verhaltens des Olfilms. Konstruktion 22:6,
(1970) 229-237
[119] Lord Rayleigh: On waves propagated along the plane surface of an elastic solid.
Proc. Math. Soc. 17, (1885) 4-11
[120] Ehlers, G.: Der Baugrund ais Federung in schwingenden Systemen. Beton und
Eisen 41, (1942) 197-203
[121] Meek, J. W., Velets, A. S.: Simple models for foundation in lateral and rocking
motion. In: 5th World congress on earth-quake. Rome (1973) pp. 2610-2613
[122] Schilling, W.: Einflufi des Baugrundes auf die Schwingungen von Turbomaschi-
nen. Techn. Univ. Darmstadt, Diss. (1986)
[123] Schollhom, K.: Koordinatenreduktion durch Anpassung kinetischer Nachgie-
bigkeiten. Techn. Univ. Darmstadt, Diss. (1984)
[124] Schirmer, K.-S.: Schwingungsberechnung eines Rotor-Fundamentsystems.
Techn. Univ. Darmstadt, Dipl.-Arb. (1987)
[125] Kramer, E.: Models for computation of turbomachine vibrations. In: ASME
Conf. Mechanical Vibration and Noise, Cincinnati, (1985) 85-DET-138
[126] Weinmann, U.: Vergleich der Schwingungen der Turbogruppe E3 bei Verwen-
dung verschiedener Fundamente. Techn. Univ. Darmstadt, Stud.-Arb. (1989)
[127] Weber, H.: Uber das gemeinsame Schwingungsverhalten von Welle und Funda-
ment bei Turbinenanlagen. VDI-Ber. 48, (1961) 55-62
[128] Pons, A.: Experimental and numerical analysis on a large nuclear steam turbo-
generator group. In: [177], (1986) pp. 269-275
[129] Gasch, R.: Vibration oflarge turbo-rotors in fluid film bearings on elastic foun-
dation. Journ. of Sound and Vibr. 47:1, (1976) 53-73
References 377

(130) Kriimer, E.: Computation of vibrations of the coupled system machine-founda-


tion. In: (176), (1980) pp. 333-338
(131) Vinsonneau, B., Lemant, F.: Characterizing the influence of the structure and
connection dampings on dynamical rotor behaviour. In: (177), (1990) pp. 367-
372
(132) Grunau, R.: Einflufi des Fundaments auf die Stabilitiit von Turbogruppen.
Techn. Univ. Darmstadt, Dipl.-Arb. (1984)
(133) Kriimer, E., Eckert, L.: Comparison between computed and measured vibrations
of turbomachines. In: (176), (1988) pp. 411-418
(134) Henry, T. A., Okah-Avae, B. E.: Vibrations in cracked shafts. In: (176), (1976)
C162/76
(135) Gasch, R.: Dynamic behaviour of a simple rotor with a cross-sectional crack.
In: (176), (1976) C178/76
(136) Mayes, I. W., Davies, W. G. R.: The vibrational behavior of a rotating shaft
system containing a transverse crack. In: (176), (1976) C168/76
(137) Ziebarth, H., Schwerdtfeger, H., Miihle, E. E.: Auswirkung von Querrissen auf
das Schwingungsverhalten von Rotoren. VDI-Berichte Nr. 320, (1978), 37-43
(138) Mayes, I. W., Davies, W. G. R.: A method of calculating the vibrational be-
havior of coupled rotating shafts containing a transverse crack. In: (176), (1980)
C254/80
(139) Inagaki, T., Kanki, H., Shiraki, K.: Transverse vibrations of a general cracked
rotor bearing system. ASME Publ., Paper 81-DET-45, Design engineering tech-
nical conference, Hartford, Connecticut, (September 1981)
(140) Sol, J. C.: Vibration detection of a transverse crack in a rotating machine shaft.
Canada Inst. for Scientific and Tech. Information, Ottawa, Ontario, Rept. No.
ISSN-0077-5606, NRD/CNR TT-2018 (1982)
(141) Inagaki, T., Kanki, H., Shiraki, K.: Transverse vibrations of a general cracked
rotor-bearing system. ASME, Journal of Mechanical Design. Vol. 104/345,
(April 1982)
(142) Grabowski, B.: Schwingungsberechnung eines angerissenen Turbinenliiufers.
VDI-Berichte Nr. 320, (1978) 31-36
(143) Grabowski, B., Poppel, R.: Das Schwingungsverhalten eines Rotors mit Querrifl
- experimentelle und theoretische Ergebnisse. VDI-Berichte Nr. 456, (1982)
167-175
(144) Baumgartner, R. J., Ziebarth, H.: Vibration monitoring criteria for early dis-
cernment of turbine rotor cracks. Presented at EPRI Workshop on incipient
failure detection, Hartford, Connecticut, (25-27 August 1982)
(145) Grabowski, B.: Shaft vibrations in turbomachinery excited by cracks. In: (12),
(1982)
(146) Muszynska, A.: Shaft crack detection. Seventh Machinery Dynamics Seminar,
Canada, (1982)
(147) Dimarogonas, A. D., Papadopoulos, C. A.: Vibration of cracked shafts in bend-
ing. Journ. of Sound and Vibr. 91 (4), (1983) 583-593
(148) Mayes, I. W., Davies, W. G. R.: Analysis of the response of a multi-rot or-
bearing system containing a transverse crack in a rotor. ASME, Journ. of Vi-
bration, acoustics, stress and reliability in design, Vol. 106, (January 1984)
378 References

[149J Davies, W. G. R., Mayes, I. W.: The vibrational behavior of a multi-shaft,


multi-bearing system in the presence of a propagating transverse crack. ASME,
Journ. of Vibration, acoustics, stress and reliability in design, Vol. 106, (January
1984)
[150J Schmied, J., Krii.mer, E.: Vibrational behavior of a rotor with a cross-sectional
crack. In: [176J, (1984) C 279/84
[151J Nataraj, C., Nelson, H. D.: The dynamics of a rotor system with a cracked
shaft. ASME Paper No. 85-DET-31, Cincinnati, (September 1985)
[152J Bently, D. E., Muszynska, A.: Detection of rotor cracks. In: Fifteenth Turbo-
machinery Symposium, (1986)
[153J Schmied, J.: Schwingungsverhalten von Rotoren mit angerissenem Wellenquer-
schnitt. VDI Fortschr. Ber. 11:77, (1986)
[154J Schwingungsiiberwachung von Turbosatzen - ein Weg zur Erkennung von
Wellenrissen. Allianz Ber. Nr. 24, Miinchen (1987)
[155J Leis, B. N., et al.: Fatigue and fracture analysis of two turbine shafts. ASME
paper 81-PVP-27 [Joint Conf., Denver/Colorado 1981J
[156J Coyle, M. B., Watson, S. J.: Fatigue strength of turbine shafts with shrunk-on
discs. Inst. Mech. Eng. Proc. (1963/64) 178 [Ref. in: Konstruktion 17 (1965) H.
6J
[157J Haas, H.: Grofischiiden durch Turbinen- oder Generatorliiufer, entstanden im
Bereich bis zur Schleuderdrehzahl. In: Der Maschinenschaden 50 (1977) H. 6,
195-200
[158J H8xtermann, E.: Erfahrungen mit Schaden in Form von Anrissen und Briichen
an Dampfturbinenwellen, Radscheiben und Generatorlaufern. In: VGB Techn.-
wiss. Ber. TW 107
[159J Greco, J. et al.: Cumberland steam plant - cracked IP rotor coupling of unit
2. In: Proceedings of the Americal Power Conference [Chicago 1978J 40 (1978)
502-514
[160J Passaleva, G., Pira, G.: Cracked shaft vibration sensitivity to steam temperature
variations. In: [177J, (1982) (not publ.)
[161 J Bently, D. E.: Anatomy of a shaft crack. In: Orbit 7 (1986) H. 1 [Bently /NevadaJ
[162J Stewart, A. T.: Fatigue cracking in the rotor of a 500 megawatt alternator.
CEGB Research (1978) October
[163J Fretting fatigue in generator rotors. NEI Parsons, CEGB (1979)
[164J Kudrjavtsev, I. V., Shokov, N. A.: Fracture analysis of large turbo-generators
(in Russian). In: Energiemaschinenbau (1980) H. 11
[165J Krii.mer, E., Haapala, E., Paavola, M.: Field experiences with cracked rotors.
In: Allianz-Berichte Nr. 24 (1987) Abschnitt 6
[166J Nilsson, L. R. K.: On the vibration behaviour of a cracked rotor. In: [177J,
(1982) pp. 515-524
[167J Jenkins, L. St.: Cracked shaft detection on a large vertical nuclear reactor cool-
ing pump. In: Bently. Symposium on Inst. in Rot. Mach., Carson City, Nevada
(1985)
[168J Lohberg, L., Ullrich, W., Gaffal, K.: Shafts of main cooling pumps; failure analy-
sis and remedies. Seminar Sicherheit und Verfiigbarkeit. Techn. Univ. Stuttgart,
MPA (1987)
References 379

[169] Bohanik, J. S., Thomas, R.: Reactor recirculation pump shaft crack. Orbit, 11:3,
(1990)
[170] Zurmiihl, R.: Matrizen, 4. Aufl.. Berlin, ... : Springer (1964)
[171] Bathe, K. J., Wilson, E. L.: Numerical methods in finite element analysis. En-
glewood Cliffs: Prentice-Hall (1976)
[172] Meirovitch, L.: Computational methods in structural dynamics. Alphen aan den
Rijn, Rockville: Sijthoff & Noordhof (1980)
[173] Kramer, E.: Shortened modal analysis as an approximation in structure dyna-
mics. In: Symp. Dynamics of machine foundations, Bucharest (1985)
[174] Kramer, E.: Approximative computation of unbalance vibrations of multi-
bearing rotors. In: [177], (1986) pp. 523-527
[175] Corr, R. B., Jennings, A.: A simultaneous iteration algorithm for symmetric
eigenvalue problems. Int. Journ. for Num. Meth. in Eng., 10, (1976) 633-647
[176] IMechE Conf.: Vibrations in rotating machinery. Cambridge 1976, Cambridge
1980, York 1984, Edinburgh 1988
[177] IFToMM Conf.: Rotordynamics. Rome 1982, Tokyo 1986, Lyon 1990
Index

Adjacent coordinates 216, 229 Decay constant 8


Angular velocity 21 Deformation work 177
Anisotropic bearing 30 Directional transformation 272
Argaud diagram 359 Displacement influence coefficients 207
Dynamic stiffness 353
Beam element 221,269 Dynamic viscosity 79
Bearing
- clearance 78, 115 Eccentricity
- pedestal model 294 - of journal 78
Bimodal method 362 - of mass 19
Breathing crack 330 - ratio 84
Eigenvalue 350, 351
Campbell diagram 59 - algorithm 368
Centrifugal moment 52, 67 Eigensolution 350, 351
Characteristic equation 8, 351 Eigenvector 350, 351
Circular cross-sections 191 - left 361
Clearance - right 361
- function 78 Elements
- ratio 87, 115 - beam 221
Complete solution 7, 350 - bearing support 227
Coordinates of models 217, 226 - flexible coupling 224
Coulomb damping 181 - rigid mass 217
Crack - ring-hub 219
- avoidance 346 - shaft seals 228
- breathing 330 - steam whirl 228
- closed 330 Elliptical path 234
- depth 330 Entrance loss coefficient 152
Critical speed 317 Equation of motion 7, 349
Cross coupling stiffness 99 - discussion of the 230

Damping Finite element method 211


- coefficient 7 Force influence coefficients 208
- modal 352 Forced vibration 11
- modulus 172 Foundation types 261
- moment 178
- of rolling-element bearings 140 General solution 7, 350
- ratio 10 Generalized
- work 176 - damping 354
382 Index

- force 354 Mode shapes 236


- mass 354 Moment of intertia
- stiffness 354 - polar, diametral 44, 69
Geometric damping 264 Momental ellipse 192
Global coordinates 271 Moments and products of intertia 267
Growth coefficient 101, 106 Myklestad Prohl method 211
Guyan reduction 228
Natural
Gyroscopic
- circular frequency 9
- elements 41
- vibration 9
- matrix 41, 219
- vibrations, general discussion of 233
Half power points 14 Node 206
Half value width 14 n-, 2n-, 3n- vibration 335
Harmonic
Oil-film
- displacement 9
- force, resultant 82
- excitation 11, 353
- thickness 78
Hermite polynomials 222
Hysteresis loop 176 One degree of freedom system 7
Optimal bearing damping 36
Influence coefficients
- displacement 207 Particular solution 7
Period time 9, 11
- force 208
Phase angle 12
Jeffcott rotor 19 Polygon procedure 17
Journal bearing Preferred coordinates 216
- damping coefficients 92 Principle coordinates 355
- stiffness coefficients 82, 92 Principle of virtual work 230
Journal eccentricity 78 Proportional damping 231, 354

Kinematic viscosity 153 Q-factor 12

Laval shaft 19 Radius of gyration 43


Left eigenvectors 361 Random excitation 17
Local coordinates 271 Reduction of coordinates 228
Logarithmic decrement 10 Relative damping 175
Loss factor 175 Resonance
Loss modulus 176 - curve 14
- factor 307
Magnification - sharpness 12
- factor 12 Resistance parameter 155
- function 12, 15 Restoring force 144
Master coordinates 229 Reynolds
Maxwells reciprocal theorem 39 - Equation 81
Modal - Number 153
- damping ratio 352, 355 Rigid body element 220,270,274
- matrix 354 Rotation number 235
- method 353 Rotor
Modal principal vibration 355 - short, spherical, long 65
Index 383

Secular equation 350 Steam whirl force 161


Second moment of intertia 192 Stiffness
Section 206 - coefficient 7
Segment clearance 115 - in line 135
Short bearing - lateral 135
- stiffness and damping coefficients 88 - of rolling element bearings 133
- theory 83
Simple natural frequency 10 Transfer function 353, 360
Skew-symmetric matrix 219 Transfer-matrix method 211
Sommerfeld Number 87
Unbalance 52, 66
Specified damping capacity 175
- excitation 14
Spring foundation 262
- moment 52
Stability 99
- border line 102 Viscous damping 7
- chart 103, 108
Startup curve 102, 107 Whirl
Superelement 216 - forward, backward 45, 234
State equation 361
Spri nger-Verlag
and the Environment

We at Springer-Verlag firmly believe that an


international science publisher has a special
obligation to the environment, and our corpo-
rate policies consistently reflect this conviction.

We also expect our busi-


ness partners - paper mills, printers, packag-
ing manufacturers, etc. - to commit themselves
to using environmentally friendly materials and
production processes.

The paper in this book is made from


low- or no-chlorine pulp and is acid free, in
conformance with international standards for
paper permanency.

Das könnte Ihnen auch gefallen